Você está na página 1de 7

Annals of Physics 355 (2015) 4854

Contents lists available at ScienceDirect

Annals of Physics
journal homepage: www.elsevier.com/locate/aop

On the KleinGordon oscillator subject to a


Coulomb-type potential
K. Bakke , C. Furtado
Departamento de Fsica, Universidade Federal da Paraba, Caixa Postal 5008, 58051-970, Joo Pessoa, PB,
Brazil

highlights

Interaction between the KleinGordon oscillator and a modified mass term.


Relativistic bound states for both attractive and repulsive Coulomb-type potentials.
Dependence of the KleinGordon oscillator frequency on the quantum numbers.
Relativistic analogue of a position-dependent mass system.

article

info

Article history:
Received 12 August 2014
Accepted 28 January 2015
Available online 7 February 2015
Keywords:
KleinGordon oscillator
Coulomb-type potential
Biconfluent Heun equation
Relativistic bound states

abstract
By introducing the scalar potential as modification in the mass
term of the KleinGordon equation, the influence of a Coulombtype potential on the KleinGordon oscillator is investigated. Relativistic bound states solutions are achieved to both attractive and
repulsive Coulomb-type potentials and the arising of a quantum
effect characterized by the dependence of angular frequency of the
KleinGordon oscillator on the quantum numbers of the system is
shown.
2015 Elsevier Inc. All rights reserved.

1. Introduction
In recent decades, the relativistic generalization of the harmonic oscillator has attracted a great
deal of attention. The most known relativistic model of the harmonic oscillator was introduced by
Moshinsky and Szczepaniak [1], which is known as the Dirac oscillator, and has been investigated

Corresponding author.
E-mail addresses: kbakke@fisica.ufpb.br (K. Bakke), furtado@fisica.ufpb.br (C. Furtado).

http://dx.doi.org/10.1016/j.aop.2015.01.028
0003-4916/ 2015 Elsevier Inc. All rights reserved.

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854

49

by several authors [211]. In particular, the Dirac oscillator has attracted a great interest in studies
of JaynesCummings model [8,4], the Ramsey-interferometry effect [9] and quantum phase transitions [10,11]. On the other hand, a relativistic model of the harmonic oscillator has also been proposed
for a scalar particle by Bruce and Minning [12] based on the Dirac oscillator [1]. Bruce and Minning [1]
showed that an analogous coupling to the linear Dirac oscillator coupling can be introduced into the
KleinGordon equation in such a way that one can recover the Schrdinger equation for a harmonic
oscillator in the nonrelativistic limit. This coupling proposed by Bruce and Minning is known as the
KleinGordon oscillator [1215]. For example, by considering the isotropic KleinGordon oscillator in
(2 + 1) dimensions, the KleinGordon equation becomes:

E 2 m2 = p + im p im ,

(1)

where m
is the rest mass of the scalar particle, is the angular frequency of the KleinGordon oscillator, = x2 + y2 and is a unit vector in the radial direction. In recent years, the KleinGordon oscillator has been investigated in noncommutative space [16,17], in noncommutative phase space [18]
and in P T -symmetric Hamiltonian [19].
Recently, several authors have shown their interest in investigating relativistic effects [2023] on
systems where the motion of a particle is governed by harmonic oscillations, such as the vibrational
spectrum of diatomic molecules [24], the binding of heavy quarks [25,26] and the oscillations of atoms
in crystal lattices, by mapping them as a position-dependent mass system [2730]. The importance of
these potentials arises from the presence of a strong potential field. In particular, Bahar and Yasuk [20]
dealt with the quarkantiquark interaction as a problem of a relativistic spin-0 particle possessing
a position-dependent mass, where the mass term acquires a contribution given by a interaction
potential that consists of a linear and a harmonic confining potential plus a Coulomb potential term.
It is worth mentioning other works that have explored the relativistic quantum dynamics of a scalar
particle subject to different confining potentials which can be in the interest of several areas of
physics [3136].
The aim of this work is to study the influence of a Coulomb-type potential on the KleinGordon
oscillator. In recent years, the confinement of a relativistic scalar particle to a Coulomb potential has
been discussed by several authors [3741]. As discussed in Ref. [41], the procedure in introducing
a scalar potential into the KleinGordon equation follows the same procedure in introducing the
electromagnetic 4-vector potential. This occurs by modifying the momentum operator p = i in the
form: p p q A (x). Another procedure
was proposed
in Ref. [42] by making a modification in
the mass term in the form: m m+S
r , t , where S
r , t is the scalar potential. This modification in
the mass term has been explored in recent decades, for instance, by analysing the behaviour of a Dirac
particle in the presence of static scalar potential and a Coulomb potential [43] and a relativistic scalar
particle in the cosmic string spacetime [44]. In this work, we investigate the influence of a Coulombtype potential on the KleinGordon oscillator by introducing the scalar potential as modification in
the mass term in the KleinGordon equation. We obtain bound state solutions to the KleinGordon
equation for both attractive and repulsive Coulomb-type potentials and show a quantum effect
characterized by the dependence of angular frequency of the KleinGordon oscillator on the quantum
numbers of the system, which means that not all values of the angular frequency are allowed.
The structure of this paper is as follows: in Section 2, we study the KleinGordon oscillator subject
to a Coulomb-type potential in the Minkowski spacetime in (2 + 1) dimensions; in Section 3, we
present our conclusions.
2. KleinGordon oscillator under the influence of a Coulomb-type potential
In this section, we study the behaviour of the KleinGordon oscillator subject to a Coulomb-type
potential in (2+1) dimensions. We consider the cylindrical symmetry, then, we write the line element
of the Minkowski spacetime in the form (with c = h = 1):
ds2 = dt 2 + d 2 + 2 d 2 .

(2)

50

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854

Thereby, the KleinGordon equation describing the interaction of the KleinGordon oscillator and
static scalar potential is given by (with c = h = 1)
[m + S ()]2 =

2
p + im p im ,
2
t

(3)

where S () is a scalar potential, m is the restmass of the scalar particle, is the angular frequency of
, = x2 + y2 and is a unit vector in the radial direction. In
the KleinGordon oscillator, p = i
this way, by considering a Coulomb-type potential
S () =

|f |
,

(4)

where f is a constant, the KleinGordon equation (3) becomes

2
2
1
1 2
2mf
f2
+
+
+ 2 2 + m m2 2 2 m2
2 = 0.
2
2
t

(5)

In what follows, we shall be considering particular stationary solutions to Eq. (5) that are eigenfunctions of the operator L z = i . In this way, we have

= eiE t eil R () ,

(6)

where l = 0, 1, 2, . . .. Then, substituting (6) into Eq. (5), we obtain


d2 R
d 2

1 dR

2
2mf
R
R m2 2 2 R + 2 R = 0,
2

(7)

where we have defined the following parameters in Eq. (7):

2 = E 2 m2 + m;

(8)

2 = l2 + f 2 .
From now on, let us consider =
d2 R
d

1 dR

m , thus, we rewrite the radial equation (7) in the form:

2
R R 2 R +
R = 0,
2

(9)

where we have defined a new parameter


2mf

(10)

Let us discuss the asymptotic behaviour of the possible solutions to Eq. (9), which are determined
for 0 and . From Refs. [4446], the behaviour of the possible solutions to Eq. (9) at 0
and allows us to write the function R ( ) in terms of an unknown function H ( ) as it follows:

R ( ) = exp

2
2

| | H ( ) ,

(11)

where = 1. Note that one should have in mind the possibility of existing a singular solution to
Eq. (9). As pointed out in Refs. [45,47], = 1 yields a non-singular solution whereas = 1 yields
a singular solution to Eq. (9). In this contribution, we consider only non-singular solutions to Eq. (9),
that is, = 1. Then, substituting the radial wave function given in Eq. (11) into Eq. (9), we obtain

1
dH

|
|
|
|
+
2
+
1

H = 0.
(
)
d 2

d
m

d2 H

(12)

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854

51

The second order differential equation (12) corresponds to the Heun biconfluent equation [48,44,
49,50] and the function H ( ) is the Heun biconfluent function
H ( ) = H

2 | | , 0,

2
, 2, .
m

(13)

Proceeding with our discussion about bound states solutions, let us use the Frobenius method
[51,52]. Thereby, the solution to Eq. (12) can be written as a power series expansion around the origin:
H ( ) =

aj j .

(14)

j=0

Substituting the series (14) into (12), we obtain the following recurrence relation:
aj+2 =

( 2j)
aj + 1
aj ,
(j + 2) (j + 1 + )
(j + 2) (j + 1 + )

(15)

where = 2 | | + 1 and = m 2 2 | |. By starting with a0 = 1 and using the relation (15), we


can calculate other coefficients of the power series expansion (14). For instance,
a1 =

2mf

=
a2 =

m (2 | | + 1)

;
(16)

2 (1 + )
2 (1 + )
2mf 2

(2 | | + 1) (2 | | + 2)

.
2 (2 | | + 2)

It is well-known that the quantum theory requires that the wave function (6) must be normalizable.
Therefore, we assume that the function R ( ) vanishes at 0 and . This means that we
have a finite wave function everywhere, that is, there is no divergence of the wave function at 0
and , then, bound state solutions can be obtained. However, we have written the function
H ( ) as a power series expansion around the origin in Eq. (14). Thereby, bound state solutions can be
achieved by imposing that the power series expansion (14) or the Heun biconfluent series becomes a
polynomial of degree n. In this way, we guarantee that R ( ) behaves as | | at the origin and vanishes
at [46,45]. Through the recurrence relation (15), we can see that the power series expansion
(14) becomes a polynomial of degree n by imposing two conditions [5254,49,44,50,46,45]:

= 2n and an+1 = 0,

(17)

where n = 1, 2, 3, . . .. From the condition = 2n, we can obtain:

En2, l

= m + 2m n, l n + | | +
2

1
2

(18)

Hence, Eq. (18) is a general expression for the relativistic energy levels of the KleinGordon oscillator subject to a Coulomb-type potential, where we have coupled this scalar potential as a modification of the mass term in the KleinGordon equation. In contrast to the results of Ref. [12], we
have that the influence of the Coulomb-like potential makes that the ground state to be defined by
the quantum number n = 1 instead of the quantum number n = 0. Note that we have written the
angular frequency in terms of the quantum numbers {n, l} in Eq. (18). From the mathematical point
of view, this dependence of the angular frequency of this relativistic oscillator on the quantum numbers {n, l} results from the fact that the exact solutions to Eq. (12) are achieved for some values of
the KleinGordon oscillator frequency. From the quantum mechanics point of view, this is an effect
which arises from the influence of the Coulomb-type potential on the KleinGordon oscillator.

52

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854

Henceforth, let us discuss this behaviour of the KleinGordon oscillator frequency. First of all, we
must note that, at first glance, the result obtained in Eq. (18) shows that the relativistic energy levels do not depend on the parameter which gives rise to the Coulomb-type potential (defined in
Eq. (10)). However, we need to analyse the second condition established in Eq. (17), that is, the condition an+1 = 0. By analysing this second condition, we obtain an expression involving specific values
of the angular frequency of the KleinGordon oscillator and the parameter . It is worth mentioning
other analyses which have been made in recent years, for example, a relation involving the harmonic
oscillator frequency, the Lorentz symmetry-breaking parameters and the total angular momentum
quantum number obtained in Ref. [53], a relation involving a coupling constant of a Coulomb-like
potential, the cyclotron frequency and the total angular momentum quantum number in semiconductors threaded by a dislocation density obtained in Ref. [49], and a relation involving the mass of
a relativistic particle, a scalar potential coupling constant and the total angular momentum quantum
number achieved in Ref. [44].
Thereby, let us assume that the angular frequency of the KleinGordon oscillator can be adjusted
in such a way that the condition an+1 = 0 is satisfied in order that a relation between the angular
frequency of the KleinGordon oscillator and the Coulomb-type potential parameter can be obtained.
The meaning of achieving this relation is that not all values of the angular frequency are allowed,
but some specific values of which depend on the quantum numbers {n, l}. For this reason, we label

= n, l .

(19)

Therefore, the conditions established in Eq. (17) are satisfied and a polynomial solution to the function
H ( ) given in Eq. (14) is achieved [46,45,44,50]. As an example, let us consider n = 1, which
corresponds to the ground state, and analyse the condition an+1 = 0. For n = 1, we have a2 = 0.
The condition a2 = 0, thus, yields

1, l =

2mf 2

(2 | | + 1)

(20)

In this way, substituting Eq. (20) into Eq. (18), the expression of the energy level (for n = 1) becomes:

E1, l = m2 + 2m1, l n + | | +

1/2
| | + 23
2
= m 1 + 4f
.
(2 | | + 1)

(21)

In what follows, let us consider the simplest case of the function H ( ) which corresponds to a
polynomial of first degree. In this way, for n = 1, we can write
H1, l ( ) = 1 +

(22)

Thereby, from Eq. (22), thus, Eq. (11) can be written in the form: R1, l ( ) = e /2 | | 1 + .
Hence, the effects of the Coulomb-type potential introduced by a coupling with the mass term
on the spectrum of energy of the KleinGordon oscillator correspond to a change of the energy levels, where the ground state is defined by the quantum number n = 1. Moreover, a quantum effect
characterized by the dependence of angular frequency of the KleinGordon oscillator on the quantum
numbers {n, l} of the system arises, whose meaning is that not all values of the angular frequency are
allowed. In this way, the conditions established in Eq. (17) are satisfied and a polynomial solution to
the function H ( ) given in Eq. (14) is achieved in agreement with Refs. [46,45,44,50].
2

3. Conclusions
We have studied the influence of a Coulomb-type potential on the KleinGordon oscillator and
shown that the ground state of the KleinGordon oscillator is defined by the quantum number

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854

53

n = 1 instead of the quantum number n = 0. Other quantum effect obtained is the dependence
of angular frequency of the KleinGordon oscillator on the quantum numbers {n, l} of the system,
whose meaning is that not all values of the angular frequency are allowed. As an example, we have
calculated the angular frequency of the ground state n = 1 and obtained the expression of the energy
level of the ground state.
It is worth mentioning that we have introduced the scalar potential as a modification of the mass
term of the KleinGordon equation. However, as discussed in Ref. [41], one can use the same procedure
of introducing the electromagnetic 4-vector potential by modifying the momentum operator as p
p q A (x). By dealing with a Coulomb-type potential via a minimal coupling, new interesting
results associated with the KleinGordon oscillator can be obtained. Moreover, the interaction
between a relativistic scalar particle and the KleinGordon oscillator can be of interest in studies of
the quarkantiquark interaction [20], position-dependent mass systems [2730], the Casimir effect
[55,56] and the KaluzaKlein theory [57].
Acknowledgment
The authors would like to thank CNPq (Conselho Nacional de Desenvolvimento Cientfico e
Tecnolgico Brazil) (300860/2013-7) for financial support.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]

M. Moshinsky, A. Szczepaniak, J. Phys. A: Math. Gen. 22 (1989) L817.


V.M. Villalba, Phys. Rev. A 49 (1994) 586.
D. It, K. Mori, E. Carrieri, Nuovo Cimento A 51 (1967) 1119.
P. Rozmej, R. Arvieu, J. Phys. A 32 (1999) 5367.
K. Bakke, Gen. Relativity Gravitation 45 (2013) 1847.
K. Bakke, C. Furtado, Ann. Phys. (NY) 336 (2013) 489.
J. Carvalho, C. Furtado, F. Moraes, Phys. Rev. A 84 (2011) 032109.
A. Bermudez, M.A. Martin-Delgado, E. Solano, Phys. Rev. A 76 (2007) 041801(R).
A. Bermudez, M.A. Martin-Delgado, A. Luis, Phys. Rev. A 77 (2008) 033832.
A. Bermudez, M.A. Martin-Delgado, A. Luis, Phys. Rev. A 77 (2008) 063815.
A. Bermudez, M.A. Martin-Delgado, E. Solano, Phys. Rev. Lett. 99 (2007) 123602.
S. Bruce, P. Minning, Nuovo Cimento A 106 (1993) 711.
V.V. Dvoeglazov, Nuovo Cimento A 107 (1994) 1413.
A. Boumali, A. Hafdallah, A. Toumi, Phys. Scr. 84 (2011) 037001.
N.A. Rao, B.A. Kagali, Phys. Scr. 77 (2008) 015003.
B. Mirza, R. Narimani, S. Zare, Commun. Theor. Phys. (Beijing) 55 (2011) 405;
B. Mirza, M. Mohadesi, Commun. Theor. Phys. (Beijing) 42 (2004) 664.
M.-L. Liang, R.-L. Yang, Internat. J. Modern Phys. A 27 (2012) 1250047.
Y. Xiao, Z. Long, S. Cai, Internat. J. Theoret. Phys. 50 (2011) 3105.
J.-Y. Cheng, Internat. J. Theoret. Phys. 50 (2011) 228.
M.K. Bahar, F. Yasuk, Advances in High Energy Physics, Vol. 2013, 2013, p. 6. http://dx.doi.org/10.1155/2013/814985.
Article ID 814985.
R. Kumar, F. Chand, Phys. Scr. 85 (2012) 055008.
S.H. Dong, Internat. J. Theoret. Phys. 39 (2000) 1119.
S.H. Dong, Internat. J. Theoret. Phys. 40 (2001) 559.
S.M. Ikhdair, Internat. J. Modern Phys. C 20 (2009) 1563.
C. Quigg, J.L. Rosner, Phys. Rep. 56 (1979) 167.
M. Chaichian, R. Kgerler, Ann. Phys. (NY) 124 (1980) 61.
L. Dekar, L. Chetouani, T.F. Hammann, Phys. Rev. A 59 (1999) 107.
A.D. Alhaidari, Phys. Rev. A 66 (2002) 042116.
A.D. Alhaidari, Phys. Lett. A 322 (2004) 72.
L. Serra, E. Lipparini, Europhys. Lett. 40 (1997) 667.
A. de Souza Dutra, C.-S. Jia, Phys. Lett. A 352 (2006) 484.
W.-C. Qiang, R.-S. Zhou, Y. Gao, Phys. Lett. A 371 (2007) 201.
A.S. de Castro, Phys. Lett. A 338 (2005) 81.
A.D. Alhaidari, H. Bahlouli, A. Al-Hasan, Phys. Lett. A 349 (2006) 87.
F. Domingues-Adame, Phys. Lett. A 136 (1989) 175.
Y. Xu, S. He, C.-S. Jia, Phys. Scr. 81 (2010) 045001.
Q. Wen-Chao, Chin. Phys. 12 (2003) 1054.
H. Motavalli, A.R. Akbarieh, Modern Phys. Lett. A 25 (2010) 2523.
F. Yasuk, A. Durmus, I. Boztosun, J. Math. Phys. 47 (2006) 082302.
A.L. Cavalcanti de Oliveira, E.R. Bezerra de Mello, Classical Quantum Gravity 23 (2006) 5249.
W. Greiner, Relativistic Quantum Mechanics: Wave Equations, third ed., Springer, Berlin, 2000.

54
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]

K. Bakke, C. Furtado / Annals of Physics 355 (2015) 4854


H.G. Dosch, J.H. Jansen, V.F. Mller, Phys. Nor. 5 (1971) 2.
G. Soff, B. Mller, J. Rafelski, W. Greiner, Z. Naturf. a 28 (1973) 1389.
E.R. Figueiredo Medeiros, E.R. Bezerra de Mello, Eur. Phys. J. C 72 (2012) 2051.
J. Myrhein, E. Halvorsen, A. Verin, Phys. Lett. B 278 (1992) 171.
A. Verin, Phys. Lett. B 260 (1991) 120.
J. Grundberg, T.H. Hansson, A. Karlhede, J.M. Leinaas, Modern Phys. Lett. B 5 (1991) 539.
A. Ronveaux (Ed.), Heuns Differential Equations, Oxford University Press, Oxford, 1995.
K. Bakke, F. Moraes, Phys. Lett. A 376 (2012) 2838.
K. Bakke, Ann. Phys. (NY) 341 (2014) 86.
G.B. Arfken, H.J. Weber, Mathematical Methods for Physicists, sixth ed., Elsevier Academic Press, New York, 2005.
C. Furtado, B.G.C. da Cunha, F. Moraes, E.R. Bezerra de Mello, V.B. Bezerra, Phys. Lett. A 195 (1994) 90.
K. Bakke, H. Belich, Eur. Phys. J. Plus 127 (2012) 102.
K. Bakke, H. Belich, Ann. Phys. (NY) 333 (2013) 272.
N.R. Khusnutdinov, M. Bordag, Phys. Rev. D 59 (1999) 064017.
H.F. Mota, K. Bakke, Phys. Rev. D 89 (2014) 027702.
C. Furtado, F. Moraes, V.B. Bezerra, Phys. Rev. D 59 (1999) 107504.

Você também pode gostar