Sunteți pe pagina 1din 13

Gen Relativ Gravit (2013) 45:18471859

DOI 10.1007/s10714-013-1561-6
RESEARCH ARTICLE

Rotating effects on the Dirac oscillator in the cosmic


string spacetime
K. Bakke

Received: 23 April 2013 / Accepted: 21 June 2013 / Published online: 30 June 2013
Springer Science+Business Media New York 2013

Abstract In this contribution, we study the Dirac oscillator under the influence of
noninertial effects of a rotating frame in the cosmic string spacetime. We show that
both noninertial effects and the topology of the cosmic string spacetime restrict the
physical region of the spacetime where the quantum particle can be placed, and discuss
two different cases of bound states solutions of the Dirac equation by analysing the
behaviour of the Dirac oscillator frequency.
Keywords Dirac oscillator Rotating frame Hard-Wall confining potential
Noninertial effects Cosmic string Relativistic bound states solutions
1 Introduction
In recent decades, the Dirac oscillator [13] has attracted a great interest in studies of
the JaynesCummings model [4,5], the Ramsey-interferometry effect [6] and quantum
phase transitions [7,8]. In Ref. [5], an exact mapping of the Dirac oscillator onto the
JaynesCummings model [9] has achieved a description of the interaction of a twolevel atom with a quantized single-mode field. Besides, the results of Ref. [4] have
shown that the entanglement of the spin with the orbital motion is produced in a way
similar to what is observed in the model of the JaynesCummings model when the
limit the strong spin-orbit coupling of the Dirac oscillator is considered.
The Dirac oscillator was proposed in Ref. [1] with the aim of recovering the harmonic oscillator Hamiltonian from the nonrelativistic limit of the Dirac equation. In
short, by considering the nonrelativistic limit of the Dirac equation describing the

K. Bakke (B)
Departamento de Fsica, Universidade Federal da Paraba, Caixa Postal 5008,
Joo Pessoa, PB 58051-970, Brazil
e-mail: kbakke@fisica.ufpb.br

123

1848

K. Bakke

interaction between a Dirac particle with the harmonic oscillator potential, the impossibility of recovering the harmonic oscillator Hamiltonian occurs due to the presence
of a quadratic potential [13,10] in the second order differential equation. Therefore,
in order to solve this problem, the relativistic harmonic oscillator has been investigated recently by introducing scalar and vector potentials which are quadratics in
coordinates [11] and by introducing a new coupling into the Dirac equation [1]. In
particular, the new coupling proposed in Ref. [1] is introduced in such a way that the
Dirac equation remains linear in both spatial coordinates and momenta, and recover
the Schrdinger equation for a harmonic oscillator in the nonrelativistic limit of the
Dirac equation. The Dirac oscillator has also been analyzed in a series of physical systems, such as in a thermal bath [12], in the presence of an external magnetic field [13],
in the point of view of the Lie algebra [14], in studies of covariance properties [15], in
the hidden supersymmetry [1518], by using the shape invariant method [19], in the
presence of the AharonovBohm quantum flux [2023], conformal invariance properties [24], in the FermiWalker reference frame [25] and for a system of a charged
particle interacting with a topological defect [26]. Recently, a model of a quantum
ring based on the Dirac oscillator has been proposed to confine relativistic spin-1/2
particles [27].
The aim of this paper is the study the Dirac oscillator under the influence of noninertial effects of a rotating frame in the cosmic string spacetime. Studies of noninertial effects have discovered interesting quantum effects such as the Sagnac effect
[2830], the Mashhoon effect [31], persistent currents in quantum rings [32], the coupling between the angular momentum and the angular velocity of the rotating frame
[3335] and spin currents [36]. Other works that worth mentioning are studies involving rotational and gravitational effects in quantum interference [3739], Dirac fields
[40], scalar fields [41,42], the influence of Lorentz transformations [43], the weak
field approximation [44], the LandauHeMcKellarWilkens quantization [45] and
confinement of a neutral particle to a quantum dot [4648]. In this work, we show
that both noninertial effects and the topology of the cosmic string spacetime restrict
the physical region of the spacetime where the quantum particle can be placed. By
analysing the behaviour of the oscillator frequency, we discuss two different cases of
bound states solutions of the Dirac equation.
The structure of this paper is: in Sect. 2, we introduce the Dirac oscillator [13]
and discuss the influence of the noninertial effects of a rotating frame on the Dirac
oscillator in the cosmic string spacetime. We show the restriction imposed by both
noninertial effects and the topology of the cosmic string spacetime on the physical
region of the spacetime where the quantum particle can be placed and analyse two
different cases of bound states solutions of the Dirac equation; in Sect. 3, we present
our conclusions.

2 Dirac oscillator in a rotating frame in a topological defect spacetime


We start this section by introducing the Dirac oscillator [13]. In the following, we
introduce the classical background of this work, build a rotating frame and discuss the
influence of noninertial effects of a rotating frame and the topology of the classical

123

Rotating effects on the Dirac oscillator

1849

background on the Dirac oscillator. Finally, we analyse the behaviour of the oscillator frequency by showing two different cases of bound states solutions of the Dirac
equation.
As we have pointed out in the introduction, the term Dirac oscillator was denominated by Moshinsky and Szczepaniak [1] based on the introduction of a coupling
in such a way that the Dirac equation remains linear in both spatial coordinates and
momenta. This new coupling introduced into the Dirac equation is given by
p p im ,

(1)

where m is the mass of the Dirac neutralparticle, is the oscillator frequency, is


one of the standard Dirac matrices, = x 2 + y 2 is the radial coordinate and is a
unit vector in the radial direction.
Let us introduce the scenario of general relativity which corresponds to the cosmic
string spacetime. The cosmic string spacetime is described by a line element given by
[49,50] (we work with the units h = c = 1)
ds 2 = dT 2 + d R 2 + 2 R 2 d2 + d Z 2 ,

(2)

where the parameter is related to the deficit angle which is defined as = 1


4 , with  being the linear mass density of the cosmic string. In the cosmic string
spacetime, the parameter can assume only values in which < 1. All values > 1
would correspond to a spacetime with negative curvature which does not make sense
in the general relativity context [5153]. The azimuthal angle varies in the interval:
0 < 2 . Besides, the geometry described by the line element (2) possesses a
conical singularity represented by the curvature tensor
,
=
R,

1
2 (
r ),
4

(3)

r ) is the two-dimensional delta function, and we have considered the units


where 2 (
h = c = 1. The behaviour of the curvature tensor (3) is denominated as a conical
singularity [54] which gives rise to the curvature concentrated on the cosmic string axis
with a null curvature in all other points of the spacetime. In recent decades, a great deal
of works have discussed the influence of the topology of the cosmic string spacetime,
for instance, in the problem of string sources in (2 + 1)-dimensional gravity [55], in
the KaluzaKlein theory [56], in the gravitational analogue of the AharonovBohm
effect [57], in the FermiWalker transport [38], in the relativistic quantum scattering
[58], in the Lorentz symmetry violation background [59] and the AharonovBohm
effect for bound states [60].
Now, let us make a coordinate transformation: T = t, R = ,  = + t and
Z = z, where is the constant angular velocity of the rotating frame. Thereby, the
line element (2) becomes


ds 2 = 1 2 2 2 dt 2 + 22 2 d dt + d 2 + 2 2 d 2 + dz 2 .

(4)

123

1850

K. Bakke

We can note that the line element (4) is defined for values of the radial coordinate
inside the range:
0 < < 1/.

(5)

Therefore, all values for the radial coordinate given by > 1/ mean that the
particle is placed outside of the light-cone because the velocity of the particle is
greater than the velocity of the light [61]. In this way, the range (5) imposes a spatial
constraint where the wave function must be defined. This restriction on the radial
coordinate imposed by both noninertial effects and the topology of the cosmic string
spacetime gives rise to a hard-wall confining potential, that is, it imposes that the wave
function of the Dirac particle must vanish at 1/.
Further, in order to work with Dirac spinors in a rotating frame in a topological
defect spacetime having a non-null curvature concentrated on the symmetry axis,
we use the formulation of spinors in curved spacetime [62,63]. In a curved spacetime background, spinors are defined locally. Each spinor transforms according to
the infinitesimal Lorentz transformations, that is,  (x) = D ( (x)) (x), where
D ( (x)) corresponds to the spinor representation of the infinitesimal Lorentz group
and (x) corresponds to the local Lorentz transformations [62]. Locally, the reference frame of the observers can be build via a noncoordinate basis a = ea (x) d x ,
where the components ea (x) are called tetrads and satisfy the relation [6264]:
g (x) = ea (x) eb (x) ab , where ab = diag( + ++) is the Minkowski ten
sor. The inverse of tetrads is defined as d x = e a (x) a , where we have the relations:

a
a
a
e (x) e b (x) = b and e a (x) e (x) = . Observe that the indices , . . . indicate the spacetime indices, while the indices a, b, c = 0, 1, 2, 3 indicate the local
reference frame. From this perspective, let us consider the following local reference
frame for the observers [65,66]:

ea (x)

1 2 0

22

0
0

1
0

1 2

0
; e (x) =
a

1 2

1 2

0
0
0

1 2

1 0
0
,

1 2

0
0

0
0
1
(6)

where we call = in (6).


The Dirac equation is defined in a curved spacetime background by considering the
covariant derivative of a spinor, whose components are defined as = +  (x),
where is the partial derivative and  (x) = 2i ab (x)  ab is the spinorial connec

tion [64,63]. The term  ab of the spinorial connection is defined as  ab = 2i a , b .
The a matrices correspond to the standard Dirac matrices in the Minkowski spacetime [67]:
0 = =

123

I 0
0 I


;

i = i =

0 i
i 0


;

i =

i 0
0 i


,

(7)

Rotating effects on the Dirac oscillator

1851

 and I being the spin vector and the 2 2 identity matrix, respectively. The
with 


matrices i are the Pauli matrices and satisfy the relation i j + j i = 2i j . The

matrices are related to the a matrices via = e a (x) a [63]. Moreover, the term
ab (x) of the spinorial connection is called the spin connection and can be obtained
a b(x) d x b = 0 in the
by solving the MaurerCartan structure equations d a +
absence of torsion [64]. By solving the MaurerCartan structure equations, we obtain
the following components of the spinorial connection [65,66]:

1 2 2 1 i


3;
2
2 1
2 1 2

1
 2 ;
 = 
2 1 2
t =

(8)

1 2
i
 = 
1 
3.
2
2 1
2 1 2
Hence, by introducing the Dirac oscillator coupling (1) into the Dirac equation in
curved spacetime background, we have [25]
i + i  (x) + i m0 0 = m.

(9)

By using the tetrads given in Eq. (6) and the components of the spinorial connection
given in Eq. (8), we can write the Dirac equation (9) in the form [66]:



1
2
1
0
+ i 
+ i +
+ m0
m = 

t
2
1 2
1 2 t
i

2
2

0
 .


+
 
i 2 
+i 3

2
2

z
2
1

2
1
1
(10)
 in the last term of Eq. (10) which is related
Observe the presence of the term
 
to the rotation-spin coupling as pointed in Ref. [35].
From now on, our goal is to solve the Dirac equation (10), then, we rewrite it in the
form:





1
2
0
1
0
2
2
+ i
= m 1 i 1 +
+ m0
i
t
t
2

2

i
+ i2 2
i 1 2 3

z

 
1
 .
+ 
(11)
2 1 2
The solution of Eq. (11) is given in the form: = ei E t ( ), where and
are two-spinors. Substituting this solution into (11), we obtain two coupled equations

123

1852

K. Bakke

for and , where the first coupled equation is









1
1
3
1
2
2
+
= i
1
E m 1 
m0
2 1 2
2
2

+ i2 2

i 1 2 3
E 2 ,
z
i

(12)
while the second coupled equation is







1
1
3
2
+ m0
E +m 1 
= i 1 1 2 +
2 1 2
2
2

+ i2 2

i 1 2 3
E 2 .
z
i

(13)
In order to solve the two coupled Eqs. (12) and (14), we consider the velocity of
rotation being small
 compared with the velocity of the light, that is, 1. In this
way, we can write 1 2 1 21 2 + . After some calculations, we obtain the
following second order differential equation:


3 2
2
1
1
i 3
2
E

+ 2+
m0
m = 2
2

E 3
+i2 3

2m0 3
+m 2 02 2 + i

+ 2 m0 2 3

2
1 2
2

2
2 2
+ 2 2 2 + i 2E
2

z
 



+ O 2 + O m 2 .
(14)
i 2m0 2
z
Note in Eq. (14) that is an eigenfunction of 3 , whose eigenvalues are s = 1.
Therefore, we label 3 s = s = ss , where s = (+ , )T . Furthermore, we
can see that the right-hand side of Eq. (14) commutes with the operators Jz = i
[68] and p z = iz , then, we can take the solutions of Eq. (14) in the form:
s = e

123



i l+ 21 ikz


R+ ()
,
R ()

(15)

Rotating effects on the Dirac oscillator

1853

where l = 0, 1, 2, . . . , k is a constant. Let us consider k = 0 in order to work with


a planar system. Thus, substituting (15) into (14), we obtain two noncoupled equations
for R+ () and R () given by
s2
d 2 Rs
1 d Rs
+

Rs
d 2
d
2 2


s
+ 1 Rs 2 s2 Rs
2m0 s

[E + (l + 1/2)]2 Rs m 2 Rs =

2s 2 s Rs 2 2 Rs + m 2 02 s2 2 Rs


 
(16)
+O 2 + O m 2 ,
where we have defined in (16) the parameter s and the effective angular momentum
s as

s = 1 + 2s
1
s
s = l + (1 s) + (1 ).
(17)
2
2
By making a change of variables given by = m 0 s 2 , we can rewrite the radial
Eq. (16) in the form:

d Rs
2
d 2 Rs

s
+
s2 Rs Rs +
Rs = 0,
d
d
4
4
4m0 s

(18)

where

s = [E + (l + 1/2)] m + 2m0
2


s
s
+ 1 + 2 s2 + 2 2 + 2s 2 s .

(19)

A regular solution at the origin is achieved if the solution of the Eq. (18) has the

form Rs ( ) = e 2

|s |
2

Fs ( ). Thus, substituting this solution into (18), we obtain





|s | 1
|s |
d Fs
s
d 2 Fs
+
1

Fs = 0,
+

d 2

d
4m0 s
2
2

(20)

which is the confluent hypergeometric equation or the Kummer equation [69]. The
solution of Eq. (20) regular at the 
origin is called the Kummer function
of first kind,

|s |
|s |
s
1
which is given by Fs ( ) = 1 F1 2 + 2 4m0 s , + 1; . It is well known
in the literature that the radial part of the wave function becomes finite everywhere
s
of the confluent hypergeometric function
when the parameter A = |2s | + 21 4m
0 s
is equal to a nonpositive integer number, which makes the confluent hypergeometric
series to be a polynomial of degree n [70]. Thereby, in order to have a wave function

being normalized inside the range 0 < < 1/, we assume that m0 .

123

1854

K. Bakke

This assumption makes the amplitude of probability being very small for values where
> 1/, because we have that = m0 s 2 1 when 1/. Therefore,
without loss of generality, we consider the wave function being normalized in the
range 0 < < 1/, since Rs ( ) 0 when 1/. In this way, the relativistic
energy levels are


En, l =





|s |
s
1
+
2m0 s
+ 1 2 s2 2 2 2s 2 s
m 2 + 4m0 s n +
2
2

[l + 1/2] .

(21)

where n = 0, 1, 2, . . . and s and s are given in Eq. (17). The relativistic energy levels
(21) correspond to the spectrum of energy of the Dirac oscillator under the influence of
noninertial effects and the topology of the cosmic string spacetime which are obtained

by imposing the condition m0 .


Moreover, we have that the effects of curvature break the degeneracy of the relativistic energy levels of the Dirac oscillator (21). By taking the limit 0, we have
the spectrum of energy of the Dirac oscillator in the Minkowski spacetime under the
influence of noninertial effects of the rotating frame (6). Recently, we have studied
the influence of the noninertial effects of the FermiWalker reference frame on the
Dirac oscillator in the cosmic string spacetime [25]. By comparing the results obtained
in Ref. [25] with the relativistic energy levels (21), we have that new contributions
to the energy levels arise from the noninertial effects of the rotating frame (6). This
stems from the fact that the local reference frameof the observers (6) is not defined
in the rest frame of the observers at each instant 0 = e0t (x) dt in contrast to the
FermiWalker reference frame [43,71]. Finally, we have that in last term of Eq. (21)
the coupling between the quantum number l and the angular velocity .
Let us discuss the nonrelativistic limit of the energy levels (21). The nonrelativistic
limit of Eq. (21) can be obtained by applying the Taylor expansion up to the first order
terms. In this way, the nonrelativistic limit of the energy levels (21) are

|s |
En l m + 20 s n +
2

2
s
s
l +
m




1
2 s2
2 2
s
+1

0 s
2

2m
2m

1
.
2
+

(22)

The first term of the nonrelativistic energy levels (22) corresponds to the rest energy
of the quantum particle. The remaining terms of the energy levels (22) corresponds
to the energy levels of a harmonic oscillator under the influence of noninertial effects
of the rotating frame (6) and the topology of a disclination [72]. By taking 0,
we recover the results of Ref. [72], but for a spin-1/2 quantum particle. Hence, we
also have that influence of the topological defect breaks the degeneracy of the energy
levels of the harmonic oscillator as pointed out in Ref. [72]. In the limit 0, we
have the spectrum of energy of the harmonic oscillator in the absence of defects, but
under the influence of noninertial effects of the rotating frame (6). Besides, we have

123

Rotating effects on the Dirac oscillator

1855

the presence of the coupling between the quantum number l and the angular velocity
, which corresponds to the PageWerner et al. term [33,34].
Hence, the bound states solutions of the Dirac oscillator in the rotating frame (6)

in the cosmic string spacetime is obtained by assuming the condition m0 .


Without this condition, we cannot consider the amplitude of probability of finding
the Dirac neutral particle in the non-physical region of the spacetime being very
small.
On the other hand, let us discuss the behaviour of the Dirac oscillator without

assuming that m0 . Returning to the solution of Eq. (20), we have that by


imposing the condition where he confluent hypergeometric series becomes a polynomial of degree n, then, the radial wave function becomes finite everywhere including
the non-physical region of the spacetime established in Eq. (5). Thereby, a normalized
wave function can be obtained if we impose first m0 is very small. The assumption
m0 1 allows us to consider a fixed radius 0 = 1/ and a fixed value for the
parameter of the confluent hypergeometric function B = |s | + 1 in such a way that
s
the parameter of the confluent hypergeometric function A = |2s | + 21 4m
can
0 s
be considered large, without loss of generality. In this way, we can write the confluent
hypergeometric function in the form [25,47,48,69]:


 1B


2
 (B) 0 B0
2
2

A
A,
B,

F
=
m

1 1
0
0 s 0
0
2




B

, (23)
cos
2B0 4 A0
+
2
4
where  (B) is the gamma function. Hence, our last step in order to obtain a normalized
wave function in the range 0 < < 1/ is to impose that the radial wave function
vanishes at 1/, that is,


Rs (0 ) = Rs ms 02 = 0,

(24)

where 0 = 1/. Therefore, by writing the radial wave function Rs ( ) =

|s |

e 2 2 1 F1 (A, B, ) in terms of the expression given in Eq. (23) and by applying the boundary condition (24), we obtain


En, l





|s |
3 2
1
s

+
+
1
2 s2 2 2 2s 2 s

2
m
n
+
s
0
2
4

02


1
,
l +
(25)
2
m2 +

where n = 0, 1, 2, . . . and the parameters s and s are defined in Eq. (17).


The relativistic energy levels (25) correspond the spectrum of energy of the Dirac
oscillator under the influence of noninertial effects of the rotating frame (6) and the
topology of the cosmic string spacetime by imposing the conditions of vanishing the
radial wave function at 1/ and m0 1. This corresponds to the spectrum

123

1856

K. Bakke

of energy of the Dirac oscillator under the influence of a hard-wall confining potential
induced by noninertial effects. In this case, the geometry of the spacetime plays the role
of a hard-wall confining potential due to the presence of noninertial effects that restricts
the physical region of the spacetime [25,47,48]. Observe the difference between the
relativistic energy levels obtained in Eqs. (21) and (25) stems from the restriction of
the physical region of the spacetime imposed by noninertial effects and the conditions
imposed on the Dirac oscillator frequency 0 . We can also see that the curvature
effects on the relativistic energy levels (25) change the degeneracy of the spectrum
of energy. Besides, by taking the limit 1, the effects of curvature vanish and
we obtain that Eq. (25) is the relativistic spectrum of energy of the Dirac oscillator
under the influence of noninertial effects of the rotating frame (6) in the Minkowski
spacetime. Again, we can see that the noninertial effects of the rotating frame (6) yield
new contributions to the relativistic energy levels (25) in contrast to the results of Ref.
[25] obtained in the FermiWalker reference frame [71].
Finally, let us obtain the nonrelativistic limit of the energy levels (25). By applying
the Taylor expansion, then, the nonrelativistic limit of the energy levels (25) are
En l





|s |
s
3 2
2 s2
2 2
1
n
+
s

+
+
1

m+

0
2
4

2m
2m
2m02


1
2 s
l +
,
(26)
s
m
2

where the first term of the nonrelativistic energy levels (26) corresponds to the rest
energy of the quantum particle.
The nonrelativistic energy levels (26) correspond to the spectrum of energy of a
harmonic oscillator under the influence of a hard-wall confining potential induced by
noninertial effects and the topology of a disclination [51,52,72]. We can also observe
that the degeneracy of the energy levels of the harmonic oscillator is broken by the
topology of the defect. By taking the limit 1, we obtain the spectrum of energy of
the harmonic oscillator under the influence of a hard-wall confining potential induced
by noninertial effects in the absence of defects. Furthermore, we have that the energy
levels (26) are proportional to n 2 in contrast to the previous result given in Eq. (22),
whose spectrum of energy is proportional to the quantum number n. However, in both
cases, we have the presence of the PageWerner et al. term [33,34], which is the
coupling between the quantum number l and the angular velocity given by the last
term of Eq. (26).
3 Conclusions
In this paper, we have studied the influence of noninertial effects of a rotating frame and
the topology of the cosmic string spacetime on the Dirac oscillator. We have shown that
both the topology of the cosmic string spacetime and the noninertial effects restrict the
physical region of the spacetime where the quantum particle can be placed. Moreover,
by analysing the behaviour of the oscillatory frequency, we have seen that two different
bound states solutions for the Dirac equation can be obtained.

123

Rotating effects on the Dirac oscillator

1857

We have shown by assuming m0 that the amplitude of the wave function


becomes very small for values of the radial coordinate given by > 1/, therefore,
the wave function could be considered to be normalized inside the physical region
of the spacetime defined in the range 0 < < 1/. As a consequence of the

assumption m0 , we have obtained the relativistic spectrum of energy of the


Dirac oscillator under the influence of the noninertial effects of a rotating frame and
the topology of the cosmic string spacetime. Moreover, we have seen that the effects of
curvature concentrated on the cosmic string axis break the degeneracy of the spectrum
of energy of the Dirac oscillator.

On the other hand, without assuming the condition m0 , we have seen that
we cannot normalize the radial wave function by imposing that the confluent hypergeometric series becomes a polynomial of degree n, because the radial wave function
becomes finite either in the physical region of the spacetime and in the non-physical
region of the spacetime. Thereby, we have shown that a normalized wave function can
be obtained by assuming that m0 is very small and by imposing that the radial wave
function vanishes at 1/. In this case, we have obtained the relativistic energy
levels that corresponds to the spectrum of energy of the Dirac oscillator under the
influence of a hard-wall confining potential induced by noninertial effects, where the
geometry of the spacetime plays the role of a hard-wall confining potential [25,47,48].
Again, we have observed that the effects of curvature concentrated on the cosmic string
axis change the degeneracy of the spectrum of energy of the Dirac oscillator under the
influence of a hard-wall confining potential induced by noninertial effects.
Finally, we have discussed in all cases above the nonrelativistic limit of the energy

levels. By assuming the condition m0 , we have seen that the nonrelativistic


energy levels correspond to the harmonic oscillator spectrum under the influence of
noninertial effects of the rotating frame (6) and the topology of a disclination [72].

Without assuming the condition m0 , but by considering m0 1 and


by imposing that the radial wave function vanishes at 1/, we have seen that
the nonrelativistic energy levels (26) correspond to the harmonic oscillator spectrum
under the influence of a hard-wall confining potential induced by noninertial effects
and the topology of a disclination [51,52,72].
Acknowledgments The author would like to thank CNPq (Conselho Nacional de Desenvolvimento Cientfico e TecnolgicoBrazil) for financial support.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.

Moshinsky, M., Szczepaniak, A.: J. Phys. A Math. Gen. 22, L817 (1989)
Villalba, V.M.: Phys. Rev. A 49, 586 (1994)
It, D., Mori, K., Carrieri, E.: Nuovo Cimento A 51, 1119 (1967)
Rozmej, P., Arvieu, R.: J. Phys. A 32, 5367 (1999)
Bermudez, A., Martin-Delgado, M.A., Solano, E.: Phys. Rev. A 76, 041801(R) (2007)
Bermudez, A., Martin-Delgado, M.A., Luis, A.: Phys. Rev. A 77, 033832 (2008)
Bermudez, A., Martin-Delgado, M.A., Luis, A.: Phys. Rev. A 77, 063815 (2008)
Bermudez, A., Martin-Delgado, M.A., Solano, E.: Phys. Rev. Lett. 99, 123602 (2007)
Jaynes, E.T., Cummings, F.W.: Proc. IEEE 51, 89 (1963)
Martnez-y-Romero, R.P., Nez-Ypez, H.N., Salas-Brito, A.L.: Eur. J. Phys. 16, 135 (1995)
de Castro, A.S., Alberto, P., Lisboa, R., Malheiro, M.: Phys. Rev. C 73, 054309 (2006)

123

1858
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.
42.
43.
44.

45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.

K. Bakke

Pacheco, M.H., Landim, R.R., Almeida, C.A.S.: Phys. Lett. A 311, 93 (2003)
Mandal, B.P., Verma, S.: Phys. Lett. A 374, 1021 (2010)
Quesne, C., Moshinsky, M.: J. Phys. A Math. Gen. 23, 2263 (1990)
Moreno, M., Zanttella, A.: J. Phys. A Math. Gen. 22, L821 (1989)
Bentez, J., Martnez y Romero, R.P., Nez-Ypez, H.N., Salas-Brito, A.L. Phys. Rev. Lett. 64, 1643
(1990)
Bentez, J., Martnez y Romero, R.P., Nez-Ypez, H.N., Salas-Brito, A.L. Phys. Rev. Lett. 65, 2085
(1990)
Quesne, C.: Int. J. Mod. Phys. A 6, 1567 (1991)
Quesne, C., Tkachuk, V.M.: J. Phys. A Math. Gen. 38, 1747 (2005)
Ferkous, N., Bounames, A.: Phys. Lett. A 325, 21 (2004)
Gonzlez-Daz, L., Villalba, V.M.: Phys. Lett. A 352, 202 (2006)
Villalba, V.M., Rojas, C.: Int. J. Mod. Phys. A 21, 313 (2006)
Gonzlez-Daz, L.A., Villalba, V.M.: Mod. Phys. Lett. A 20, 2245 (2005)
Martnez y Romero, R.P., Salas Brito, A.L.: J. Phys. A Math. Gen. 33, 1831 (1992)
Bakke, K.: Eur. Phys. J. Plus 127, 82 (2012)
Carvalho, J., Furtado, C., Moraes, F.: Phys. Rev. A 84, 032109 (2011)
Bakke, K., Furtado, C.: Phys. Lett. A 376, 1260 (2012)
Sagnac, M.G.: C. R. Acad. Sci. (Paris) 157, 708 (1913)
Sagnac, M.G.: C. R. Acad. Sci. (Paris) 157, 1410 (1913)
Post, E.J.: Rev. Mod. Phys. 39, 475 (1967)
Mashhoon, B.: Phys. Rev. Lett. 61, 2639 (1988)
Vignale, G., Mashhoon, B.: Phys. Lett. A 197, 444 (1995)
Page, L.A.: Phys. Rev. Lett. 35, 543 (1975)
Werner, S.A., Staudenmann, J.-L., Colella, R.: Phys. Rev. Lett. 42, 1103 (1979)
Hehl, F.W., Ni, W.-T.: Phys. Rev. D 42, 2045 (1990)
Papini, G.: Phys. Lett. A 377, 960 (2013)
Anandan, J.: Phys. Rev. D 15, 1448 (1977)
Anandan, J.: Phys. Lett. A 195, 284 (1994)
Dresden, M., Yang, C.N.: Phys. Rev. D 20, 1846 (1979)
Iyer, B.R.: Phys. Rev. D 26, 1900 (1982)
Letaw, J.R., Pfautsch, J.D.: Phys. Rev. D 22, 1345 (1980)
Konno, K., Takahashi, R.: Phys. Rev. D 85, 061502(R) (2012)
Damio Soares, I., Tiomno, J.: Phys. Rev. D 54, 2808 (1996)
Anandan, J., Suzuki, J.: In: Rizzi, G., Ruggiero, M.L. (eds). Relativity in Rotating Frames, Relativistic
Physics in Rotating Reference Frame. Kluwer Academic Publishers. Dordrecht, pp. 361369 (2004);
arXiv:quant-ph/0305081
Bakke, K.: Brazilian J. Phys. 42, 437 (2012)
Bakke, K.: Phys. Lett. A 374, 4642 (2010)
Bakke, K.: Eur. Phys. J. B 85, 354 (2012)
Bakke, K.: Mod. Phys. Lett. B 27, 1350018 (2013)
Vilenkin, A.: Phys. Rep. 121, 263 (1985)
Kibble, T.W.B.: J. Phys. A 19, 1387 (1976)
Katanaev, M.O., Volovich, I.V.: Ann. Phys. (NY) 216, 1 (1992)
Furtado, C., Moraes, F.: Phys. Lett. A 188, 394 (1994)
Moraes, F.: Brazilian J. Phys. 30, 304 (2000)
Sokolov, D.D., Starobinskii, A.A.: Sov. Phys. Dokl. 22, 312 (1977)
Deser, S., Jackiw, R.: Ann. Phys. (NY) 192, 352 (1989)
Furtado, C., Moraes, F., Bezerra, V.B.: Phys. Rev. D 59, 107504 (1999)
Bezerra, V.B.: Phys. Rev. D 35, 2031 (1987)
Spinelly, J., Bezerra de Mello, E.R., Bezerra, V.B.: Class. Quantum Grav. 18, 1555 (2001)
Belich, H., et al.: Phys. Rev. D 83, 25025 (2011)
Bezerra, V.B., dos Santos, I.B.: Eur. J. Phys. 13, 122 (1992)
Landau, L.D., Lifshitz, E.M.: The Classical Theory of Fields, Course of Theoretical Physics, vol. 2,
4th edn. Elsevier, Oxford (1980)
Weinberg, S.: Gravitation and Cosmology: Principles and Applications of the General Theory of
Relativity. IE-Wiley, New York (1972)

123

Rotating effects on the Dirac oscillator

1859

63. Birrel, P., Davies, P.: Quantum Fields in Curved Space. Cambridge University Press, Cambridge, UK
(1982)
64. Nakahara, M.: Geometry, Topology and Physics. Institute of Physics Publishing Bristol, Bristol (1998)
65. Bakke, K., Furtado, C.: Phys. Rev. D 80, 024033 (2009)
66. Bakke, K., Furtado, C.: Phys. Rev. D 82, 084025 (2010)
67. Greiner, W.: Relativistic Quantum Mechanics: Wave Equations, 3rd edn. Springer, Berlin (2000)
68. Schlter, P., Wietschorke, K.-H., Greiner, W.: J. Phys. A 16, 1999 (1983)
69. Abramowitz, M., Stegum, I.A.: Handbook of Mathematical Functions. Dover Publications Inc.,
New York (1965)
70. Landau, L.D., Lifshitz, E.M.: Quantum Mechanics, the Nonrelativistic Theory, 3rd edn. Pergamon,
Oxford (1977)
71. Misner, C.W., Thorne, K.S., Wheeler, J.A.: Gravitation. W. H Freeman, San Francisco (1973)
72. Furtado, C., Moraes, F.: J. Phys. A Math. Gen. 33, 5513 (2000)

123

S-ar putea să vă placă și