Вы находитесь на странице: 1из 8

Annals of Physics 346 (2014) 5158

Contents lists available at ScienceDirect

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

Torsion and noninertial effects on a


nonrelativistic Dirac particle
K. Bakke
Departamento de Fsica, Universidade Federal da Paraba, Caixa Postal 5008, 58051-970,
Joo Pessoa, PB, Brazil

highlights
Torsion effects on a spin-1/2 particle in a noninertial reference frame.
FermiWalker reference frame in the cosmic dislocation spacetime background.
Torsion and noninertial effects on the confinement to a hard-wall confining potential.

article

info

Article history:
Received 14 February 2014
Accepted 8 April 2014
Available online 18 April 2014
Keywords:
Torsion
Screw dislocation
Noninertial effects
Cosmic dislocation
FermiWalker reference frame
Hard-wall confining potential

abstract
We investigate torsion and noninertial effects on a spin-1/2 quantum particle in the nonrelativistic limit of the Dirac equation. We
consider the cosmic dislocation spacetime as a background and
show that a rotating system of reference can be used out to distances which depend on the parameter related to the torsion of
the defect. Therefore, we analyse torsion effects on the spectrum
of energy of a nonrelativistic Dirac particle confined to a hard-wall
potential in a FermiWalker reference frame.
2014 Elsevier Inc. All rights reserved.

1. Introduction
In recent decades, a great deal of works has studied the influence of torsion on several physical
systems [111]. The interaction between fermions and torsion and possible physical effects have been
discussed in Refs. [46]. In crystalline solids, torsion has been studied in the continuum picture of
defects by using the differential geometry in order to describe the strain and the stress induced by
the defect in an elastic medium [1217]. The study of the influence of torsion on quantum systems

Tel.: +55 83 8801 5202.


E-mail addresses: kbakke@fisica.ufpb.br, kbakkef@gmail.com.

http://dx.doi.org/10.1016/j.aop.2014.04.003
0003-4916/ 2014 Elsevier Inc. All rights reserved.

52

K. Bakke / Annals of Physics 346 (2014) 5158

has been extended to the electronic properties of graphene sheets [18], Berrys phase [19], quantum
scattering [20], Landau levels for a nonrelativistic scalar particle [21] and holonomic quantum
computation [22]. The influence of torsion on a two-dimensional quantum ring has been discussed
in Ref. [23] and on a quantum dot in Ref. [24].
In this paper, we discuss torsion effects on the spectrum of energy of a nonrelativistic spin-1/2
particle confined to a hard-wall confining potential in a noninertial reference frame by considering the
cosmic dislocation spacetime as a background and by showing that a rotating system of reference can
be used out to distances which depend on the parameter related to the torsion of the defect. Studies
of noninertial effects have discovered quantum effects related to geometric phases [2527] and the
coupling between the angular momentum and the angular velocity of the rotating frame [2830].
In other quantum systems, the study of noninertial effects have been extended to the weak field
approximation [31], the influence of Lorentz transformations [32], scalar fields [33], Dirac fields [34],
persistent currents in quantum rings [35], spin currents [36] and rotational and gravitational effects in
quantum interference [3739]. An interesting discussion made in Ref. [40] is by making a coordinate
transformation (in the Minkowski spacetime) from a system at rest to a uniformly rotating frame,
then, the line element of the Minkowski spacetime is not well-defined at large distances. This means
that the coordinate system becomes singular at large distances, which is associated with the velocity
of the particle would be greater than the velocity of the light. In recent years, the behaviour of
external fields and the wave function of a neutral particle have been analysed under the influence of
noninertial effects and the presence or absence of curvature [41]. In Refs. [42,43], it has been shown
that noninertial effects can provide a confinement of a neutral particle interacting with external fields
analogous to a two-dimensional quantum dot. Thereby, our focus is to analyse torsion and noninertial
effects and fill a lack in the study of the quantum dynamics of a nonrelativistic Dirac particle.
This paper is structured as follows: in Section 2, we study the nonrelativistic quantum dynamics of
a spin-1/2 particle in the FermiWalker reference frame and in the presence of torsion. Then, we show
that a rotating system of reference can be used out to distances which depend on the parameter related
to the torsion of the defect and use the coordinate singularity to impose where the wave function must
vanish, which allows us to obtain bound states analogous to the confinement of a spin-1/2 quantum
particle to a hard-wall confining potential; in Section 3, we present our conclusions.
2. Torsion effects on the nonrelativistic quantum dynamics of a Dirac particle in a noninertial
frame
In this section, we study the confinement of a nonrelativistic Dirac particle to a hard-wall confining
potential under the influence of noninertial effects and torsion. We present the mathematical tools
to describe spinors under the influence of torsion and noninertial effects, when we consider the local
reference frame of the observers is a FermiWalker reference frame. In this paper, we work with the
units h = c = 1. We start by writing the line element of the cosmic dislocation spacetime [9,10,44]
(in the rest frame of the observers):
ds2 = dT

+ dR2 + R2 d 2 + (dZ + d )2 .

(1)

The parameter is a constant, and it is related to the torsion of the defect. From the crystallography
= b z , where = b [7,9,44].
language, the parameter is related to the Burgers vector b
2
In the following, we consider a coordinate transformation given by T = t , R = , = + t,
and Z = z, where is the constant angular velocity of the rotating frame. Thus, the line element (1)
becomes
ds2 = 1 2 2 2 2 dt 2 + 2 2 + 2 2 d dt + d 2 + 2 d 2

+ 2 dz dt + (dz + d)2 .

(2)

We can note that the line element (2) is defined for values of the radial coordinate inside the range:
1 2 2

0<<

(3)

K. Bakke / Annals of Physics 346 (2014) 5158

53

Therefore, for all values of the radial coordinate given by

1 2 2

, the line element (2)

1 2 2

is not well-defined. Observe that for values of the radial coordinate given by
, we

have that the component gtt of the metric tensor becomes positive. The meaning of this restriction
of the
radial coordinate is that, in the rotating system, the coordinate system becomes singular at

1 2 2

, which is associated with the velocity of the particle would be greater than the velocity

of the light as discussed in Ref. [40]. Moreover, the range (3) shows that the restriction of the radial
coordinate in the rotating frame depends on the angular velocity and the parameter associated with
the torsion of the defect. Therefore, based on the coordinate singularity given in Eq. (3), we analyse
the behaviour of the quantum particle under the influence of torsion and noninertial effects
inside
1 2 2

,
the region 0 < < 1 2 2 / by imposing that the wave function vanishes at

which yields bound states analogous to the confinement of a spin-1/2 particle to a hard-wall confining
potential [4548].
From the line element (2), we can see that we have chosen to work with curvilinear coordinates,
thus, the Dirac spinors should be worked with the mathematical formulation of the spinor theory in
curved space [49]. In a curved spacetime background, spinors are defined locally, where 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 [49]. Locally, the reference frame of the observers
can be build via a noncoordinate basis a = ea (x) dx , where the components ea (x) are called
tetrads and satisfy the relation: g (x) = ea (x) eb (x) ab [4951], where ab = diag( + ++) is

the Minkowski tensor. The inverse of the tetrads are defined as dx = e a (x) a , and the relations
ea (x) e b (x) = a b and e a (x) ea (x) = are satisfied. In the FermiWalker reference frame,
noninertial effects can be observed from the action of external forces without any influence of
arbitrary rotations of the spatial axis of the local frame [52]. This reference frame can be built by taking
0 = e0 t (x) dt, which means that the components of the noncoordinate basis form a rest frame for the
observers at each instant, and the spatial components of the noncoordinate basis i must be chosen
in such a way that they do not rotate [52]. With these conditions, we can write

0 = dt ;

1 = d;

2 = dt + d;

3 = dt + d + dz .

(4)

Recently, the FermiWalker reference frame has been used in studies of the analogue effect of
the AharonovCasher effect [53], and the Dirac oscillator [54]. Hence, by solving the MaurerCartan
structure equations in the presence of torsion T a = d a + a b b [51], where T a = 21 T a dx dx
is the torsion 2-form, a b = a b (x) dx is the connection 1-form, the operator d corresponds to
the exterior derivative and the symbol means the wedge product, we obtain the following non-null
components of the connection 1-form:
T 3 = 2 () () d d;

1 2 (x) = 2 1 (x) = 1;
t

1
2

(x) =

2
t 1

(5)

(x) = .

In order to write the Dirac equation in a curved spacetime background and in the presence of torsion, we need to take into account that the partial derivative becomes the covariant derivative, where
the covariant derivative is given by = + (x)+ K (x), with (x) = 4i ab (x) ab be-

ing the spinorial connection [50,51], and K (x) = 4i Kab (x) ab , with ab = 2i a , b . The indices
(a, b, c = 0, 1, 2, 3) indicate the local reference
frame. The connection 1-form
Kab (x) is related to

the contortion tensor by [5]: Kab = K e a (x) e b (x) e b (x) e a (x) . Following the definitions

of Ref. [5], the contortion tensor is related to the torsion tensor via K = 12 T T T ,
where we have that the torsion tensor is antisymmetric in the last two indices, while the contortion tensor is antisymmetric in the first two indices. Moreover, it is usually convenient to write
the torsion tensor into three irreducible components: the trace vector T = T , the axial vector

54

K. Bakke / Annals of Physics 346 (2014) 5158

S = T and in the tensor q , which satisfies the conditions q = 0 and q = 0.

Thus, the torsion tensor becomes: T = 31 T g T g 16 S + q . The a matrices are defined in the local reference frame and correspond to the Dirac matrices in the Minkowski
spacetime [50,55], i.e.,

0 = =

I
0

0
I

i = i =

(6)

and I being the spin vector and the 2 2 identity


with
matrix, respectively. The matrices are the
Pauli matrices and satisfy the relation i j + j i = 2ij . The matrices are related to the a
matrices via = e a (x) a [50]. In this way, the general expression for the Dirac equation describing
torsion effects on a quantum particle in the FermiWalker reference frame (4) is

i 0
+ i 1
+i
m = i
t

+ i 3

+ i (x)
z

1
S .
S0 0 5
(7)
8
8
From the results obtained in Eq. (5), we can calculate the non-null components of the spinorial

connection (x), and obtain i = i 2 [42]. Moreover, from the definition of the axial vector S
given above and the results given in Eq. (5), we have that the only non-null component of the axial
4
vector is S 0 = () () [56]. Hence, for = 0, the Dirac equation (7) becomes

+ i
i
= m
i 1
t

2
i

i 3

.
z

(8)

Now, let us discuss the nonrelativistic behaviour of the spin-1/2 particle. We can obtain the nonrelativistic dynamics of the spin-1/2 particle by writing the solution of the Dirac equation (8) in the form

=e

imt

(9)

where and are two-spinors, and we consider being the large component and being the
small component [55]. Substituting (9) into the Dirac equation (8), we obtain two coupled equations of and . The first coupled equation is

i1
i2

1
3
i
= i

i
,
i
t

z
z

(10)

while the second coupled equation is


i

i1
i2

+ 2m i
= i 1

i3
.
t

z
z

(11)

With being the small component of the wave function, we can consider |2m| i t , and

|2m| i
, thus, we can write

1
i1
i2

i 1

i3
.
(12)
2m

z
z
Substituting of the expression (12) into (10), we obtain a second order differential equation given
by

1
i
=
t
2m
+

2
1
1
+
+ 2
2

1
8m

+ i

2
i 3

+ 2 +

z
2m 2
z
(13)

K. Bakke / Annals of Physics 346 (2014) 5158

55

which corresponds to the SchrdingerPauli equation for a spin-1/2 particle under the influence of
the noninertial effects of the FermiWalker reference frame, and a screw dislocation (torsion). Note
that is an eigenfunction of 3 in Eq. (13), whose eigenvalues are s = 1. Thus, we can write
3 s = s = ss . We can see that the operators p z = iz and Jz = i [57] commute
with the Hamiltonian of the right-hand side of (13), thus, we write the solution of (13) in terms of
the eigenvalues of the operator p z = iz , and the z-component of the total angular momentum
Jz = i 1 :

s = e

1
iE t i l+ 2 ikz

R+ ()
,
R ()

(14)

where l = 0, 1, 2, . . . and k is a constant. Substituting the solution (14) into the SchrdingerPauli
equation (13), we obtain two noncoupled equations for R+ and R . After some calculations, we can
write the noncoupled equations for R+ and R in the following compact form:
Rs +

2
1
Rs s2 Rs + 2 Rs = 0,

(15)

where we have defined the following parameters:


1

= s = l + (1 s) k
2

1
k2
2
= 2m E + l +

.
2

(16)

2m

The second order differential equation (15) corresponds to the Bessel differential equation [58]. The
general solution of (15) is given by: Rs () = A Js ()+ B Ns (), where Js () and Ns () are the
Bessel functions of first and second kinds [58]. In order to have a regular solution at the origin, we must
take B = 0 in the general solution of Eq. (15), since the Neumann function diverges at the origin. Thus,
the solution of (15) becomes: Rs () = A J|s | (). Moreover, we wish to obtain a normalized wave

12 2

function inside the region 0 < <


, therefore we consider the spin-1/2 quantum particle

is confined to
a hard-wall confining potential by imposing that the radial wave function vanishes at

0 =

Rs

12 2

, that is,

0 =

1 2 2

= 0.

(17)

12 2

Then, by assuming that 0 1 (where 0 =

J|s | (0 )

cos 0

|s |
2

), and we can take [43,58]

(18)

Hence, substituting (18) into (17), we obtain

En, l

n +
2m 1 2 2
2
1

2
2

l + 1 (1 s) k + 3 + k [l + 1/2] .

2
4
2m

(19)

Eq. (19) is the spectrum of energy of a nonrelativistic Dirac particle confined to a hard-wall confining potential under the influence of torsion and noninertial effects. We can observe the influence of

1 It has been shown in Ref. [57] that the z-component of the total angular momentum in cylindrical coordinates is given by
Jz = i , where the eigenvalues are = l 21 .

56

K. Bakke / Annals of Physics 346 (2014) 5158

torsion on the energy levels (19) in the effective angular momentum given by s = l + 12 (1 s) k,

12 2

and by the presence of the fixed radius 0 =


. Note that the spectrum of energy obtained in

2
Eq. (19) is proportional to n (parabolic energy spectrum) in contrast to recent studies of the influence
of noninertial effects on the Landau quantization for neutral particles [41], whose spectrum of energy
is proportional to the quantum number n. This results from the imposition of having the wave function being normalized in the space
restricted by the range (3), where we have considered the wave
12 2

function vanishing at 0 =
and assumed 0 1. It is worth mentioning an analogy

between the present study and studies of confinement of quantum particle to a quantum dot in condensed matter systems. The result obtained in Eq. (19) agrees with the studies of the confinement of
quantum particles to a quantum dot made in Refs. [4548] where the quantum dot models provide a
parabolic spectrum of energy (in relation to the quantum number n). The main difference between the
present work and the quantum dot models of Refs. [4548] is that the hard-wall confining potential is
determined by the geometry of the manifold (described by the line element (2)). Hence, the geometrical approach yielded in the present work can be useful in studies of noninertial effects in condensed
matter systems possessing the presence of screw dislocations [9,43].
We also obtain in Eq. (19) the PageWerner et al. term [2830], which corresponds to the coupling
between the angular velocity and the quantum number l. Returning to the analogy with condensed
matter systems, we have that the presence of the PageWerner et al. term in Eq. (19) agrees with
the studies of the confinement of a neutral particle to a quantum dot analogous to the TanInkson
model [59] made in Ref. [42]. Finally, we can see in the last term of Eq. (19) that there is no influence
of the torsion on the PageWerner et al. term [2830].
3. Conclusions
In this brief report, we have discussed torsion and noninertial effects on the confinement of a
nonrelativistic Dirac particle to a hard-wall confining potential.
We have seen, in the rotating system,

12 2

, where the restriction of the values of


that the coordinate system becomes singular at

the radial coordinate depends on the parameter related to the torsion of the defect and the
angular
12 2

velocity. We have also shown that by considering the wave function vanishing at 0 =

and 0 1, we can obtain bound states whose spectrum of energy is parabolic in relation to the
quantum number n. We also have shown that the influence of torsion on the energy levels is given
by the presence of an effective angular momentum s = l + 12 (1 s) k and by a fixed radius

0 =

12 2

. We have also obtained the coupling between the angular velocity and the quantum

number l, which is known as the PageWerner et al. term [2830]. Moreover, this study has shown
that there exists no influence of the torsion on the PageWerner et al. term [2830].
We would like to add a comment on the influence of torsion and noninertial effects in quantum
systems. It has been shown in Ref. [11] that the presence of torsion modifies the electromagnetic field
in the rest frame of the observers. Following the study made in Ref. [11], it should be interesting
to consider either a charged particle or a neutral particle interacting with external fields in the
presence of torsion in a noninertial reference frame. Both torsion and noninertial effects can yield
new field configurations and, consequently, new contributions to the energy levels can be obtained.
Furthermore, the geometrical approach used in this work can be useful in studies of quantum dots
in condensed matter systems described by the Dirac equation such as graphene [18,60], topological
insulators [61] and cold atoms [62]. A different context of using the present geometrical approach can
be on the behaviour of the Dirac oscillator [54] in the cosmic dislocation spacetime and the Casimir
effect [63,64]. Another interesting quantum effect arises from the presence of a quantum flux in the
energy levels of bound states called persistent currents. For instance, in Ref. [65], persistent currents
arise from the dependence of the energy levels on the Berry phase [66] and the AharonovAnandan
quantum phase [67]. In Ref. [35], persistent currents have been investigated in a quantum ring from
rotating effects. Hence, the study of persistent currents should be interesting in a scenario where there
exists the presence of torsion in an elastic medium in a noninertial frame.

K. Bakke / Annals of Physics 346 (2014) 5158

57

Acknowledgement
The author would like to thank CNPq (Conselho Nacional de Desenvolvimento Cientfico e
TecnolgicoBrazil) for financial support.
References
[1] F.W. Hehl, Gen. Relativity Gravitation 4 (1973) 333;
F.W. Hehl, et al., Rev. Modern Phys. 48 (1976) 393.
[2] J. Audretsch, Phys. Rev. D 24 (1981) 1470.
[3] J. Audretsch, J. Phys. A: Math. Gen. 14 (1981) 411;
F. Cianfrani, G. Montani, Europhys. Lett. EPL 84 (2008) 30008.
[4] L.H. Ryder, I.L. Shapiro, Phys. Lett. A 247 (1998) 21.
[5] I.L. Shapiro, Phys. Rep. 357 (2002) 113.
[6] P. Singh, L.H. Ryder, Classical Quantum Gravity 14 (1997) 3513;
C. Lmmerzahl, Phys. Lett. A 228 (1997) 223.
[7] H. Kleinert, Gauge Fields in Condensed Matter, Vol. 2, World Scientific, Singapore, 1989.
[8] F. Moraes, Phys. Lett. A 214 (1996) 189;
A. de Padua, et al., Phys. Lett. A 238 (1998) 153.
[9] F. Moraes, Braz. J. Phys. 30 (2000) 304.
[10] R.A. Puntingam, H.H. Soleng, Classical Quantum Gravity 14 (1997) 1129.
[11] L. Dias, F. Moraes, Braz. J. Phys. 35 (2005) 636.
[12] E. Krner, Appl. Mech. Rev. 15 (1962) 599;
B.A. Bilby, et al., Proc. R. Soc. Lond. Ser. A 231 (1955) 263.
[13] I.E. Dzyaloshinskii, G.E. Volovik, Ann. Phys. 125 (1980) 67;
R. Bausch, et al., Phys. Rev. Lett. 80 (1998) 2257.
[14] M.O. Katanaev, I.V. Volovich, Ann. Physics (NY) 216 (1992) 1.
[15] K.C. Valanis, V.P. Panoskaltsis, Acta Mech. 175 (2005) 77.
[16] A.L. Silva Netto, C. Furtado, J. Phys.: Condens. Matter 20 (2008) 125209.
[17] K. Bakke, F. Moraes, Phys. Lett. A 376 (2012) 2838.
[18] F. de Juan, et al., Nuclear Phys. B 828 (2010) 625;
A. Mesaros, et al., Phys. Rev. B 82 (2010) 073405.
[19] C. Furtado, V.B. Bezerra, F. Moraes, Europhys. Lett. 52 (2000) 1.
[20] C. Furtado, V.B. Bezerra, F. Moraes, Phys. Lett. A 289 (2001) 160.
[21] C. Furtado, F. Moraes, Europhys. Lett. 45 (1999) 279.
[22] K. Bakke, C. Furtado, Phys. Lett. A 375 (2011) 3956.
[23] A.L. Silva Netto, C. Chesman, C. Furtado, Phys. Lett. A 372 (2008) 3894.
[24] E. Aurell, J. Phys. A: Math. Gen. 32 (1999) 571.
[25] M.G. Sagnac, C. R. Acad. Sci. Paris 157 (1913) 708;
M.G. Sagnac, C. R. Acad. Sci. Paris 157 (1913) 1410.
[26] E.J. Post, Rev. Modern Phys. 39 (1967) 475.
[27] B. Mashhoon, Phys. Rev. Lett. 61 (1988) 2639.
[28] L.A. Page, Phys. Rev. Lett. 35 (1975) 543.
[29] S.A. Werner, J.-L. Staudenmann, R. Colella, Phys. Rev. Lett. 42 (1979) 1103.
[30] F.W. Hehl, W.-T. Ni, Phys. Rev. D 42 (1990) 2045.
[31] J. Anandan, J. Suzuki, in: G. Rizzi, M.L. Ruggiero (Eds.), Relativity in Rotating Frames, Relativistic Physics in Rotating
Reference Frame, Kluwer Academic Publishers, Dordrecht, 2004, pp. 361369. arXiv:quant-ph/0305081.
[32] I. Damio Soares, J. Tiomno, Phys. Rev. D 54 (1996) 2808.
[33] J.R. Letaw, J.D. Pfautsch, Phys. Rev. D 22 (1980) 1345;
K. Konno, R. Takahashi, Phys. Rev. D 85 (2012) 061502(R).
[34] B.R. Iyer, Phys. Rev. D 26 (1982) 1900.
[35] G. Vignale, B. Mashhoon, Phys. Lett. A 197 (1995) 444.
[36] G. Papini, Phys. Lett. A 377 (2013) 960.
[37] J. Anandan, Phys. Rev. D 15 (1977) 1448.
[38] J. Anandan, Phys. Lett. A 195 (1994) 284.
[39] M. Dresden, C.N. Yang, Phys. Rev. D 20 (1979) 1846.
[40] L.D. Landau, E.M. Lifshitz, The Classical Theory of Fields, fourth ed., in: Course of Theoretical Physics, vol. 2, Elsevier, Oxford,
1980.
[41] K. Bakke, Ann. Phys. (Berlin) 523 (2011) 762.
[42] K. Bakke, Phys. Lett. A 374 (2010) 4642.
[43] K. Bakke, Modern Phys. Lett. B 27 (2013) 1350018.
[44] D.V. Galtsov, P.S. Letelier, Phys. Rev. D 47 (1993) 4273;
P.S. Letelier, Classical Quantum Gravity 12 (1995) 471.
[45] N. Aquino, E. Castao, E. Ley-Koo, Chinese J. Phys. 41 (2003) 276.
[46] L. Solimany, B. Kramer, Solid State Commun. 96 (1995) 471.
[47] E. Tsitsishvili, G.S. Lozano, A.O. Gogolin, Phys. Rev. B 70 (2004) 115316.
[48] R. Rosas, R. Riera, J.L. Marn, J. Phys.: Condens. Matter 12 (2000) 6851.
[49] S. Weinberg, Gravitation and Cosmology: Principles and Applications of the General Theory of Relativity, IE-Wiley, New
York, 1972.

58
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]

[61]
[62]
[63]
[64]
[65]
[66]
[67]

K. Bakke / Annals of Physics 346 (2014) 5158


N.D. Birrell, P.C.W. Davies, Quantum Fields in Curved Space, Cambridge University Press, Cambridge, UK, 1982.
M. Nakahara, Geometry, Topology and Physics, Institute of Physics Publishing, Bristol, 1998.
C.W. Misner, K.S. Thorne, J.A. Wheeler, Gravitation, W. H. Freeman, San Francisco, 1973.
K. Bakke, C. Furtado, Eur. Phys. J. C 69 (2010) 531.
K. Bakke, Eur. Phys. J. Plus 127 (2012) 82.
W. Greiner, Relativistic Quantum Mechanics: Wave Equations, third ed., Springer, Berlin, 2000.
K. Bakke, C. Furtado, J.R. Nascimento, Eur. Phys. J. C 60 (2009) 501.
P. Schlter, K.-H. Wietschorke, W. Greiner, J. Phys. A 16 (1983) 1999.
M. Abramowitz, I.A. Stegum, Handbook of Mathematical Functions, Dover Publications Inc., New York, 1965.
W.-C. Tan, J.C. Inkson, Semicond. Sci. Technol. 11 (1996) 1635;
W.-C. Tan, J.C. Inkson, Phys. Rev. B 53 (1996) 6947.
P.E. Lammert, V.H. Crespi, Phys. Rev. Lett. 85 (2000) 5190;
P.E. Lammert, V.H. Crespi, Phys. Rev. B 69 (2004) 035406;
J. Gonzales, F. Guinea, M.A.H. Vozmediano, Phys. Rev. Lett. 69 (1992) 172;
C. Furtado, et al., Phys. Lett. A 372 (2008) 5368.
M.A. Hasan, C.L. Kane, Rev. Modern Phys. 82 (2010) 3045.
O. Boada, A. Celi, J.I. Latorre, M. Lewenstein, New J. Phys. 13 (2011) 035002.
V.B. Bezerra, V.M. Mostepanenko, H.F. Mota, C. Romero, Phys. Rev. D 84 (2011) 104025.
H.F. Mota, K. Bakke, Phys. Rev. D 89 (2014) 027702.
D. Loss, P. Goldbart, A.V. Balatsky, Phys. Rev. Lett. 65 (1990) 1655.
M.V. Berry, Proc. R. Soc. Lond. Ser. A 392 (1984) 45.
Y. Aharonov, J. Anandan, Phys. Rev. Lett. 58 (1987) 1593.

Вам также может понравиться