Jamp 2020120914361061
Jamp 2020120914361061
https://www.scirp.org/journal/jamp
ISSN Online: 2327-4379
ISSN Print: 2327-4352
Martin Tcoffo1,2, Germain Yinde Deuto1, Issofa Nsangou1, Armel Azangue Koumetio1,
Lylyane S. Yonya Tchapda1, Alain G. Tene1
Research Unit of Condensed Matter, Electronics and Signal Processing, University of Dschang, Dschang, Cameroon
1
Centre d’Etude et de Recherche en Agronomie et en Biodiversité, FASA, Université de Dschang, Dschang, Cameroun
2
1. Introduction
The theory of open quantum systems has become one of the main interests of
physicists during the last few decades due to their strong applications in several
DOI: 10.4236/jamp.2020.812207 Dec. 10, 2020 2801 Journal of Applied Mathematics and Physics
M. Tcoffo et al.
domains of physics, especially in solid state physics [1], quantum optics [2], na-
notechnology [3], quantum measurement theory [4], etc. Indeed, various re-
search works conducted in the field proved that perfect isolated quantum system
does not exist, since any realistic system permanently interacts with its environ-
ment, which typically presents a large number of degrees of freedom [5]. A
physical system interacting with the environment is referred to an open system
[6]. Thus, every physical system is affected by the presence of its surroundings
leading to dissipation, thermalization and in the quantum case, decoherence [6].
Many reasons lead scientists to pay particular attention to the above mentioned
phenomena, here, we refer particularly to decoherence and dissipation, which
are the main obstacle in the realization of quantum computers and other quan-
tum devices on one hand [7] [8] [9]. On the other hand, several interesting pro-
posals to exploit the interaction between quantum systems and their environ-
ment to dissipatedly engineer challenging state of matter have emerged very re-
cently [6].
In fact, dissipation in an open system results from microscopic reversible in-
teractions between the observable system and the environment. It implies irre-
versibility and therefore, a preferred direction in time. However, we are particu-
larly interested in this work on the phenomenon of decoherence, which some-
what “regroups dissipation” and introduces the (partial) destruction of quantum
coherence through the interaction of quantum mechanical system with its sur-
rounding [10]. From the theoretical analysis point of view, decoherence can be
described with the help of microscopic models based on the interaction of
quantum mechanical systems with a collection of an infinite number of har-
monic oscillators, representing the environment [11] [12]. The strength and na-
ture of the coupling with environment play a crucial role in selecting the pre-
ferred states leading to a distance-dependent nature of typical interaction. Fur-
thermore, decoherence simply characterizes the transition of a given quantum
system to classical one due to interactions, thus, the particular time that a quan-
tum system starts behaving classically is known as “decoherence time” [12] [13]
[14]. Whereas, the decoherence time scale defines the time during which quan-
tum interferences between quantum states disappear (i.e. loss of quantum prop-
erties) [13] [15]. Deep studies have been made to determine this particular time
and to see whether or not it can be controlled due to its relevance in many inter-
esting physical problems such as quantum computation and quantum informa-
tion processing [16], heavy ion collisions [17], etc. In many cases, physicists are
also interested in understanding the specific causes of quantum decoherence in
order to prevent it from damaging quantum states and protect the information
stored inside [18]. Thus, it was discovered that, for most of superposed quantum
states, the associated decoherence time is often found to be smaller than the cor-
responding relaxation (damping) time. Several research works have been suc-
cessfully investigated to determine this time in the literature, many of them
found that the main tool for its investigation is the master equations (ME) dy-
namics approach, which describes the evolution in time of the density operator
2. Model Hamiltonian
The model described in this work forms the standard pattern of a damped
quantum harmonic oscillator in NC phase-space. Similar model was already in-
troduced in the literature [29] [30] [31] [32] [33], but with different approach
and for different motivations. For instance, in [30] it was used to analyze the
Landau diamagnetism in NC space, while in [32] it was introduced to study the
connection between dissipation and NCty. However, in this work our oscillator
is assumed to be a non-relativistic charged particle of mass m moving in a
two-dimensional NC phase-space under the effect of a homogeneous magnetic
field, and confined by a damped anisotropic harmonic potential with frequencies
ω1 and ω2 in the x and y directions respectively. The Hamiltonian of this sys-
tem is given by:
2 2
1 e ˆ 1 e ˆ m 2 2
H= pˆ1 − A1 +
2m c 2m
pˆ 2 − A2 +
c 2
( )
ω1 qˆ1 + ω22 qˆ22 + k12 qˆ1qˆ2 , (1)
where ( Aˆ , Aˆ )
1 2 are the components of the potential vector of the magnetic field,
k12 the damping parameter, ( qˆ1 , qˆ2 ) and ( pˆ1 , pˆ 2 ) are respectively the non-
commutative coordinates and momenta of the quantum oscillator satisfying the
following commutation relations [34] [35] [36] [37] [38] [ qˆk , qˆl ] = iθε kl ,
θη
[ pˆ k , pˆ l ] = iηε kl , [ qˆk , pˆ l ] = i δ kl , with =
1 + 2 the effective Planck constant,
4
θ and η the NC parameters with respect to the coordinates and momenta,
respectively. Due to the fact that the NCty structure of the phase-space is a pure
geometrical property, its physical effects are independent of the particle nature.
Therefore, NCty between momenta arises naturally as a consequence of NC
coordinates, since momenta are defined to be partial derivatives of the action
with respect to the NC coordinates. The NC parameters θ and η have the
dimensions of ( length ) and ( momentum ) respectively [35] [39]. The NC
2 2
coordinates and momenta in terms of the commuting ones are defined as:
θ θ
qˆ1 =q1 −
2
p2 , qˆ2 =
q2 +
2
p1
(2)
pˆ = η η
p1 + q2 , pˆ 2 = p2 − q1
1
2 2
where the coordinates qk and momenta pk obey the usual commutation rela-
tions [ qk , ql ] = 0 , [ pk , pl ] = 0 and [ qk , pl ] = iδ kl ,
k , l = 1, 2 . By the gauge
symmetric relation, the vector potentials of the magnetic field Aˆk can be de-
fined as:
B B
Aˆ = − qˆ2 , qˆ1 (3)
2 2
Using the latest relation and Equation (2), the Hamiltonian (1) can be rewrit-
ten as follows:
1 2 1 2 m 2 2
H=
2m1
p1 +
2m2 2
(
p2 + ϖ 1 q1 + ϖ 22 q22 )
θ θ2
+ k12 q1q2 + ( q1 p1 − q2 p2 ) − 2 p1 p2 (4)
2 2
− ( λ1q1 p2 − λ2 q2 p1 ) ,
where
ωc2 ωcη η2
ϖ 1 = ω12 + + + ,
4 2 m 4m 2 2
ωc2 ωcη η2
ϖ 2 = ω22 + + + ,
4 2 m 4m 2 2
(5)
mθ 2 ωc2 ωc θη η
λ=1 ω1 + + 1 + 2 + ,
2 4 2 4 2 m
mθ 2 ωc2 ωc θη η
λ=
2 ω1 + + 1 + 2 + ,
2 4 2 4 2 m
m m
with m1 = 2 2
and m2 = 2 2
. It is
mωcθ mθω2 mωcθ mθω1
1 + 4 + 2 1 + 4 + 2
important to notice that the effective masses m1 and m2 are different due to
eB
the anisotropy of the system, and ωc = is the cyclotron frequency. The last
mc
term appearing in the Hamiltonian (4) is due to the NCty effects. This term in-
troduces some difficulties to solve the eigenvalue equation of the corresponding
Hamiltonian, thus the whole Hamiltonian requires further transformations. For
this reason, let’s define a canonical transformation involving the set of coordi-
nates-momenta ( q1 , q2 , p1 , p2 ) [31] [32] [40] by:
1 1
q1 = χ cos ( Φ ) Q1 + sin ( Φ ) P2 , p1 = − ρ sin ( Φ ) Q2 + cos ( Φ ) P1 ,
ρ χ
and (6)
q = 1 1
2
χ cos ( ) 2
Φ Q + sin ( ) 1
Φ P , p
2
= − ρ sin ( ) 1
Φ Q + cos ( ) 2
Φ P .
ρ χ
The new coordinates ( Q1 , Q2 ) and momenta ( P1 , P2 ) also satisfy the ca-
nonical commutation relations:
[=
Q1 , P1 ] [Q
=2 , P2 ] i , (7)
while all other permutations vanish. Using the transformations of Equation (6),
the Hamiltonian (4) becomes:
1 1 M M
H
= P12 + P22 + 1 Ω12 Q12 + 2 Ω 22 Q22 + c12 Q1Q2
2M1 2M 2 2 2 (8)
+ µ11Q1 P2 + µ22 Q2 P2 + d12 P1 P2 ,
where
1 1
M
= 1 2
, M=
2 ,Ω
=1 2 I1 K1 , Ω=
2 2I2 K2 , (9)
2 I1 2 I 22
with
2 1 ρ2 λρ
= I1 2
cos 2 ( Φ ) + mϖ 22 sin 2 ( Φ ) + 2 sin ( 2Φ ) ,
2 ρ m1 χ 2
χ
I 2 1 ρ2 λρ
= cos 2 ( Φ ) + mϖ 12 sin 2 ( Φ ) + 1 sin ( 2Φ ) ,
2 2 ρ 2 m2 χ 2 χ
(10)
2 χ ρ
2 2
λρ
= K1 sin 2 ( Φ ) + mϖ 12 cos 2 ( Φ ) + 1 sin ( 2Φ ) ,
2 m2 χ 2 χ
χ ρ2 λρ
= K 22 sin 2 ( Φ ) + mϖ 22 cos 2 ( Φ ) + 2 sin ( 2Φ ) ,
2 m1 χ 2
χ
and
ρ χ 2 θ 2 θχ
= µ11 k12 2 + 2 sin ( 2Φ ) + ,
2χ ρ 4 ρ
ρ χ 2 θ 2 θχ
= µ22 k12 2 + 2 sin ( 2Φ ) − ,
2 χ ρ 4 ρ
(11)
θ 2ρ2
=c12 k12 χ cos 2 ( Φ ) − 2 2 sin 2 ( Φ ) ,
2
4 χ
k12 2 θ ρ
2 2
sin ( Φ ) − 2 2 cos ( Φ ) .
2
= d12 ρ2 4 χ
The set ( ρ, χ, Φ) satisfies the following relations:
ρ = (
mm1m2 λ2ϖ 12 + λ1ϖ 22 )
χ m2 λ1 + m1λ2
(12)
(
2 m1m2 ( m1λ2 + m2 λ1 ) λ2ϖ 12 + λ1ϖ 22 )
tan ( 2Φ ) =
(
m m2ϖ 12 − m1ϖ 22 )
Equation (8) is the Hamiltonian of a harmonic oscillator in two-dimensional
NC phase-space, where the mass varies in terms of the NC parameters, which
can provide simple solution to the eigenvalues equation. Table 1 depicts the
values of all the parameters and constants that appear in this Hamiltonian and in
the entire work, and which will be necessary for numerical simulations in Sec-
tion 5. In order to investigate the dynamics of two bosonic modes (Harmonic
oscillators) in weak interaction with thermal reservoir in NC phase-space, we
employ the axiomatic formalism based on the completely positive quantum dy-
namical semigroups [41].
Table 1. Recapitulation of all the parameters and constants that are used in this work.
Magnetic field B ∝ 10 −2
T
Effective mass m 1 kg
Light constant c 1 m∙s−1
Squeezing parameter δ 0.5 s−1
dρ i 1
L(ρ) =
dt
− [H , ρ ] +
=
2 j
(
∑ V j ρ ,V j + V j , ρV j , ) (13)
where H denotes the open system’s Hamiltonian and the operators Vk and Vk†
defined on the Hilbert space of H, describe the interaction of open system with
the environment. Given that, we have an interest in the set of Gaussian states, we
introduce such quantum dynamical semi group that preserves this set. Conse-
quently, H is considered to be a polynomial of second degree in coordinates Q1 ,
Q2 and momenta P1 , P2 of the quantum oscillators. In this case, V j and
V j† are taken as polynomials of first degree in these canonical observables. Then,
in the linear space spanned by the coordinates and momenta, there exist only
four linearly independent operators V j ( j = 1, 2,3, 4 ) [16] [44]. Moreover, the
operators are non-Hermitian and describe the dissipation and decoherence due
to interaction between the system and its environment. Using the transforma-
tions (2) and (6), the Lindblad operators in the NC phase-space, are given by [42]
[43]:
V j = Aj1 P1 + Aj 2 P2 + B j1Q1 + B j 2 Q2
† * * * *
(14)
V j = Aj1 P1 + Aj 2 P2 + B j1Q1 + B j 2 Q2
with
1 θ 1 η
= A j1 a j1 + b j 2 cos ( Φ ) + b j 2 + a j1 sin ( Φ ) ,
χ 2 ρ 2
1 θ 1 η
= Aj 2 a j1 − b j1 cos ( Φ ) + b j1 − a j 2 sin ( Φ ) ,
χ 2 ρ 2
(15)
B =− ρ a − b θ sin ( Φ ) + χ b − a η cos ( Φ ) ,
j1 j2 j1 j1 j2
2 2
B j 2 =− ρ a j1 + b j 2 θ sin ( Φ ) + χ b j 2 + a j1 η cos ( Φ ) ,
2 2
2 kT
∀l =1, 2 .
2
Let σ
= AA A2 − A the dispersion (variance) of a given operator A, where
A = Tr [ ρ A] is the expectation value of the operator A, ρ the statistical
AB + BA
operator (density matrix), = and σ AB − A B the correlation
2
(covariance) of the operators A and B. It was demonstrated in [46] [47] that, for
any Hermitian operators A and B, and for pure quantum states, the following
generalized uncertainty relation holds:
1 2
σ AAσ BB − σ AB
2
≥ [ A, B ] . (20)
4
From the master Equation (18), we obtain the following equation of motion
for the expectation values of coordinates and momenta [22] [48]:
dn ( t )
= Xn ( t ) , (21)
dt
T
where n ( t ) = σ Q1 , σ Q2 , σ P1 , σ P2 denotes the expectation vector and X a 4×4
matrix defined by:
1
−λ11 + µ11 −λ21
M1
−α12 + d12
1
X = −λ12 −λ22 + µ22 α12 + d12
(22)
M2
− M 1Ω1
2
β12 − c12 −λ11 − µ11 −λ12
−β − c − M 2 Ω 22 −λ21 −λ22 − µ22
12 12
However, for this limit to exist, X must have only negative real eigenvalues
parts. For this reason, we set k12 = 0 , such that µ=
ll d=
12 c=
12 0 , which lead
the system to uncoupled oscillators. Under these conditions, the matrix X be-
comes:
1
−λ11 −λ21
M1
−α12
1
X = −λ12 −λ22 α12 (25)
M2
− M 1Ω12 β12 −λ11 −λ12
− β12 −M 2Ω 2
2 −λ21 −λ22
In order to evaluate the exponential matrix N ( t ) , it is important to first di-
agonalize the matrix X by solving the corresponding secular equation i.e.
det ( X − z ) =
0, (26)
where z is the eigenvalues of X and the unit matrix. From Equation (26), one
obtains an equation of fourth order with respect to the eigenvalues z, which can
be easily solved. Let us for simplicity consider the particular case where
β=
12 α=
12 λ=
12 λ=
21 0 , then the secular equation obtained is:
( λ11 + z )2 + Ω12 ( λ22 + z )2 + Ω 22 = 0. (27)
The resolution of this equation provides z1 =−λ+ + iΩ + , z2 =−λ− + iΩ − ,
z3 =−λ+ − iΩ + and z4 =−λ− − iΩ − , where λ+ = λ11 , λ− = λ22 , Ω + =Ω1 , and
Ω − =Ω 2 . To fulfill the condition of Equation (24), only the positive values of
λ+ and λ− are considered. Thus, using the eigenvalues zn of X, the
time-dependent matrix N ( t ) can be obtained as follows:
N ij ( t ) = ∑ M in exp ( zn t ) M nj−1 , (28)
n
Considering Equation (23) and the initial conditions defined above, one get:
−λ t 1
σ Qk ( t ) e kk σ Qk ( 0 ) cos ( Ω k t ) +
= σ Pk ( 0 ) sin ( Ω k t ) ,
M k Ωk k = 1, 2. (31)
− M k Ω k σ Qk ( 0 ) sin ( Ω k t ) + σ Pk ( 0 ) cos ( Ω k t ) ,
σ Pk ( =
t) e − λkk t
Thus, we can observe that, the expectation values of the coordinates and mo-
menta decay quickly due to the exponential factors e − λ11t and e − λ22t and va-
nish when t → ∞ . Let us now determine the variance and covariance matrices.
For this reason, let σ ( t ) and D be respectively the covariance and the diffusion
coefficients matrices where the elements are defined as follows:
DQ1Q1 DQ1Q2 DQ1P1 DQ1P2 σ Q1Q1 σ Q1Q2 σ Q1P1 σ Q1P2
DQ2Q1 DQ2Q2 DQ2 P1 DQ2 P2 σ Q2Q1 σ Q2Q2 σ Q2 P1 σ Q2 P2
=D = and σ ( t ) ,
DP Q DP1Q2 DP1P1 DP1P2 σ PQ σ P1Q2 σ P1P1 σ P1P2
1 1 1 1
DP2Q1 DP2Q2 DP2 P1 DP2 P2 σ P2Q1 σ P2Q2 σ P2 P1 σ P2 P2
(32)
Thus, from the master Equation (18), the equations of motion corresponding
to the quantum correlations of canonical observables Q1 , Q2 and P1 , P2 are the
following [48] [49]:
dσ
=X σ + σ X T + 2 D, (33)
dt
where the matrix X is defined in Equation (25), with X T its transposed matrix.
The time-dependent solution of (33) is given by:
σ ( t ) N ( t ) (σ ( 0 ) − σ ( ∞ ) ) N T ( t ) + σ ( ∞ ) ,
= (34)
where the remaining elements can be easily found following the same approach.
It can be observed that the matrix elements σ ij decay rapidly to zero due to the
exponential terms exp ( −2λ11t ) , exp ( −2λ22 t ) and exp ( − ( λ11 + λ22 ) t ) . The
matrix elements of σ ( ∞ ) depend on X and D, thus should be evaluated using
the relation [22]:
∞
σ ( ∞ ) =2∫0 N ( t ′ ) DN ( t ′ ) dt ′.
T
(36)
dρ 2 i ∂ 2 ∂2 iM k Ω k2 2
= ∑ M 2− − ( )
Qk − Qk′2 ρ
k ∂Qk ∂Qk′
2
dt k =1 2
λkk ∂ ∂
− ( Qk − Qk′ ) − ρ
2 ∂Q
k ∂Q ′
k
2
λ ∂ ∂
+ ∑ kk ( Qk + Qk′ ) + + 2 ρ
2
k =1 ∂Qk ∂Qk′
∂ ∂
2
DPk Pk
ρ − 2 ( Qk − Qk′ ) ρ
2
+ DQk Qk +
∂Qk ∂Ql′
2 iDQk Pl ∂ ∂
+ ∑ − ( Qk − Qk′ ) + ρ
k ≠l =
1 ∂Q
l ∂Ql′
∂ ∂ ∂ ∂
+ DQk Ql + + ρ
∂Qk ∂Qk′ ∂Ql ∂Ql′ (39)
2 D
− ∑ k2 l ( Qk − Qk′ )( Ql − Ql′ ) ρ
PP
k ≠l =1
2 2iD ∂ ∂
−∑
Qk Pk
( Qk − Qk′ ) + ρ
k =1 ∂Qk ∂Qk′
Following [21], we can also transform the master Equation (18) for the density
operator into the following Fokker-Planck-type equation satisfied by the Wigner
distribution function W ( Q1 , Q2 , P1 , P2 , t ) :
∂W 2 P ∂W ∂W ∂ ( PkW ) ∂ ( QkW )
∂t
= ∑ − Mk + M k Ω 2k + λkk +
∂Qk
k =1 k ∂Qk ∂Pk ∂Pk
2
∂ 2W ∂ 2W ∂ 2W
+ ∑ DQk Qk 2 + DPk Pk 2 + 2 DQk Pk (40)
k =1 ∂ Qk ∂ Pk ∂Qk ∂Pk
2 ∂ 2W ∂ 2W ∂ 2W
+ ∑ 2 DQk Ql
∂Qk ∂Ql
+ 2 DPk Pl
∂Pk ∂Pl
+ 2 DQk Pl .
∂Qk ∂Pl
k ≠ l =1
In Equation (40), the first four terms on the right-hand side give a purely uni-
tary evolutions, the second four terms are the dissipative terms and have a
damping effect (exchange of energy with environment), the last twelve terms are
diffusive (noise) terms and produce fluctuation effects in the evaluation of the
system. Moreover DPk Pk promotes diffusion in momentum and generates de-
coherence in coordinate Qk . It reduces the off-diagonal terms, responsible for
correlations between spatially separated pieces of the wave packet. Analogically,
DQk Qk promotes diffusion in coordinate and generates decoherence in momen-
tum Pk . The terms DQk Ql , DQk Pl , DQk Pk , and DPk Pl ( k ≠ l = 1, 2 ) are the
so-called “anomalous diffusion” terms. They promote diffusion in the variables
Pk Pl + Pl Pk , Qk Pl + PQ
l k , Qk Pk + Pk Qk , and Qk Ql + Ql Qk respectively. Just like
the other diffusion terms but they do not generate decoherence [5]. To solve
Equation (40), we have to first recall that the diffusion coefficients DQ1Q2 , DQ1P2 ,
DQ2 P1 , DP1P2 are well known and are generally zero for uncoupled oscillators. In
addition, introducing the matrix notation Qk Ql = Gkl , Pk Pl = Ekl and
det ( 2πσ ( t ) ) 2
Q1 + Q1′
2 σQ Q − Q1′
where R = , Q = 1 , r = 1 , H = G −1 F T and
Q2 + Q2 ′ σ Q − Q ′
2Q 2 2
2
σ P P 0 σ Q1P1 0
β= E − FG −1 F T , with E = 1 1 , F = , and
0 σ P2 P2 0 σ Q2 P2
σ Q1Q1 0
G= , y = ( Q1 , Q2 , P1 , P2 ) . n ( t ) = (σ Q1 , σ Q2 , σ P1 , σ P2 ) denotes the
0 σ Q2Q2
expectation vector, and σ ( t ) is the dispersion (correlation) matrix defined by:
G F
σ (t ) = . (43)
F E
Thus, in the case of a thermal bath and when the asymptotic state is a Gibbs
state and considering Equation (38), we obtain the following steady state solu-
tion:
Q1Q2 ρ ( ∞ ) Q1′Q2′ =
ρ1 ρ 2
M 1Ω1 − M 1Ω1 ( Q1 + Q1′ )2 2
= exp + coth ε1 ( Q1 − Q1′ ) (44)
π coth ε1 4 coth ε1
M 2 Ω2 − M Ω ( Q2 + Q2′ )2
2
× exp 2 2
+ coth ε 2 ( Q2 − Q2′ )
π coth ε 2 4 coth ε 2
and
W∞ = W1W2
1 1 2 1
= exp − M 1Ω1Q1 + P12 (45)
π coth ε1 coth ε1 M 1Ω1
1 1 2 1
× exp − M 2 Ω 2 Q2 + P22 ,
π coth ε 2 coth ε 2 M 2 Ω 2
Ω k
with ε k = . We observe that, all stationary solutions to the evolution equa-
kT
tions obtained in the long time limit are possible as a result of a balance between
the wave packet spreading induced by the Hamiltonian and the localizing effect
of the Lindblad operators. Figure 1 and Figure 2 depict the behavior of the den-
sity matrix and the Wigner distribution functions respectively. It can be ob-
served that both exhibit a Gaussian behavior with the amplitudes which strongly
depend on the NC parameters. The diagonal elements of the density matrix (i.e.
Qk = Qkp , k = 1, 2 ) represent the probability of finding the system in this posi-
tion, while the off-diagonal elements introduce the correlation in the density
matrix between the points Qk and Qkp , k = 1, 2 .
(a) (b)
Figure 1. Density matrix ρ ∞ with respect to the coordinates in the Q1-direction (Figure
1(a)) and in the Q2-direction (Figure 1(b)) respectively.
(a) (b)
Figure 2. Wigner function W∞ with respect to both the coordinates and momentum in
the Q1-direction (Figure 2(a)) and in the Q2-direction (Figure 2(b)) respectively.
DP P
the terms of Equation (39) containing the quantity − k2 k dominate on others.
To kindly appreciate its effects, let us rewrite an explicit solution of the master
Equation (Equation (39)) approximating the right hand side by the dominating
terms. In that case, the following time depend evolution of the density matrix is
obtained:
Q1Q2 ρ ( t ) Q1′Q2′
DP P DP P 2
Q1Q2 ρ ( 0 ) Q1′Q2′ exp − 12 1 ( Q1 − Q1′ ) + 22 2 ( Q2 − Q2′ ) t
2
= (46)
1
= Q1Q2 ρ ( 0 ) Q1′Q2′ exp − t .
tdeco
( Qk − Qk′ )
2
where the quantities are in order of the initial dispersion in coordi-
nates σ Qk Qk ( 0 ) , in the particular case of thermal bath (i.e. at zero temperature).
Considering Equation (37), Equation (48) becomes:
4
tdeco = , (49)
λ11δ coth ε1 + λ22δ coth ε 2
2kT 1 2kT 1
τ1
with= = τ2 =
and= . The quantity tdeco defines the so-called
Ω1 ε1 Ω 2 ε 2
decoherence time scale which as we previously mentioned defines the particular
time that the system initially quantum, starts behaving classically (i.e. the partic-
ular time that the system starts losing its quantum properties). Based on the
above analytical expressions, we have numerically simulated the decoherence
time scale as function of the temperature on one hand and as function of the
cutting frequency (magnetic field) on the other hand. One can easily observe on
Figure 3(a) that, the NCty effects present significant impacts on this time scale,
and thus, considerably affect decoherence in the system in the sense that, the
decoherence time scale increases as the NC parameters increase. However, this
time scale decreases asymptotically in the case of NC phase-space as well as in
the commutative phase-space case, and similar effects are also observed as in the
case of total energy since the NCty effects are not more observable for high tem-
perature. It is important to mention that, the NCty effects contribution to deco-
herence for zero temperature coincides with that of the commutative space since
the system receives very small feedback from the environment at this tempera-
ture. We can conclude that, even if the NCty effects increase decoherence in the
system at high temperature, the system losses its coherence properties and be-
comes decoherent, because the decoherence time goes to zero for high tempera-
ture. Moreover, it is observed in Figure 3(b) that, the decoherence time scale
increases as the magnetic field increases, in the case of NC phase-space as well as
in the case of commutative phase-space, but very slowly in the latest case. Under
lower magnetic field effects, the decoherence time scale is different from zero
implying that even at this temperature, the system still receives feedback from its
environment. In addition, the decoherence time scale increases asymptotically to
a fixed value in the case of NC phase-space as well as in the case of commutative
phase-space. This constant value defines the time during which the system loses
its coherence properties and becomes decoherent (the system exhibits classical
behaviour). The above results are confirmed by Figure 4, plotting the decohe-
rence time scale versus simultaneously the NC parameters in the x- and
y-directions, where we can easily observe that it increases with both parameters.
(a) (b)
Since the expectation and variances values of the coordinates and momenta
decrease exponentially in time, the energy is dissipated to a minimum value.
Assuming that the conditions (38) hold, the minimum value of the total energy
of the system is given by the following expression:
Ω Ω1 Ω 2 Ω 2
Emin = E ( ∞ )= 1 coth + coth . (51)
2 2kT 2 2kT
At zero temperature, i.e. T = 0 , the minimal energy becomes
Ω Ω
Emin 1 + 2 , which corresponds to the energy of the system in its
=
2 2
ground state. In the particular case where ω1 = ω2 implying Ω1 =Ω 2 =Ω and
θ= η= 0 , the minimum energy of the ground state can be reduced to
′ = Ω , and we recover the ground state energy of the usual harmonic oscil-
Emin
lator in the commutative case. Figure 5 depicts therefore, the energy with re-
spect to the cutting frequency on one hand and with respect to the real temper-
ature on the other hand both for different values of NC phase-space parameters.
It can be observed that, for fixed values of the temperature T (the thermal bath
temperature), the total energy increases quickly with the magnetic field (cutting
frequency), and that under lower magnetic field effects, the total energy for
(a) (b)
different NC parameter values coincides. But under intense magnetic field ef-
fects, the NCty structure of the phase-space considerably affects the system since
the energy increase with these effects. In addition, it is observed that, for ex-
tremely low temperature, the total energy of the system increase significantly
with the NCty effects, however as the temperature increases, the energy increases
too and for high temperature, the energy of the system in NC phase-space coin-
cides with that of the system in commutative space (blue solid curve). This
represents the fact that, when the temperature increases in the system, the
coupling between the particle and the bath becomes considerably strong, the
whole system takes a large dimension and then, the effects of quantum gravity
manifest in NC phases-space are not more detectable. Similar results were found
by Tchoffo et al. [26] in the case of Brownian particle in NC space. Analogically
to the decoherence time scale, Figure 6 plots the energy of the system with re-
spect to simultaneously the NC parameters in the x- and y-directions, where we
can easily observe that it increases with both parameters, confirming the above
results.
6. Conclusions
In this paper, we studied decoherence of a damped anisotropic harmonic oscil-
lator under magnetic field effects in two-dimensional NC phase-space. For this
reason, the evolution of the system was studied within the framework of the
Lindblad ME theory for open quantum systems, considering the general case of
an environment consisting of a thermal bath at arbitrary temperature. Based on
the above mentioned theory, the damping of the expectation values of coordi-
nates and momenta was evaluated as functions of time. Then, from the ME of an
NC damped anisotropic oscillator approach, the time evolution of the density
matrix and the Wigner function were derived systematically. It turned out by
solving these equations that:
The solutions follow Gaussian distribution. Moreover, from the expectation
values and the variances of coordinate and momenta, the total energy of the
system was evaluated.
The total energy increases significantly with the NCty effects, and as the
temperature increases, the energy increases too. However, for high tempera-
ture, the energy of the system in NC phase-space coincides with that of the
system in commutative space. Similar effects were observed with the mag-
netic field effects.
In addition, the decoherence time scale was analytically derived, and its si-
mulation proved that, the NCty effects present significant impact on this
time scale, and thus, considerably affect decoherence in the system in the
sense that, the decoherence time scale increases as the NC parameters in-
crease. It turned out therefore that, the decoherence time scale was improved
by NCty effects.
The decoherence time was found to have a similar scale as the time after
which statistic fluctuations become comparable with quantum fluctuations as
expected, and the values of the scales become closer with the growth of tem-
perature, magnetic field, and NCty effects.
Conflicts of Interest
The authors declare no conflicts of interest regarding the publication of this pa-
per.
References
[1] Weiss, U. (2012) Quantum Dissipative Systems. World Scientific, Singapore.
https://doi.org/10.1142/8334
[2] Hu, B. and Zhang, Y. (1993) Squeezed States and Uncertainty Relation at Finite
Temperature. Modern Physics Letters A, 8, 3575-3584.
https://doi.org/10.1142/S0217732393002312
[3] Breuer, H.-P. and Petruccione, F. (2003) Concepts and Methods in the Theory of
Open Quantum Systems. In: Irreversible Quantum Dynamics, Springer, Berlin,
65-79. https://doi.org/10.1007/3-540-44874-8_4
[4] Zurek, W.H. (1982) Environment-Induced Superselection Rules. Physical Review D,
26, 1862. https://doi.org/10.1103/PhysRevD.26.1862
[5] Isar, A., Sandulescu, A. and Scheid, W. (2000) Dissipative Tunneling through a Pa-
rabolic Potential in the Lindblad Theory of Open Quantum Systems. The European
Physical Journal D, 12, 3-10. https://doi.org/10.1007/s100530070035
[6] Lampo, A., Lim, S.H., Wehr, J., Massignan, P. and Lewenstein, M. (2016) Lindblad
Model of Quantum Brownian Motion. Physical Review A, 94, Article ID: 042123.
https://doi.org/10.1103/PhysRevA.94.042123
[7] Griessner, A., Daley, A., Clark, S., Jaksch, D. and Zoller, P. (2006) Dark-State Cool-
ing of Atoms by Superfluid Immersion. Physical Review Letters, 97, Article ID:
220403. https://doi.org/10.1103/PhysRevLett.97.220403
[8] Diehl, S., Yi, W., Daley, A. and Zoller, P. (2010) Dissipation-Induced d-Wave Pair-
ing of Fermionic Atoms in an Optical Lattice. Physical Review Letters, 105, Article
ID: 227001. https://doi.org/10.1103/PhysRevLett.105.227001
[9] Khiari, L., Boudjedaa, T., Makhlouf, A. and Meftah, M. (2019) Coupled Oscillators
in Non-Commutative Phase Space: Path Integral Approach. European Physical
Journal, 134, Article No. 396. https://doi.org/10.1140/epjp/i2019-12770-3
[10] Breuer, H.-P. and Petruccione, F. (2000) Radiation Damping and Decoherence in
Quantum Electrodynamics. In: Relativistic Quantum Measurement and Decohe-
rence, Springer, Berlin, 31-65. https://doi.org/10.1007/3-540-45369-5_3
[11] Zurek, W.H. and Paz, J.P. (1995) Quantum Chaos: A Decoherent Definition. Phy-
sica D, 83, 300-308. https://doi.org/10.1016/0167-2789(94)00271-Q
[12] Paz, J.P. and Zurek, W.H. (1993) Environment-Induced Decoherence, Classicality,
and Consistency of Quantum Histories. Physical Review D, 48, 2728.
https://doi.org/10.1103/PhysRevD.48.2728
[13] Lombardo, F.C. and Nacir, D.L. (2005) Decoherence during Inflation: The Genera-
tion of Classical Inhomogeneities. Physical Review D, 72, Article ID: 063506.
https://doi.org/10.1103/PhysRevD.72.063506
[14] Bai, X.-F., Xin, W., Liu, X.-X., et al. (2020) Asymmetric Gaussian Confinement Po-
tential and Decoherence Effect on Polaron in Quantum Disk with Electromagnetic
Field. European Physical Journal, 135, 321.
https://doi.org/10.1140/epjp/s13360-020-00321-y
[15] Isar, A. and Scheid, W. (2007) Quantum Decoherence and Classical Correlations of
the Harmonic Oscillator in the Lindblad Theory. Physica A: Statistical Mechanics
and Its Applications, 373, 298-312. https://doi.org/10.1016/j.physa.2006.04.065
[16] Krzywicki, A. (1993) Coherence and Decoherence in Radiation Off Colliding Heavy
Ions. Physical Review D, 48, 5190. https://doi.org/10.1103/PhysRevD.48.5190
[17] Nielsen, M.A. and Chuang, I. (2002) Quantum Computation and Quantum Informa-
tion. American Journal of Physics, 70, 558. https://doi.org/10.1119/1.1463744
[18] Isar, A., Sandulescu, A. and Scheid, W. (1999) Purity and Decoherence in the
Theory of a Damped Harmonic Oscillator. Physical Review E, 60, 6371.
https://doi.org/10.1103/PhysRevE.60.6371
[19] Davies, E.B. (1976) Quantum Theory of Open Systems. Academic Press, Cam-
bridge.
[20] Isar, A. (1999) Uncertainty, Entropy and Decoherence of the Damped Harmonic
Oscillator in the Lindblad Theory of Open Quantum Systems. Fortschritte der Phy-
sik: Progress of Physics, 47, 855-879.
https://doi.org/10.1002/(SICI)1521-3978(199909)47:7/8<855::AID-PROP855>3.0.C
O;2-Z
[21] Isar, A., Sandulescu, A., Scutaru, H., Stefanescu, E. and Scheid, W. (1994) Open
Quantum Systems. International Journal of Modern Physics E, 3, 635-714.
https://doi.org/10.1142/S0218301394000164
[22] Isar, A., Sandulescu, A. and Scheid, W. (1991) Use of Characteristic Function in
Open Quantum Systems and Charge Equilibrium in Deep Inelastic Reactions.
Journal of Physics G: Nuclear Physics, 17, 385.
https://doi.org/10.1088/0954-3899/17/3/015
[23] Halliwell, J.J. (1989) Decoherence in Quantum Cosmology. Physical Review D, 39,
2912. https://doi.org/10.1103/PhysRevD.39.2912
[24] Dragovich, B. and Dugić, M. (2005) On Decoherence in Noncommutative Plane
with Perpendicular Magnetic Field. Journal of Physics A: Mathematical and Gener-
al, 38, 6603. https://doi.org/10.1088/0305-4470/38/29/014
[25] Ghorashi, S. and Harouni, M.B. (2013) Decoherence of Quantum Brownian Motion
in Noncommutative Space. Physics Letters A, 377, 952-956.
https://doi.org/10.1016/j.physleta.2013.02.019
[26] Tchoffo, M., Kuetche, J.C.N., Fouokeng, G.C., Afuoti, N.E. and Fai, L.C. (2014) Ki-
Gaussian States in the Two-Reservoir Model. Romanian Reports in Physics, 68, 19.
[42] Suciu, S. and Isar, A. (2016) Quantum Coherence of Two-Mode Systems in a Ther-
mal Environment. Romanian Journal of Physics, 61, 1474.
[43] Isar, A. (2018) Generation of Quantum Steering in Gaussian Open Systems. Roma-
nian Journal of Physics, 63, 108.
[44] Isar, A., Scheid, W. and Sandulescu, A. (1991) Quasiprobability Distributions for
Open Quantum Systems within the Lindblad Theory. Journal of Mathematical
Physics, 32, 2128. https://doi.org/10.1063/1.529185
[45] Isar, A. (2009) Entanglement Generation and Evolution in Open Quantum Systems.
Open Systems & Information Dynamics, 16, 205-219.
https://doi.org/10.1142/S1230161209000153
[46] Schrödinger, E. (1930) Zum heisenbergschen unschärfeprinzip. Akademie der Wis-
senschaften.
[47] Robertson, H.P. (1929) The Uncertainty Principle. Physical Review, 34, 163.
https://doi.org/10.1103/PhysRev.34.163
[48] Mihaescu, T. and Isar, A. (2017) Gaussian Quantum Steering of Two Bosonic Mod-
es in a Thermal Environment. Quantum, 1, 3.
[49] Isar, A. (2013) Quantum Correlations of Two-Mode Gaussian Systems in a Thermal
Environment. Physica Scripta, 2013, Article ID: 014035.
https://doi.org/10.1088/0031-8949/2013/T153/014035
[50] Dodonov, V. and Manko, O. (1985) Quantum Damped Oscillator in a Magnetic
Field. Physica A: Statistical Mechanics and Its Applications, 130, 353-366.
https://doi.org/10.1016/0378-4371(85)90111-6
[51] Agarwal, G. (1971) Brownian Motion of a Quantum Oscillator. Physical Review A,
4, 739. https://doi.org/10.1103/PhysRevA.4.739
[52] Rajagopal, A. and Rendell, R. (2001) Decoherence, Correlation, and Entanglement
in a Pair of Coupled Quantum Dissipative Oscillators. Physical Review A, 63, Ar-
ticle ID: 022116. https://doi.org/10.1103/PhysRevA.63.022116