You are on page 1of 16

Phys. Scr. 96 (2021) 085705 https://doi.org/10.

1088/1402-4896/ac0273

PAPER

Decoherence dynamics of a charged particle within a non-


RECEIVED
2 March 2021
demolition type interaction in non-commutative phase-space
REVISED
7 May 2021
ACCEPTED FOR PUBLICATION
Yiande Deuto Germain1, Azangue Koumetio Armel1, Alain Giresse Tene1 , Nsangou Isofa1 and
18 May 2021 Martin Tchoffo1,2
1
PUBLISHED Research Unit of Condensed Matter, Electronics and Signal Processing, Department of Physics, University of Dschang, PO Box 67
27 May 2021 Dschang, Cameroon
2
Centre d’Etudes et de Recherches en Agronomie et en Biodiversité, FASA, Université de Dschang, PO Box 222 Dschang, Cameroon
E-mail: mtchoffo2000@yahoo.fr

Keywords: decoherence, quantum non-demolition, non-commutative phase-space

Abstract
This paper studies decoherence without dissipation of charged magneto-oscillator in the framework
of quantum non-demolition type interaction in non-commutative phase-space. The master equation
is derived considering the non-commutativity effects of a bath of harmonic oscillators to study the
dynamics of such a system, and its possible exact solution is presented. By analyzing this solution, it
turns out that the process involving decoherence without energy dissipation can be observed explicitly.
In addition, the decoherence factor and the measure of coherence via linear entropy dynamic are
deduced for two types of reservoir, namely the ohmic and super-ohmic reservoirs at low and high
temperature limits. Numerical results obtained show that the coherence is better preserved in the
system when non-commutativity effects are taken into account at low temperature, while the inverse
phenomenon is observed at high temperature. Moreover, by kindly adjusting the non-commutative
parameters, it is possible to improve the coherence time scale of the system. Another interesting result
can be observed from the system’s coherence time scale, which is very sensitive to the magnetic field
and thus adding to non-commutative parameters, it may be useful to control decoherence in the
system.

1. Introduction

The theory of non-commutative geometry has attracted extensive attention in recent studies, although it
presents a long standing story since the discovery of non-commutative effects in low-energy effective of string/
M-theory, quantum field theory and quantum mechanics [1–3]. In addition, it was suggested that space-time
non-commutativity might be assimilated to quantum effect of gravity [4, 5]. A more important feature of non-
commutative geometry is its direct connection to the breaking of Lorentz invariance. Various recent research
works have been done in non-commutative field theory, including its non-commutative extension of the
Standard Model [6]. Given that quantum mechanics can be interpreted from the one-particle non-relativistic
quantum field theory point of view, it is important to study its non-commutative aspects, particularly in open
quantum systems, which typically has a large number of degree of freedom [7].
Indeed, open quantum systems theory addresses the problem of dephasing and damping by its assertion that
no system in the nature can never be truly isolated; every system actually interacts with its environment. The
interest of such systems is evident in our everyday life phenomena [8]. In the past two decades, the research on
open quantum systems received a growing interest due to its various application in quantum optics [9],
quantum information devices [10] and quantum information theory [11]. The study of open quantum systems
was introduced by Feynmann-Veron [12] and later expanded by Caldeira and Leggett [13]. For example we
denote the quantum Brownian motion, which is a system where the environment is assumed to be a bath of non-
interacting harmonic oscillators [14, 15]. Coupling a system to its environment consequently induce firstly, an
irreversibly exchange of energy between the system of interest with its large environment leading to dissipation

© 2021 IOP Publishing Ltd


Phys. Scr. 96 (2021) 085705 Y D Germain et al

[16–19]. Secondly, the entanglement of the system and its environment destroys the coherent superposition of
quantum states giving rise to decoherence [18, 20, 21, 21]. However, extensive studies have been conducted
during the past two decades to observe the effects of the latest mentioned phenomenon for two main reasons,
including the fundamental quantum-to-classical transition concept [17, 21, 22] on one hand, and on the other
hand, the phenomenon of decoherence presents a significant impact on quantum devices technology [10, 11],
such as quantum computer and quantum cryptography, which are related to the preservation of quantum
coherence. For this reason, the superposition should not be confused with mixture, otherwise we lose the unique
advantage offered by quantum mechanics. In addition, the question of energy-preserving measurements that
leads to decoherence without dissipation for such systems when manipulating quantum states of matter and
quantum information processing tasks have triggered a renewed demand to really understand and control the
environmental impact on such systems [23]. This may be achieved by considering a particular type of interaction
between the system and its environment so-called quantum non-demolition (QND) [24–29]. A QND type
interaction has been proved to present significant impact in quantum open systems and particularly in
decoherence process due to its properties. In this idea, Anirban and Gautam [30], studied the decoherence
without dissipation in a system under fermionic bath interaction. Considering as their system, a harmonic
oscillator coupled to a fermionic environment via a QND type interaction, they were able to provide an exact
solution describing its dynamics. Banerjee et al [31] also studied the decoherence without dissipation dynamics
in a squeezed thermal bath. They obtained a general master equation describing the evolution of such systems
influenced by a squeezed thermal bath of harmonic oscillator. It results from the above that decoherence in
QND type interaction quantum open system may be of great interest, and still explorable. Most of these research
works were achieved only in the commutative space (i.e. the Hilbert space momenta and coordinates are
commutative).
Given the interest and large advantage that provide non-commutative geometry, we intend in this work to
study decoherence in a quantum magneto-oscillator that interacts with a QND bath made of independent
harmonic oscillators considering the Hilbert phase-space as non-commutative. In a very recent research paper,
we discovered that in addition to environmental effects, the decoherence time scale (i.e. the time interval in
which a system exhibits quantum behavior) increases significantly with non-commutative phase-space effects
compared to the case of commutative phase-space. It becomes thus, interesting to combine this effects to a QND
type interaction to observe the evolution of this time interval, which is the main goal of this research paper. To
achieve this goal, we therefore structured the paper as follows: section 2 presents the theoretical model and the
formalism of non-commutative phase-space geometry. Therein, our open system model for a charged magneto-
oscillator under the influence QND type interaction in non-commutative is developed. The bath correlation
function is presented in section 3. In section 4, the time evolution operator and master equation of the system
interacting with its environment are obtained. Therein, the dynamic of the evolution operator in interaction
picture is studied, following by the derivation of an exact solution of the master equation associated to the
previously mentioned system. Section 5 provides the analytical and numerical results, where we quantitatively
analyze the dynamics of decoherence (respectively of coherence) from the decoherence factor present in the
master equation point of view on one hand, and from the linear entropy on the other hand. We thus, end the
work with concluding remarks presented in section 6.

2. Model Hamiltonian of a charged particle interacting with a non-demolition


environment in non-commutative phase-space

2.1. Non-commutativitity formalism


In this work, both the momentum and coordinate space are assumed to be non-commutative. In the usual
commutative phase-space, coordinates and momenta satisfy the following relations of commutation:
[x i , x j ] = 0, [ pi , pj ] = 0, [x i , pj ] = idij . (1)

However, at a very tiny scale (or string scale), not only coordinate-momentum is non-commutative, but also
coordinate-coordinate and momentum-momentum may not commute too. Therefore, the non-commutative
phase-space in quantum mechanics is a Hilbert space where coordinate and momentum operators satisfy the
following relations of commutation:
[xˆ i , xˆ j ] = iq ij , [ pˆi , pˆj ] = ih ij , [xˆ j , pˆj ] = i dij , i , j = 1, 2, (2)
qh ⎞
where  is the effective Plank constant given by  =  ⎛1 + , xˆ i ( j) and pˆi ( j) the coordinate and
⎝ 4 2 ⎠
momentum operators in non-commutative phase-space, respectively. Here θ and η are constants and represent
the non-commutative parameters for coordinates and momentum, respectively. Moreover, θ and η have
the dimensions of (length)2 and (momentum)2, respectively. The mapping between the non-commutative

2
Phys. Scr. 96 (2021) 085705 Y D Germain et al

phase-space (x̂ i , p̂i ) and the commutative ones (xi, pi), is obtained using the Bopp-Shift linear transformation
[32–35]:
q
xˆ i = x i - eij p,
2 j i , j = 1, 2. (3)
h
pˆi = pi + eij x j ,
2

2.2. Model Hamiltonian


Let’s consider a quantum charged moving particle with mass m in the presence of an external magnetic field, and
confined by a square potential, then its Hamiltonian can be given by [36, 37]:
2
1 ⎛ eAˆ ⎞
Hˆ S = ⎜pˆ - ⎟ + V (x
ˆ1, xˆ2) , (4)
2m ⎝ c ⎠
with
1 cos (4f)
V (xˆ1, xˆ2) = mw 20 (xˆ12 + xˆ 22)(1 + ), (5)
2 5
characterized by its confinement frequency ω0 and the polar angle between the position vector and x1, f. Here,
(xˆ1, xˆ2) and ( pˆ1 , pˆ2 ) are respectively the non-commutative coordinates and momentum. Considering the
B
symetric gauge Aˆ = 2 (-xˆ2, xˆ1, 0) and the Bopp-Shift transformation given by equation (3), the Hamiltonian
can be rewritten as:
1 K
Hˆ S = ( p 2 + p22 ) + (x12 + x 22) - l (x1 p2 - x2 p1) , (6)
2M 1 2
m w2
where M = 2 2 ( )(
w w
) (
, l = ⎡ 1 + 2wc wh + 2c + w0 1 + 5
⎣ q q
cos (4f )
) ⎤,


(
⎡ 1 + wc ) ( )(
2wq
+
w0
wq
1 +
cos (4f )
) 5


2 2 2 h
W=

(
⎡ 1 + wc
2wq ) + ( ) (1 +
w0
wq
cos (4f )
5
cos (4f )
( w

) 2
)][w 20 (1 + 5 ) + wh + 2c ⎤ , wq = mq , wh = 2m ,
K eB
W = M , wc = mc is the cyclotron frequency, ωθ and ωη are the non-commutative frequency due to non-
commutativity effects. Ω depends on θ and η, and we can easily observe that when η = 0, the above Hamiltonian
reduces to the case where only the space is non-commutative, while if in addition θ = 0, then its results to the
usual 2-dimensional harmonic oscillator on commutative phase-space.
We note that the Hamiltonian (6) contains an additional term comparing to the usual harmonic oscillator
Hamiltonian with commuting operators. This is because of the non-commutativity aspect of the phase-space,
and can be assumed as the orbital angular momentum along the z-direction. In addition, the non-commutative
system has an effective mass which coincides with the actual mass m for θ = 0. Further, in the commutative case
(i.e. θ = 0, η = 0), and in the absence of magnetic field (i.e. B = 0), the last term in the Hamiltonian vanishes. In
this situation, the system reduces to a simple two dimensional harmonic oscillator with energy En1 n2 =
w¢ (n1 + n2 + 1). Otherwise (i.e. θ ≠ 0, η ≠ 0 and B ≠ 0), the system is assumed to be an harmonic oscillator in
a Hilbert phase-space with non-commutative structure, and presenting non-diagonal Hamiltonian due to the
last term in equation (6). In order to diagonalize this Hamiltonian, let’s introduce two new operators, a1 and a2
such that the position and momentum variables are rewritten as [38, 39]
W W
x1 = (a1 - a2 + a1+ - a 2+) , x2 = i ( - a1 - a2 + a1+ + a 2+) ,
4K 4K
M W M W
p1 = i (a2 - a1 + a1+ - a 2+) , p2 = - (a1 + a2 + a1+ + a 2+) , (7)
4K 4K
where ai and aj+ satisfy the usual commutation relations [aj+, a i ] = -dij , i, j = 1, 2. Considering equation (7),
the Hamiltonian becomes
1 1
HS =  W1⎛a1+a1 + ⎞ +  W2 ⎛a 2+a2 + ⎞ , (8)
⎝ 2⎠ ⎝ 2⎠
with the frequencies
W1 = W + l , W2 = W - l. (9)
It can be observed from the Hamiltonian (8) that, the non-commutativity structure of the phase space introduces
anisotropy in the system, given that the frequencies are different for the x1 − and x2 − directions. In this case, the
energy of the system strongly depends on the structure of the space and is given by:

3
Phys. Scr. 96 (2021) 085705 Y D Germain et al

1 1
En1 n2 =  W1⎛n1 + ⎞ +  W2 ⎛n2 + ⎞ , (10)
⎝ 2⎠ ⎝ 2⎠

where n1 and n2 are the quantum numbers (n1, n2 = 0, 1, 2, L), Ω1 and Ω2 the frequencies defined by
equation (9). We assume that the interaction system-environment leads only to decoherence, but not to energy
dissipation [26, 31], which has been already introduced for a two-level atom in [40–42]. The total Hamiltonian
of such system interacting with a bath of harmonic oscillators, via a QND type interaction [30, 31, 43] is defined
by:

H = HS + HSB + HB , (11)

where HS and HB are respectively the Hamiltonian of non-commutative magneto-oscillator and that of the bath;
HSB the interaction Hamiltonian between the system and the bath. In this work, the bath is modeled as an infinite
number of harmonic oscillators, considered as two independent heat bath in the x1- and x2-directions. Then, the
bath Hamiltonian is given by:
¥ 2
1
HB = å å wn,j ⎛bn+,jbn,j + 2⎠
⎞, (12)
n= 0 j= 1 ⎝

2 2
⎛1 + mn wn q )(w n2 + h
( ) ⎞  ( h + mn wn q ) are the non-commutative frequencies of
2
where wn, j =

( 2 ) 2m n 
⎠ 2m n  2
+
n oscillators in the x1 and x2 directions respectively. bn, j and bn, j (j = 1, 2) are the annihilation and creation
th

bosonic operators of the heat bath in the x1 and x2 directions respectively, and satisfy the commutation relations
[bn, j , bn+, j ] = dnn ¢ djj ¢ . Moreover, by considering the interaction part HSB as QND coupling, one has [HS, HSB] = 0.
This means that in our fully quantified model, there is no energy transfer from the system to the environment. It
implies that HSB is a constant of motion, which can be generated by HS and taken as some function of HS as:
¥ 2
HSB = å å h j (gn,j bn,j + gn*,jbn+,j ) , (13)
n= 0 j= 1

with hj =  Wj aj+a j +
( 1
2 ).
3. Bath correlation function evaluation

Let K (t - t ¢) be the correlation function quantifying the interaction between the system and the bath. It is
defined as the trace of the system over the bath and given by:
K (t - t ¢) = TrB [rB exp (iHB t ) B exp ( - iHB t ) exp (iHB t ¢) B exp (iHB t ¢)] ,
= TrB [rB exp (iHB (t - t ¢)) B exp (iHB t ) B] ,
= áB(t ) Bñ. (14)
The correlation function only depends on the quantity t = t - t ¢, for any time-independent form of the
interaction. The state of the bath is considered as a canonical Gibbs state given by:
e -bHB
rB = , (15)
Z
with Z = trB [e-bHB ]. From the interaction Hamiltonian (13), we can note that
¥
Bj = å (gn,j bn,j + gn*,jbn+,j ). (16)
n= 0

To evaluate B(t ), let us assume that


e iHB tbn, j e -iHB t = bn, j e -iwn,jt , e iHB tbn+, j e -iHB t = bn+, j e iwn,jt (17)
Therefore,
¥
Bj (t ) = å (gn,j bn,je-iw n, jt + gn*, j bn+, j e iwn,jt ). (18)
n= 0

Considering equations (16) and (18), the correlation function K(τ) can be derived in terms of the bath’s
parameters. However, it is worth noting that only the terms of the form ábn+, jbn, jñ and ábn, jbn+, jñ can possibly give
non-zero diagonal matrix elements, since they conserve the numbers of particles in modes n2 and j2. Using this
idea, it comes that:

4
Phys. Scr. 96 (2021) 085705 Y D Germain et al

¥
K j (t ) = å ∣gn,j∣2 [e-iw n, jt (1 + Nn, j ) + e iwn,jtNn, j ] , j = 1, 2, (19)
n= 0

where we have assumed that ábn+, jbn, jñ = Nn, j and ábn, jbn+, jñ = Nn, j + 1. Since the bath is in thermal equilibrium,
Nn,j is thus, defined by the Bose–Einstein distribution:
1
Nn, j = wn,j . (20)
e KB T - 1
Then the correlation function can be simplified to:
¥ wn, j ⎞
K j (t ) = å ∣gn,j∣2 ⎡⎢cos (wn,jt ) coth ⎛ 2K ⎜ - i sin (wn, jt ) ⎤ ,


(21)
n= 0 ⎣ ⎝ BT ⎠ ⎦
where T denotes the temperature of the bath and KB the Boltzmann constant.
w
Setting nj = å n¥= 0∣gn, j∣2 cos (wn, jt ) coth ( 2Kn,Tj ) and cj = å ¥
n = 0∣gn, j ∣ sin (wn, jt ), therefore the correlation
2
B
function becomes:
K j = n j - icj , (22)
where νj the real part is assimilated to the noise kernels, while χj, the imaginary part is referred to as the
dissipation kernels.

4. Evolution operator and master equation

4.1. Time evolution operator: interaction picture


To solve the dynamics of our model described by equations (8), (12) and (13), let us first consider the interaction
picture transformation. That is, the Hamiltonian (13) becomes
HSB (t ) = e iH0 t HSB e-iH0 t , (23)
where H0 = HS + HB is the free Hamiltonian in the absence of interaction. Then, considering equation (23) we
have
¥ 2
HSB (t ) = å å h j (gn*,jbn,je-iw n, jt + gn, j bn+, j e iwn,jt ). (24)
n= 0 j= 1

By defining U0 (t ) = exp (-iH0 t ), one has:


USB (t ) = U 0+(t ) U (t ) , (25)
with U(t), the usual time evolution operator depicting a unitary evolution of the total system. The time-
derivative of equation (25) shows that USB(t) verifies the following differential equation:
dUSB (t )
= - iHSB (t ) USB (t ) , (26)
dt
where its solution can be set as [44, 45]:
¥
USB (t ) = exp { å Ck (t )}, (27)
k=1

Thus, the solution of equation (26) takes the following form (details can be found in appendix A):
2 ¥
USB (t ) =  exp [ å h j (bn+, j an, j - bn, j a*n, j ) - ih j2 D j ]. (28)
j=1 n= 0

Further, the exact unitary time evolution operator according to equation (25) is given by:
2
U (t ) =  e-i (E m j¢ - E n j¢ ) t e -iHB t USB (t ). (29)
j=1

4.2. Master equation


The dynamical behavior of a magneto-oscillator can be completely described by its density matrix state. So, by
using the master equation approach, one is able to determine the time evolution of the density matrix and
thereby the complete behavior of the system. In the presence of interactions between the system and the
environment, it may not be possible to fully derive an exact master equation for the system only and
consequently, we use the Born-Markov-approximation. This approach considers the reduced density matrix
of the system in the limit of a weak interaction with dissipative environment. Thus, by considering the

5
Phys. Scr. 96 (2021) 085705 Y D Germain et al

system-environment coupling Hamiltonian as follows:


2
HSB (t ) = å h j Ä B j (t ) , (30)
j=1

where the quantities hj contain operators described in the Hilbert space of the system defined by
hj =  Wj aj+a j + 2 and Bj(t) the operators in the environmental Hilbert space given by equation (16). Then,
( )
1

assuming a weak system-environment interaction, the total initial state defined by the product state
ρ(0) = ρB(0)ρS(0), where ρB(0) and ρS(0) are the density matrix of the environment and the system respectively,
one has [46, 47]:
drS (t ) i 2 t

dt
= [rS, HS] -

å ò0 dt ¢ ([h j , h˜ j (t ¢ - t ) rS (t )] K j (t - t ¢)
j=1

- [h j , rS (t ) h˜ j (t ¢ - t )] K j (t - t ¢)*) , (31)
where h˜j (t , t ¢) = US (t , t ¢) hj US+(t ,
t ¢), with US (t , t ¢) a unitary operator describing the evolution of the system
and Kj (t - t ¢) the bath correlation function given by equation (19). In addition, the physical meaning of the
different terms of equation (31) are as follows: the first term on the right hand side introduces the coherent
evolution of the system, while the remaining terms introduce the coupling with the bath. Considering
equations (8) and (21), the master equation (31) becomes [30, 31]:
drS 2 2
2 
= - i å W j [a +
j a j , rs ] + i å W j D j [((a j a j ) + a j a j ) , rs ]
2 + 2 +
dt j=1 j=1
2
-  2 å W2j G j [(a + + + +
j a j ) rs - 2a j a j rS a j a j + rs (a j a j ) ] ,
2 2 (32)
j=1

where
t ¥ ∣gn, j ∣2
d
dt
Dj = ò0 cj (t - t ¢) dt ¢ = å
n= 0 wn, j
(cos (wn, jt ) - 1) , (33)

and

d t ¥ ∣gn, j ∣2 sin (wn, jt ) wn, j ⎞


dt
Gj = ò0 n j (t - t ¢) dt ¢ = å
n= 0 wn, j
coth ⎛ ⎜

⎝ 2KB T ⎠
. ⎟ (34)

The state of the system alone at anytime t is found by tracing over the reservoir degree of freedom as
rS (t ) = trB (U (t ) r (0) U +(t )) , (35)

with ρ(0) = ρS(0)ρB(0) the corresponding density operator of the total system (system +bath), ρS(0) and ρB(0)
the density operators of the system and the bath respectively. Then the reduced density matrix elements in the
eigenbasis are:
[rS (t )]mj¢ n j¢ = trB [U (t ) r (0) U +(t ) Tmj¢ n j¢] , (36)

where Tmj¢ nj¢ = ∣mj¢ nj¢ñ , ∣nj¢ñ is the eigenstates of hj with eigenvalues E nj¢ . Therefore, we find that the exact solution
of the reduced density matrix of equation (32) is given by (see appendix B for detailed calculations):
2
-i (E mj¢ - E nj¢ ) t - i (Em2 ¢ - En2¢ ) D j (t ) -Gj (t )(E m ¢ - E n ¢ )2
[rS (t )]n ¢ m ¢ = [rS (0)]n ¢ m ¢ e j j e j j , (37)
j=1

where
t s ¥ ∣gn, j ∣2 wn, j ⎞
Gj (t ) = ò0 ds ò0 n j (s - t ¢) dt ¢ = å
n = 0 w n, j
2
(1 - cos (wn, jt )) coth ⎛
⎝ 2KB T ⎠
, ⎜ ⎟ (38)

t s ¥ ∣gn, j ∣2
D j (t ) = ò0 ds ò0 cj (s - t ¢) dt ¢ = å
n= 0 w n2, j
(sin (wn, jt ) - wn, jt ) , (39)

-i (E 2 - E 2 ) D (t )
with 2j = 1 e mj¢ nj¢ j describing the indirect atom-atom interaction due to the common bath. Comparing
the master equation obtained here with that obtained in the case of quantum Brownian motion as in [34, 35, 43],
it turns out that the term responsible of decoherence in the system is the derivative of Γ(t) with respect to time
(G (t )), consequently, Γ(t) is responsible for coherence in the system.

6
Phys. Scr. 96 (2021) 085705 Y D Germain et al

5. Dynamical evolution of decoherence of the magneto-oscillator in non-commutative


phase-space

In this section, the decoherence dynamics is evaluated via the decoherence factor and the linear entropy of a two
dimensional magneto-oscillator interacting with ohmic and super-ohmic type environments in a non-
commutative phase-space. For the reasons of simplicity, let us introduce two parameters α and β defined such
that θ = α × 10−38m2 and η = β × 10−60kg2m2s−2, where θ and η are the non-commutative parameters, whose
values are selected following the works of [48, 49]. It is worth noting that α and β are dimensionless.

5.1. Decoherence factor for two different environments


The result in equation (37) is very interesting in the sense that, it demonstrates that the decoherence is mostly
controlled by the heat bath parameters instead of those of the system. Moreover the non-commutative
parameters and magnetic field content in the eigenvalues E nj¢ of the systems determine the rate while the
function Γj(t) is controlled by the bath coupling type. Thus, the term
2
 e-G (t )(E m j¢ - E n j¢ )
2
r (t ) = j
, (40)
j=1

in equation (37) introduces the decoherence rate. On the other hand, one can realize that Γj(t) is a sum of positive
terms, which means that, to get decoherence (i.e., for this sum to diverge as the time goes to infinity), we need
continuum frequency values and a strong interaction with the bath modes at low frequencies. Therefore, let us
define the environment function, namely the spectral density function in order to perform the sum over all
environmental modes as Jj (w ) = å n¥= 0 gn2, j d (w - wn, j ). This functions contains all relevant information of
environment and the coupling of the system, apart from the temperature. Further, we assume that the baths are
in the thermal equilibrium state with the same value of temperature T. Then, the spectral density is selected as
[50, 51]:
G0 s - wj
J j (w ) = w j e Lc , j = 1, 2 (41)
p
m¢q h
where w1,2 = (1 + aq2 w 2)(w 2 + b 2h )  (bh + aq w 2), (with aq = 2
, bh = 2m ¢ 
), G0 the coupling
constant and Λc the cut-off frequency of the bath. Therefore, we have
¥ +¥
å ∣gn,j∣2 f (wn,j )  ò0 dwJ j (w ) f j (w ). (42)
n= 0

In this work, we focus on two special cases, which include the ohmic (s = 1) and super-ohmic (s = 3 ) spectral
densities, respectively. These two cases describe different physical contexts.

5.1.1. (a) Case of ohmic spectral density


Let’s thus first consider the ohmic case, and evaluate the quantities Δj(t) and Γj(t). Substituting equations (41)
and (42) into equations (38) and (39), we obtain:
G0 +¥ (sin (w j t ) - w j t ) wj
D j (t ) =
p ò0 wj
e - Lc dw , (43)

and
G0 +¥ (1 - cos (w j t )) wj ⎞ - wj
Gj (t ) =
p ò0 wj
coth ⎛ ⎜

⎝ 2KB T ⎠
e Lc dw. ⎟ (44)

In the commutative limit (i.e. θ = η = 0), equations (43) and (44) reduce to:
G0
D1(t ) = D2 (t ) = D (t ) = [tan-1(Lc t ) - Lc t ] , (45)
p
and
G0
G1(t ) = G2 (t ) = G (t ) =
ln (1 + Lc2 t 2) , (46)
2p
respectively, for zero temperature, which confirm the results obtained in [31, 52]. For low temperature (i.e.
Λc ? KBT), equation (44) reduces to [26, 40]:
G0 ⎡ 1
G1(t ) = G2 (t ) = G (t ) = ln (1 + Lc2 t 2) + 2 ln ⎡ sinh (pKB Tt ) ⎤ ⎤. (47)
p ⎣⎢ ⎢ pKB Tt
⎣ ⎥⎥
⎦ ⎦

7
Phys. Scr. 96 (2021) 085705 Y D Germain et al

The first term arises from the quantum vacuum fluctuations while the second is due to thermal ones. The above
G (L t )2
expression reduces to G (t ) = 0 c for t  L- c (quiet regime), where the fluctuations are ineffective in the
1
p
G ln (Lc t )
decoherece process. While it reduces to G (t ) = 0 for L-
c  t  (KB T ) . In this case, the main
1 -1
p
G K Tt
causes of coherence loss (decoherence) are the quantum fluctuations. However, one has G (t ) = 0 B for
p
(KB T )-1  t (thermal regime). In this case, the thermal fluctuations play the major role in eroding the system’s
coherence.
For high temperature, equation (44) reduces to:
+¥ (1 - cos (w j t )) wj
Gj (t ) = G0 ò0 w 2j
e- Lc dw. (48)

Then, in commutative case (θ = η = 0), the previous equation becomes,


G1(t ) = G2 (t ) = G (t ) = G0 [t Lc tan-1(Lc t ) - ln (1 + Lc2 t 2)] , (49)
2G0 KB T
with G0 = p
, which coincides with those obtained in [31, 52].

5.1.2. (b) Case of super-ohmic spectral density


It is also interesting to consider the super-ohmic case (s = 3 in equation (41)). In this particular case, the quantity
Γj(t) that leads decoherence can still be evaluated in three different temperature bands, which include zero-
temperature, low-temperature and high-temperature.

• For zero-temperature, one has:


G0 +¥ wj
Gj (t ) =
p ò0 (1 - cos (w j t )) w j e - Lc dw , j = 1, 2. (50)

• For low temperature, one has:


G0 +¥ wj ⎞ - wj
Gj (t ) =
p ò0 (1 - cos (w j t )) w j coth ⎛ ⎜

⎝ 2KB T ⎠
e Lc dw ,⎟ j = 1, 2. (51)

In the commutative limit (i.e. θ = η = 0), equation (51) becomes [40]

G0 ⎡ K T K T (1 + i L c t ) ⎤ K T (1 - i L c t ) ⎤ ⎞
G1(t ) = G2 (t ) = G (t ) = (KB T )2 ⎜⎛2z ⎡2, B ⎤ - z ⎡2, B - z ⎡2, B ⎟
p ⎣⎢ ⎢ Lc ⎦
⎥ ⎢ Lc ⎥ ⎢ Lc ⎥⎠
⎝ ⎣ ⎣ ⎦ ⎣ ⎦
1 1
+ Lc2 ⎡ + - 2⎤ ⎤ ,

⎣ ( 1 + i L c t ) 2 ( 1 - i L c t ) 2
⎦⎥
⎥ ⎦
(52)
where ζ is the generalized Riemann Zeta function.
• For high temperature, one has:
+¥ wj
Gj (t ) = G0 ò0 (1 - cos (w j t )) e - Lc dw , j = 1, 2, (53)

which reduces to
G0L3c t 2
G1(t ) = G2 (t ) = (54)
1 + (t L c ) 2
in the commutative limit (i.e. θ = η = 0).

In the above expressions, we have as previously mentioned w1,2 = (1 + aq2 w 2)(w 2 + b 2h )  (bh + aq w 2),
h
(with aq = m2¢ q , bh = 2m ¢  ), G0 the coupling constant and G0 = 2G0pKB T . These expressions are helpful in
evaluating the decoherence factor given by equation (40).
It can be observed a rapid decay of the system’s coherence in time, which vanishes after a very short given
period of time when the system is considered in both commutative and non-commutative phase-space. This
means that the suppression of coherence in a system in permanent interaction with a dissipative environment
appears at a very short time scale. However, this decrease is more rapid when the system is found in commutative
phase-space at low temperature, while at high temperature, the inverse phenomenon occurs. This behavior is

8
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 1. Effects of the non-commutative phase-space on the dynamical behavior of decoherence of a magneto-oscillator interacting
with an ohmic reservoir at low temperature represented by equation (40) with Γj defined by equation (44) (figure 1(a)), and at high
temperature represented by equation (40) with Γj defined by equation (48) (figure 1(b)).

Figure 2. Evolution of decoherence factor in a system of a magneto-oscillator interacting with an ohmic reservoir with respect to both
the temperature and time (figure 2(a)) and with respect to both the magnetic field and the cut-off frequency (figure 2(b)) respectively.
Blue color corresponds to α = 0 and β = 0, Black color corresponds to α = 0.55 and β = 0.75, Red color corresponds to α = 0.9 and
β = 0.9.

observed when the system interact with an ohmic (figure 1) and super-ohmic (figure 3) reservoirs. The
oscillatory behavior observed at low temperature traduces the revivals of coherence in the system due to non-
commutativity effects. Similar behaviors are observed in figures 2 and 4 representing the decoherence-causing
term with respect simultaneously to the time and the temperature (figures 2(a) and 4(a)) and with respect to the
magnetic field and the cut-off frequency (figures 2(b) and 4(b)), respectively. As regards to figure 2(b), it can be
observed an exponential decrease of this term with the magnetic field in the commutative case (the blue color
curve), while it decreases very slowly to zero in the non-commutative system (the black and white color curves).
In addition, from those figures, it can be identified no matter if the system is in non-commutative or
commutative phase-space as well, three time regimes of decoherence: (i) the first time regime is when the
decoherence factor tends to an initial value of one. This corresponds to the quiet regime where the fluctuations
are ineffective in the system (decoherence process), (ii) the second phase is when the decoherence factor
decreases, which corresponds to the thermal regime where the quantum vacuum fluctuations are the main cause
of coherence loss and (iii) the third regime is when the decoherence factor vanishes, which also corresponds to
the thermal regime, where thermal fluctuation play the crucial role in suppressing the system coherence. Similar
results were found by Palma et al [40], when analyzing the possibility to control decoherence in systems
interacting with different environment.

9
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 3. Effects of the non-commutative phase-space on the dynamical behavior of decoherence of a magneto-oscillator interacting
with an super-ohmic reservoir at low temperature represented by equation (40) with Γj defined by equation (51) (figure 3(a)), and at
high temperature represented by equation (40) with Γj defined by equation (53) (figure 3(b)).

Moreover, from figures 2(b) and 4(b) it can be observed that in an intense magnetic field effects, the decay of
coherence is more dominated by the thermal fluctuations of the reservoir, while, in low magnetic field effect,
there is an intermediate regime where the vacuum fluctuation dominates, and which corresponds to the thermal
regime. However, for a very tiny magnetic field effects, no decay is observed, thereby this regime represent the
quiet regime. In order to provide further interpretation, we evaluate in the next section the linear entropy of the
system.

5.2. The linear entropy as measure of coherence


The linear entropy is an important quantity which help measuring the degree of purity or mixing of quantum
states (measure of coherence of a quantum open system) beside the Von Neumann entropy S [53, 54]. It is
defined as:
S (t ) = 1 - tr [r 2 (t )] , (55)
where ρ(t) defines the system’s density matrix. For pure state, we have ρ (t) = ρ(t) and tr [r 2 (t )] = 1, and then S
2

(t) = 0. For mixed state, we have tr [r 2 (t )] < 1 and thus, 0 < S(t) < 1. It is important to mention that, the
increase in linear entropy S(t) due to the system-environment interaction is closely related to the decoherence
phenomenon (which implies loss of quantum coherence in the system), induced by the diffusion process [53].
We can define the measure of coherence as follows [30]:
C (t ) = tr [r 2 (t )]. (56)
If we assume that the initial state is a coherent state, then the initial density matrix is ρ(0) = |α1, α2〉〈α1, α2|, with
aj = gj e ifj . Thus, the initial density matrix elements are:
[r (0)]n1¢ n2¢ m1¢ m2¢ = án1¢ , n2¢∣a1, a2ñáa1, a2∣m1¢, m 2¢ñ. (57)
In addition, by expanding the coherent states in terms of states number, we get:
∣a1∣n1¢ ∣a2∣n2¢ ∣ a12 ∣ ∣ a 22 ∣
áa1, a2∣n1¢ , n2¢ñ = e -( 2 + 2 ) e -i (n1¢ f1+ n2¢ f2). (58)
n1¢!n2¢!
Applying the results of equation (37) giving the elements of the density matrix, equation (56) becomes:
C (t ) =  (a1, a2) r 2 (t ) , (59)
where
¥ ¥
∣a1∣2(n1¢ + m1¢) ∣a2∣2(n2¢ + m2¢)
 (a1, a2) = e -(∣ a1∣ åå
2 + ∣ a ∣2 )
2 ´ , (60)
n1¢ m1¢ n2¢ m2¢ n1¢!m1¢! n2¢!m 2¢!
and r(t) defined by equation (40). Considering these results, the linear entropy S(t) can be derived as:
S (t ) = 1 - C (t ) = 1 -  (a1, a2) r 2 (t ). (61)

10
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 4. Evolution of decoherence factor in a system of a magneto-oscillator interacting with an ohmic reservoir with respect to both
the temperature and time (figure 2(a)) and with respect to both the magnetic field and the cutting frequency (figure 2(b)) respectively.
Blue color corresponds to α = 0 and β = 0, Black color corresponds to α = 0.55 and β = 0.75, Red color corresponds to α = 0.9 and
β = 0.9.

• Case of ohmic reservoir

In the case of ohmic reservoir (s = 1) and in the commutative limit (i.e. θ = η = 0), we obtain:
C (t ) =  (a1, a2) exp [ - 2b [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2]
´ [t Lc tan-1(Lc t ) - ln (1 + Lc2 t 2)]] , (62)
G0 KB T
at zero temperature (T = 0), with b = ,
p Lc
C (t ) =  (a1, a2) exp [ - 2a [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2]
1
´ ⎡ln (1 + Lc2 t 2) + 2 ln [ sinh (pKB Tt )] ⎤ ⎤ , (63)

⎣ pKB Tt ⎦⎥
⎥ ⎦
at low temperature (Λc ? KBT), and
C (t ) =  (a1, a2)(1 + Lc2 t 2)-a [(E n1¢- E m1¢ )
2 + (E
n2¢ - E m2¢ )
2]
, (64)
G0
at high temperature, with a = p
.

• Case of super-ohmic reservoir

In the case of super-ohmic reservoir (s = 3) and in the commutative limit (i.e. θ = η = 0), we obtain:

K T
C (t ) =  (a1, a2) exp ⎡ - 4a (KB T )2 [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2] z ⎛2, B ⎞ ⎤ ⎜ ⎟

⎣ ⎝ Lc ⎠ ⎥⎦
K T (1 + i L c t ) ⎞ ⎤
´ exp ⎡2a (KB T )2 [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2] z ⎛2, B⎜ ⎟

⎣ ⎝ Lc ⎠⎥

K T (1 - i L c t ) ⎞ ⎤
´ exp ⎡2a (KB T )2 [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2] z ⎛2, B⎜ ⎟

⎣ ⎝ Lc ⎠⎥

1 1
´ exp ⎡ - 2aLc2 [(E n1¢ - E m1¢ )2 + (E n2¢ - E m2¢ )2] ⎡ + - 2⎤ ⎤. (65)

⎣ ⎢
⎣ ( 1 + i L c t ) 2 ( 1 - i L c t ) 2
⎦⎥
⎥ ⎦
Figures 5 and 7 depict the evolution of the linear entropy in time in the system of harmonic oscillator under
the magnetic fields effects at low (figures 5(a) and 7(a)) and high (figures 5(b) and 7(b)) temperature, when the
system interacts with an ohmic and super-ohmic reservoir respectively. One can easily observe that the non-
commutativity effects present significant impact on the linear entropy. For the case of the system interacting
with an ohmic reservoir (figure 5(b)), the linear entropy is enhanced with an increase in the non-commutative

11
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 5. Effects of the non-commutative phase-space on the linear entropy of a system of magneto-oscillator interacting with an
ohmic reservoir at low temperature (figure 5(a)), and at high temperature (figure 5(b)), both represented by equation (61).

parameters, while in contrary the inverse phenomenon is observed for the case of super-ohmic reservoir
(figure 7(b)). Therefore, increasing the degree of mixedless in the system state, induce loss of coherent dynamics.
This result is in perfect agreement with the previous as regard to decoherence causing factor.
In addition, it is important to notice that, the coherence time scale in which the system is coupled to an
ohmic reservoir is much larger than that of the super-ohmic reservoir when the non-commutative effects are
taken into consideration. However, in the commutative phase-space case, the coherence is better preserved for a
larger period of time for the super-ohmic reservoir. Moreover, at a finite time, both linear entropy curves
asymptotically increase in the case of non-commutative phase-space as well as in the commutative phase-space
cases characterizing a complete decoherence state of the system.
Furthermore, plotting the linear entropy with respect simultaneously to the time and the temperature
(figures 6(a) and 8(a)) and with respect to the magnetic field and the cut-off frequency (figures 6(b) and 8(b)),
respectively leads to similar conclusion as previously mentioned in Figs 2 and 4. This implies that the
decoherence process occurs very quickly in the commutative phase-space systems than in the non-commutative
phase-space systems. We also realize from both figures that, in a lower magnetic field effects, the linear entropy
goes to zero which characterizes a total coherent state of the system. However, for intense magnetic field effects,
there is an intermediate zone, which characterize an abrupt appearance of decoherence process in the system,
afterward it reaches a linear regime where the system is completely decoherent.

6. Conclusion

This paper’s main goal was to analyze in detail the decoherence dynamics of charged particle interacting with its
environment via an energy-preserving QND type interaction in non-commutative phase-space. For this reason,
the non-commutative model Hamiltonian was first derived from commutative one. Thus, the dynamics of the
system was studied following the master equation approach within the Born-Markov approximation. It was
therefore found that in this master equation, there is no dissipation term, but only a term governing
decoherence, which implies that, such systems undergo decoherence with no energy dissipation. An exact
solution of the density matrix for the dissipationless model of decoherence of the system, interacting with the
bosonic bath was derived. We determined analytical expression of the decoherence-causing term and linear
entropy, which are an indicators of the coherence in the reduced density matrix for the cases of low and high
temperature considering the system in permanent interaction with an ohmic and super-ohmic reservoirs.
Moreover, our analysis shows that, the decay of coherence in the system interacting with its bath via a QND type
interaction depends on the eigen values of the system. In addition, the reservoirs control only asymmetry in the
decoherence rates of various non-diagonal elements in the density matrix.
Numerical analyses revealed that, the decoherence-causing term decreases with magnetic field and increases
with non-commutativity effects for both ohmic and super-ohmic reservoirs at low and high temperature. This
implied that decoherence occurs more quickly in the commutative phase-space system than in the non-
commutative phase-space system. Moreover, when the system interacts with an ohmic reservoir, the coherence
time interval is much larger for non-commutative phase-space system compared to that of the system in
commutative phase-space, while the inverse phenomenon is observed when the system interacts with super-
ohmic reservoir. In addition, we also observed that, in both cases, the linear entropy increases very quickly with

12
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 6. Evolution of linear entropy in a system of a magneto-oscillator interacting with an ohmic reservoir with respect to both the
temperature and time (figure 6(a)) and with respect to both the magnetic field and the cutting frequency (figure 6(b)) respectively. Blue
color corresponds to α = 0 and β = 0, Black color corresponds to α = 0.55 and β = 0.75, Red color corresponds to α = 0.9 and
β = 0.9.

Figure 7. Effects of the non-commutative phase-space on the linear entropy of a system of magneto-oscillator interacting with a
super-ohmic reservoir at low temperature (figure 7(a)), and at high temperature (figure 7(b)), considering equation (61).

the magnetic field and with the cut-off frequency. This prove that increasing the magnetic field in the system
favors the degree of mixed state of the system, thereby causing its decoherence. However, an increase in the non-
commutative parameters considerably improve coherence in the system.
Another interesting feature that comes out as regards to both the decoherence-causing term and the linear
entropy is that by suitably adjusting the non-commutative parameters, we can be able to control decoherence in
the system. This corroborates our previous results [55] in which it was demonstrated that non-commutativity
effects improved the coherence time scale. Our results might be very useful in the analysis of experimental
situations that deal with QND measurements. However, these results might be more interesting if instead of
using discrete variable, continuous-variable systems were used to achieve decoherence dynamics, given its
potential application in quantum computing, thus we intend to really focus on this particular point in our next
research project.

Data availability statement

The data generated and/or analysed during the current study are not publicly available for legal/ethical reasons
but are available from the corresponding author on reasonable request.

13
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Figure 8. Evolution of linear entropy in a system of a magneto-oscillator interacting with a super-ohmic reservoir in terms of both the
temperature and time (figure 8(a)) and in terms of both the magnetic field and the cutting frequency (figure 8(b)) respectively. Blue
color corresponds to α = 0 and β = 0, Black color corresponds to α = 0.55 and β = 0.75, Red color corresponds to α = 0.9 and
β = 0.9.

Appendix A. Exact unitary evolution operator


¥
USB (t ) = exp { å Ck (t )}, (66)
k=1
t 1 t t1
where C1 = -i ò USB (t ¢) dt ¢ and C2 = - 2 ò dt1 ò dt2 [USB (t1), USB (t2)].
0 0 0
It is straightforward to find
t ¥ 2
C1 = - i ò0 USB (t ¢) dt ¢ = å å h j (bn+,jan,j - bn,ja*n,j )
n= 0 j= 1
(67)

gn, j
where an, j (t ) = wn, j
(1 - e iwn, jt ). Further, using the following commutation relation,
¥ 2
[HSB (t1) , HSB (t2)] = - 2i å å h j2∣gn, j ∣2 sin (wn, j (t1 - t2)) , (68)
n= 0 j= 1

then we obtain
t t1 2
1
C2 = -
2 ò0 dt1 ò0 dt2 [USB (t1) , USB (t2)] = - i å h j2 D j ,
j=1
(69)

∣gn2, j∣
with Dj (t ) = å+¥
n=0 (sin (wn, jt ) - wn, jt ), j = 1, 2. Thus, the solution of equation (26) takes the
w n2, j
following form
2 ¥
USB (t ) =  exp [ å h j (bn+, j an, j - bn, j a*n, j ) - ih j2 D j ]. (70)
j=1 n= 0

Appendix B. Exact solution of master equation

[rS (t )]mj¢ n j¢ = trB [U (t ) r (0) U +(t ) Tmj¢ n j¢] , (71)


where Tmj¢ nj¢ = ∣mj¢ nj¢ñ , ∣nj¢ñ is the eigenstates of hj with eigenvalues E nj¢ . Considering the Heisenberg picture, the
operator Tnj¢ mj¢ (t ) is derived as:
-i (E mj¢ - E nj¢ ) t - i (Em2 ¢ - En2¢ ) D j (t ) -Fn¢ m¢ (t )
Tn j¢ mj¢ (t ) = U +(t ) Tn j¢ mj¢ U (t ) = e j j e j j , (72)
taking Fnj¢ mj¢ (t ) = (E nj¢ - E mj¢ ) å n¥= 0(bn*, jan, j (t ) - bn, ja*n, j (t )), then we have:
[rS (t )]mj¢ n j¢ = trB [Tn j¢ mj¢ (t ) r (0)]. (73)

14
Phys. Scr. 96 (2021) 085705 Y D Germain et al

Furthermore, substituting equation (72) into (73), we obtain:


-i (E mj¢ - E nj¢ ) t - i (Em2 ¢ - En2¢ ) D j (t )
[rS (t )]mj¢ n j¢ = [rS (0)]mj¢ n j¢ e j j trB [r (0) e -Fnj¢ mj¢ (t )] , (74)
where
trB [r (0) e -Fnj¢ mj¢ (t )] = áe -Fnj¢ mj¢ (t ) ñ. (75)
Here the average is taken with respect to the thermal bath state at equilibrium. Thus, for an operator D which is a
áD 2ñ
linear combination of creation and annihilation operators we get áe Dñ = e 2 , then considering this identity, we
have
¥
-1
áe -Fn¢j m¢j (t ) ñ.=  exp (E mj¢ - E n j¢ )2∣an, j (t )∣2 (2Nn, j + 1). (76)
n= 0 2
In addition, using the definition of αn,j(t) and the Bose–Einstein distribution, we find that trB [r (0) e-Fnj¢ mj¢ (t )] =
¥ 1 ∣gn, j∣2 w
n = 0 exp {- (E mj¢ - E nj¢ ) 2 (1 - cos (wn, jt )) coth ( 2Kn,Tj )}, and definitively the exact solution of the
2 w n, j
2 B

reduced density matrix of equation (35) is given by


2
-i (E mj¢ - E nj¢ ) t - i (Em2 ¢ - En2¢ ) D j (t ) -Gj (t )(E m ¢ - E n ¢ )2
[rS (t )]n ¢ m ¢ = [rS (0)]n ¢ m ¢ e j j e j j , (77)
j=1

ORCID iDs

Alain Giresse Tene https://orcid.org/0000-0002-6440-2813


Martin Tchoffo https://orcid.org/0000-0003-4989-2751

References
[1] Snyder H S 1947 The electromagnetic field in quantized space-time Phys. Rev. 72 68
[2] Bertolami O, Rosa J G, De Aragao C M L, Castorina P and Zappala D 2005 Noncommutative gravitational quantum well Phys. Rev. D 72
025010
[3] Douglas M R and Nekrasov N A 2001 Noncommutative field theory Rev. Mod. Phys. 73 977
[4] Connes A, Douglas M R and Schwarz A 1998 Noncommutative geometry and matrix theory J. High Energy Phys. 1998 003
[5] Seiberg N and Witten E 1999 String theory and noncommutative geometry J. High Energy Phys. 1999 032
[6] Ho P-M and Kao H-C 2002 Noncommutative quantum mechanics from noncommutative quantum field theory Phys. Rev. Lett. 88
151602
[7] Banerjee S, Ghosh J and Ghosh R 2007 Phase-diffusion pattern in quantum-nondemolition systems Phys. Rev. A 75 062106
[8] Nielsen M A and Chuang I L 2001 Quantum computation and quantum information (Cambridge: Cambridge University Press) http://
csis.pace.edu/~ctappert/cs837-19spring/QC-textbook.pdf
[9] Barnett S M et al 2017 Journeys from quantum optics to quantum technology Prog. Quantum Electron. 54 19–45
[10] Barnett S 2009 Quantum Information vol 16 (Oxford: Oxford University Press)
[11] Pachos J K and Knight P L 2003 Quantum computation with a one-dimensional optical lattice Phys. Rev. Lett. 91 107902
[12] Feynman R P and Vernon F L 2000 The theory of a general quantum system interacting with a linear dissipative system Ann. Phys. 281
547–607
[13] Caldeira A O and Leggett A J 1983 Path integral approach to quantum brownian motion Physica A 121 587–616
[14] Grabert H and Talkner P 1983 Quantum brownian motion Phys. Rev. Lett. 50 1335
[15] Banerjee S and Ghosh R 2003 General quantum brownian motion with initially correlated and nonlinearly coupled environment Phys.
Rev. E 67 056120
[16] Caldeira A O and Leggett A J 1983 Quantum tunnelling in a dissipative system Ann. Phys. 149 374–456
[17] Paavola J, Piilo J, Suominen K-A and Maniscalco S 2009 Environment-dependent dissipation in quantum brownian motion Phys. Rev.
A 79 052120
[18] Ambegaokar V 2006 Dissipation and decoherence in a quantum oscillator J. Stat. Phys. 125 1183–92
[19] Schlosshauer M, Hines A P and Milburn G J 2008 Decoherence and dissipation of a quantum harmonic oscillator coupled to two-level
systems Phys. Rev. A 77 022111
[20] Zurek W H 2003 Decoherence, einselection, and the quantum origins of the classical Rev. Mod. Phys. 75 715
[21] Schlosshauer M and Fine A 2007 Decoherence and the foundations of quantum mechanics Quantum Mechanics at the Crossroads. The
Frontiers Collection (Berlin: Springer) 125–48
[22] Zurek W H and Paz J P 1995 Decoherence, chaos, the quantum and the classical Il Nuovo Cimento B (1971–1996) 110 611–24
[23] Tameshtit A and Sipe J E 1993 Rates of quantum decoherence in regular and chaotic systems Phys. Rev. A 47 1697
[24] Thorne K S, Drever R W P, Caves C M, Zimmermann M and Sandberg V D 1978 Quantum nondemolition measurements of harmonic
oscillators Phys. Rev. Lett. 40 667
[25] Unruh W G 1979 Quantum nondemolition and gravity-wave detection Phys. Rev. D 19 2888
[26] Mozyrsky D and Privman V 1998 Adiabatic decoherence J. Stat. Phys. 91 787–99
[27] Gangopadhyay G and Lin S H 1997 The effect of environment induced pure decoherence on the generalized jaynes-cummings model
Pramana 49 399–416
[28] Shao J, Ge M-L and Cheng H 1996 Decoherence of quantum-nondemolition systems Phys. Rev. E 53 1243

15
Phys. Scr. 96 (2021) 085705 Y D Germain et al

[29] Caves C M, Thorne K S, Drever R W P, Sandberg V D and Zimmermann M 1980 On the measurement of a weak classical force coupled
to a quantum-mechanical oscillator: I. issues of principle Rev. Mod. Phys. 52 341
[30] Karmakar A and Gangopadhyay G 2012 Decoherence without dissipation due to fermionic bath Phys. Scr. 85 045008
[31] Banerjee S and Ghosh R 2007 Dynamics of decoherence without dissipation in a squeezed thermal bath J. Phys. A: Math. Theor. 40
13735
[32] Mirza B and Mohadesi M 2004 The klein-gordon and the dirac oscillators in a noncommutative space Commun. Theor. Phys. 42 664
[33] Hassanabadi H, Molaee Z and Zarrinkamar S 2014 Noncommutative phase space schrödinger equation with minimal length Adv. High
Energy Phys. 2014 459345
[34] Haouam I 2014 The phase-space noncommutativity effect on the large and small wavefunction components approach at dirac equation
Open Access Library Journal 5 1–10
[35] Li K, Wang J and Chen C 2005 Representation of noncommutative phase space Mod. Phys. Lett. A 20 2165–74
[36] Rajesh A, Bandyopadhyay M and Jayannavar A M 2017 Decoherence control mechanisms of a charged magneto-oscillator in contact
with different environments Eur. Phys. J. B 90 1–12
[37] Bandyopadhyay M 2009 Quantum thermodynamics of a charged magneto-oscillator coupled to a heat bath J. Stat. Mech: Theory Exp.
2009 P05002
[38] Jellal A, Schreiber M and El Kinani E H 2005 Two coupled harmonic oscillators on noncommutative plane Int. J. Mod. Phys. A 20
1515–29
[39] Martin T, Armel A K, Germain Y D, Giresse T A and Lukong Cornelius F A I 2019 Decoherence of quantum brownian particle trapped
in a penning potential in a non-commutative space Journal of Physics Communications 3 095006
[40] Palma G M, Suominen K-A and Ekert A 1996 Quantum computers and dissipation Proceedings of the Royal Society of London. Series A:
Mathematical, Physical and Engineering Sciences 452 567–84
[41] Yi X X, Tong D M, Wang L C, Kwek L C and Oh C H 2006 Geometric phase in open systems: Beyond the markov approximation and
weak-coupling limit Phys. Rev. A 73 052103
[42] Vorrath T, Brandes T and Kramer B 2004 Dynamics of a large spin with weak dissipation Chem. Phys. 296 295–300
[43] Banerjee S and Srikanth R 2008 Geometric phase of a qubit interacting with a squeezed-thermal bath Eur. Phys. J. D 46 335–44
[44] Blanes S, Casas F, Oteo J-A and Ros J 2009 The magnus expansion and some of its applications Phys. Rep. 470 151–238
[45] Chaudhry A Z and Gong J 2013 Amplification and suppression of system-bath-correlation effects in an open many-body system Phys.
Rev. A 87 012129
[46] Wang X and Wang J 2019 Nonequilibrium effects on quantum correlations: Discord, mutual information, and entanglement of a two-
fermionic system in bosonic and fermionic environments Phys. Rev. A 100 052331
[47] Vorrath T 2003 Dissipation-induced collective effects in two-level systems PhD Thesis Staats-und Universitätsbibliothek Hamburg Carl
von Ossietzky https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.463.7773&rep=rep1&type=pdf
[48] Hinchliffe I, Kersting N and Ma Y L 2004 Review of the phenomenology of noncommutative geometry Int. J. Mod. Phys. A 19 179–204
[49] Banerjee R, Roy B D and Samanta S 2006 Remarks on the noncommutative gravitational quantum well Phys. Rev. D 74 045015
[50] Biercuk M J, Uys H, VanDevender A P, Shiga N, Itano W M and Bollinger J J 2009 Optimized dynamical decoupling in a model
quantum memory Nature 458 996–1000
[51] Paavola J and Maniscalco S 2010 Decoherence control in different environments Phys. Rev. A 82 012114
[52] Bar-Gill N, Bhaktavatsala Rao D D and Kurizki G 2011 Creating nonclassical states of bose-einstein condensates by dephasing collisions
Phys. Rev. Lett. 107 010404
[53] Isar A, Sandulescu A and Scheid W 1999 Purity and decoherence in the theory of a damped harmonic oscillator Phys. Rev. E 60 6371
[54] Anastopoulos C and Halliwell J J 1995 Generalized uncertainty relations and long-time limits for quantum brownian motion models
Phys. Rev. D 51 6870
[55] Tchoffo M, Deuto G Y, Nsangou I, Koumetio A A, Tchapda L S Y and Tene A G 2020 Decoherence of a damped anisotropic harmonic
oscillator under magnetic field effects in a two-dimensional noncommutative phase-space Journal of Applied Mathematics and Physics 8
2801–23

16

You might also like