Brought to you by:
Paper

Decoherence dynamics of a charged particle within a non-demolition type interaction in non-commutative phase-space

, , , and

Published 27 May 2021 © 2021 IOP Publishing Ltd
, , Citation Yiande Deuto Germain et al 2021 Phys. Scr. 96 085705 DOI 10.1088/1402-4896/ac0273

1402-4896/96/8/085705

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.

Export citation and abstract BibTeX RIS

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 [13]. 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 [1619]. 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) [2429]. 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:

Equation (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:

Equation (2)

where $\check{{\hslash }}$ is the effective Plank constant given by $\check{{\hslash }}={\hslash }\left(1+\displaystyle \frac{\theta \eta }{4{{\hslash }}^{2}}\right)$, ${\hat{x}}_{i(j)}$ and ${\hat{p}}_{i(j)}$ the coordinate and 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 phase-space (${\hat{x}}_{i}$ , ${\hat{p}}_{i}$) and the commutative ones (xi , pi ), is obtained using the Bopp-Shift linear transformation [3235]:

Equation (3)

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]:

Equation (4)

with

Equation (5)

characterized by its confinement frequency ω0 and the polar angle between the position vector and x1, ϕ. Here, $({\hat{x}}_{1},{\hat{x}}_{2})$ and $({\hat{p}}_{1},{\hat{p}}_{2})$ are respectively the non-commutative coordinates and momentum. Considering the symetric gauge $\hat{A}=\tfrac{B}{2}(-{\hat{x}}_{2},{\hat{x}}_{1},0)$ and the Bopp-Shift transformation given by equation (3), the Hamiltonian can be rewritten as:

Equation (6)

where $M=\displaystyle \frac{m}{\left[{\left(1+\tfrac{{\omega }_{c}}{2{\omega }_{\theta }}\right)}^{2}+{\left(\tfrac{{\omega }_{0}}{{\omega }_{\theta }}\right)}^{2}\left(1+\tfrac{\cos (4\phi )}{5}\right)\right]}$, $\lambda =\left[\left(1+\tfrac{{\omega }_{c}}{2{\omega }_{\theta }}\right)\left({\omega }_{\eta }+\tfrac{{\omega }_{c}}{2}\right)+\tfrac{{\omega }_{0}^{2}}{{\omega }_{\theta }}\left(1+\tfrac{\cos (4\phi )}{5}\right)\right]$,

${\rm{\Omega }}=\sqrt{\left[{\left(1+\tfrac{{\omega }_{c}}{2{\omega }_{\theta }}\right)}^{2}+{\left(\tfrac{{\omega }_{0}}{{\omega }_{\theta }}\right)}^{2}(1+\tfrac{\cos (4\phi )}{5})][{\omega }_{0}^{2}(1+\tfrac{{\cos }(4\phi )}{5})+{\left({\omega }_{\eta }+\tfrac{{\omega }_{c}}{2}\right)}^{2}\right]}$, ${\omega }_{\theta }=\tfrac{2{\hslash }}{m\theta }$, ${\omega }_{\eta }=\tfrac{\eta }{2m{\hslash }}$, ${\rm{\Omega }}=\sqrt{\tfrac{K}{M}}$, ${\omega }_{c}=\tfrac{{eB}}{{mc}}$ 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 ${E}_{{n}_{1}{n}_{2}}\,={\hslash }\omega ^{\prime} ({n}_{1}+{n}_{2}+1)$. 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]

Equation (7)

where ai and ${a}_{j}^{+}$ satisfy the usual commutation relations $[{a}_{j}^{+},{a}_{i}]=-{\delta }_{{ij}}$, i, j = 1, 2. Considering equation (7), the Hamiltonian becomes

Equation (8)

with the frequencies

Equation (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:

Equation (10)

where n1 and n2 are the quantum numbers (n1, n2 = 0, 1, 2, ⋯), 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 [4042]. The total Hamiltonian of such system interacting with a bath of harmonic oscillators, via a QND type interaction [30, 31, 43] is defined by:

Equation (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:

Equation (12)

where ${\omega }_{n,j}=\sqrt{\left(1+{\left(\tfrac{{m}_{n}{\omega }_{n}\theta }{2{\hslash }}\right)}^{2})({\omega }_{n}^{2}+{\left(\tfrac{\eta }{2{m}_{n}{\hslash }}\right)}^{2}\right)}\pm (\tfrac{\eta }{2{m}_{n}{\hslash }}+\tfrac{{m}_{n}{\omega }_{n}^{2}\theta }{2{\hslash }})$ are the non-commutative frequencies of nth oscillators in the x1 and x2 directions respectively. bn, j and ${b}_{n,j}^{+}$ (j = 1, 2) are the annihilation and creation bosonic operators of the heat bath in the x1 and x2 directions respectively, and satisfy the commutation relations $[{b}_{n,j},{b}_{n,j}^{+}]={\delta }_{{nn}^{\prime} }{\delta }_{{jj}^{\prime} }$. 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:

Equation (13)

with ${h}_{j}={\hslash }{{\rm{\Omega }}}_{j}\left({a}_{j}^{+}{a}_{j}+\tfrac{1}{2}\right)$.

3. Bath correlation function evaluation

Let $K(t-t^{\prime} )$ 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:

Equation (14)

The correlation function only depends on the quantity $\tau =t-t^{\prime} $, for any time-independent form of the interaction. The state of the bath is considered as a canonical Gibbs state given by:

Equation (15)

with $Z={{tr}}_{B}[{e}^{-\beta {H}_{B}}]$. From the interaction Hamiltonian (13), we can note that

Equation (16)

To evaluate $\check{B}(\tau )$, let us assume that

Equation (17)

Therefore,

Equation (18)

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 $\langle {b}_{n,j}^{+}{b}_{n,j}\rangle $ and $\langle {b}_{n,j}{b}_{n,j}^{+}\rangle $ 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:

Equation (19)

where we have assumed that $\langle {b}_{n,j}^{+}{b}_{n,j}\rangle ={N}_{n,j}$ and $\langle {b}_{n,j}{b}_{n,j}^{+}\rangle ={N}_{n,j}+1$. Since the bath is in thermal equilibrium, Nn,j is thus, defined by the Bose–Einstein distribution:

Equation (20)

Then the correlation function can be simplified to:

Equation (21)

where T denotes the temperature of the bath and KB the Boltzmann constant.

Setting ${\nu }_{j}={\sum }_{n=0}^{\infty }| {g}_{n,j}{| }^{2}\cos ({\omega }_{n,j}\tau )\coth (\tfrac{{\omega }_{n,j}}{2{K}_{B}T})$ and ${\chi }_{j}={\sum }_{n=0}^{\infty }| {g}_{n,j}{| }^{2}\sin ({\omega }_{n,j}\tau )$, therefore the correlation function becomes:

Equation (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

Equation (23)

where H0 = HS + HB is the free Hamiltonian in the absence of interaction. Then, considering equation (23) we have

Equation (24)

By defining ${U}_{0}(t)=\exp (-{{iH}}_{0}t)$, one has:

Equation (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:

Equation (26)

where its solution can be set as [44, 45]:

Equation (27)

Thus, the solution of equation (26) takes the following form (details can be found in appendix A):

Equation (28)

Further, the exact unitary time evolution operator according to equation (25) is given by:

Equation (29)

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 system-environment coupling Hamiltonian as follows:

Equation (30)

where the quantities hj contain operators described in the Hilbert space of the system defined by ${h}_{j}={\hslash }{{\rm{\Omega }}}_{j}\left({a}_{j}^{+}{a}_{j}+\tfrac{1}{2}\right)$ and Bj (t) the operators in the environmental Hilbert space given by equation (16). Then, 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]:

Equation (31)

where ${\tilde{h}}_{j}(t,t^{\prime} )={U}_{S}(t,t^{\prime} ){h}_{j}{U}_{S}^{+}(t,t^{\prime} )$, with ${U}_{S}(t,t^{\prime} )$ a unitary operator describing the evolution of the system and ${K}_{j}(t-t^{\prime} )$ 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]:

Equation (32)

where

Equation (33)

and

Equation (34)

The state of the system alone at anytime t is found by tracing over the reservoir degree of freedom as

Equation (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:

Equation (36)

where ${T}_{{m}_{j}^{{\prime} }{n}_{j}^{{\prime} }}=| {m}_{j}^{{\prime} }{n}_{j}^{{\prime} }\rangle $ , $| {n}_{j}^{{\prime} }\rangle $ is the eigenstates of hj with eigenvalues ${E}_{{n}_{j}^{{\prime} }}$. Therefore, we find that the exact solution of the reduced density matrix of equation (32) is given by (see appendix B for detailed calculations):

Equation (37)

where

Equation (38)

Equation (39)

with ${\prod }_{j=1}^{2}{e}^{-i({E}_{{m}_{j}^{{\prime} }}^{2}-{E}_{{n}_{j}^{{\prime} }}^{2}){{\rm{\Delta }}}_{j}(t)}$ 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 ($\dot{{\rm{\Gamma }}}(t)$), consequently, Γ(t) is responsible for coherence in the system.

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 × 10−38 m2 and η = β × 10−60 kg2 m2 s−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 ${E}_{{n}_{j}^{{\prime} }}$ of the systems determine the rate while the function Γj (t) is controlled by the bath coupling type. Thus, the term

Equation (40)

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 ${J}_{j}(\omega )={\sum }_{n=0}^{\infty }{g}_{n,j}^{2}\delta (\omega -{\omega }_{n,j})$. 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]:

Equation (41)

where ${\omega }_{\mathrm{1,2}}=\sqrt{(1+{\alpha }_{\theta }^{2}{\omega }^{2})({\omega }^{2}+{\beta }_{\eta }^{2})}\pm ({\beta }_{\eta }+{\alpha }_{\theta }{\omega }^{2})$, (with ${\alpha }_{\theta }=\tfrac{m^{\prime} \theta }{2{\hslash }}$, ${\beta }_{\eta }=\tfrac{\eta }{2m^{\prime} {\hslash }}$), G0 the coupling constant and Λc the cut-off frequency of the bath. Therefore, we have

Equation (42)

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:

Equation (43)

and

Equation (44)

In the commutative limit (i.e. θ = η = 0), equations (43) and (44) reduce to:

Equation (45)

and

Equation (46)

respectively, for zero temperature, which confirm the results obtained in [31, 52]. For low temperature (i.e. Λc KB T), equation (44) reduces to [26, 40]:

Equation (47)

The first term arises from the quantum vacuum fluctuations while the second is due to thermal ones. The above expression reduces to ${\rm{\Gamma }}(t)=\displaystyle \frac{{G}_{0}{\left({{\rm{\Lambda }}}_{c}t\right)}^{2}}{\pi }$ for $t\leqslant {{\rm{\Lambda }}}_{c}^{-1}$ (quiet regime), where the fluctuations are ineffective in the decoherece process. While it reduces to ${\rm{\Gamma }}(t)=\displaystyle \frac{{G}_{0}\mathrm{ln}({{\rm{\Lambda }}}_{c}t)}{\pi }$ for ${{\rm{\Lambda }}}_{c}^{-1}\leqslant t\leqslant {\left({K}_{B}T\right)}^{-1}$. In this case, the main causes of coherence loss (decoherence) are the quantum fluctuations. However, one has ${\rm{\Gamma }}(t)=\displaystyle \frac{{G}_{0}{K}_{B}{Tt}}{\pi }$ for ${\left({K}_{B}T\right)}^{-1}\leqslant t$ (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:

Equation (48)

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

Equation (49)

with ${{\rm{\Gamma }}}_{0}=\tfrac{2{G}_{0}{K}_{B}T}{\pi }$, 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:
    Equation (50)
  • For low temperature, one has:
    Equation (51)
    In the commutative limit (i.e. θ = η = 0), equation (51) becomes [40]
    Equation (52)
    where ζ is the generalized Riemann Zeta function.
  • For high temperature, one has:
    Equation (53)
    which reduces to
    Equation (54)
    in the commutative limit (i.e. θ = η = 0).

In the above expressions, we have as previously mentioned ${\omega }_{\mathrm{1,2}}=\sqrt{(1+{\alpha }_{\theta }^{2}{\omega }^{2})({\omega }^{2}+{\beta }_{\eta }^{2})}\pm ({\beta }_{\eta }+{\alpha }_{\theta }{\omega }^{2})$, (with ${\alpha }_{\theta }=\tfrac{m^{\prime} \theta }{2{\hslash }}$, ${\beta }_{\eta }=\tfrac{\eta }{2m^{\prime} {\hslash }}$), G0 the coupling constant and ${{\rm{\Gamma }}}_{0}=\tfrac{2{G}_{0}{K}_{B}T}{\pi }$. 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 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).

Figure 1.

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)).

Standard image High-resolution image
Figure 2.

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.

Standard image High-resolution image
Figure 3.

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)).

Standard image High-resolution image
Figure 4.

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.

Standard image High-resolution image

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.

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:

Equation (55)

where ρ(t) defines the system's density matrix. For pure state, we have ρ2(t) = ρ(t) and ${tr}[{\rho }^{2}(t)]=1$, and then S(t) = 0. For mixed state, we have ${tr}[{\rho }^{2}(t)]\lt 1$ 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]:

Equation (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 ${\alpha }_{j}={\gamma }_{j}{e}^{i{\phi }_{j}}$. Thus, the initial density matrix elements are:

Equation (57)

In addition, by expanding the coherent states in terms of states number, we get:

Equation (58)

Applying the results of equation (37) giving the elements of the density matrix, equation (56) becomes:

Equation (59)

where

Equation (60)

and r(t) defined by equation (40). Considering these results, the linear entropy S(t) can be derived as:

Equation (61)

  • Case of ohmic reservoir

In the case of ohmic reservoir (s = 1) and in the commutative limit (i.e. θ = η = 0), we obtain:

Equation (62)

at zero temperature (T = 0), with $b=\displaystyle \frac{{G}_{0}{K}_{B}T}{\pi {{\rm{\Lambda }}}_{c}}$,

Equation (63)

at low temperature (Λc KB T), and

Equation (64)

at high temperature, with $a=\tfrac{{G}_{0}}{\pi }$.

  • 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:

Equation (65)

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 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.

Figure 5.

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).

Standard image High-resolution image

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.

Figure 6.

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.

Standard image High-resolution image
Figure 7.

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).

Standard image High-resolution image
Figure 8.

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.

Standard image High-resolution image

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 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.

Appendix A.: Exact unitary evolution operator

Equation (66)

where ${C}_{1}=-i{\int }_{0}^{t}{U}_{{SB}}(t^{\prime} ){dt}^{\prime} $ and ${C}_{2}=-\tfrac{1}{2}{\int }_{0}^{t}{{dt}}_{1}{\int }_{0}^{{t}_{1}}{{dt}}_{2}[{U}_{{SB}}({t}_{1}),{U}_{{SB}}({t}_{2})]$.

It is straightforward to find

Equation (67)

where ${\alpha }_{n,j}(t)=\tfrac{{g}_{n,j}}{{\omega }_{n,j}}(1-{e}^{i{\omega }_{n,j}t})$. Further, using the following commutation relation,

Equation (68)

then we obtain

Equation (69)

with ${{\rm{\Delta }}}_{j}(t)={\sum }_{n=0}^{+\infty }\displaystyle \frac{| {g}_{n,j}^{2}| }{{\omega }_{n,j}^{2}}(\sin ({\omega }_{n,j}t)-{\omega }_{n,j}t),j\,=\,1,2$. Thus, the solution of equation (26) takes the following form

Equation (70)

Appendix B.: Exact solution of master equation

Equation (71)

where ${T}_{{m}_{j}^{{\prime} }{n}_{j}^{{\prime} }}=| {m}_{j}^{{\prime} }{n}_{j}^{{\prime} }\rangle $ , $| {n}_{j}^{{\prime} }\rangle $ is the eigenstates of hj with eigenvalues ${E}_{{n}_{j}^{{\prime} }}$. Considering the Heisenberg picture, the operator ${T}_{{n}_{j}^{{\prime} }{m}_{j}^{{\prime} }}(t)$ is derived as:

Equation (72)

taking ${F}_{{n}_{j}^{{\prime} }{m}_{j}^{{\prime} }}(t)=({E}_{{n}_{j}^{{\prime} }}-{E}_{{m}_{j}^{{\prime} }}){\sum }_{n=0}^{\infty }({b}_{n,j}^{* }{\alpha }_{n,j}(t)-{b}_{n,j}{\alpha }_{n,j}^{* }(t))$, then we have:

Equation (73)

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

Equation (74)

where

Equation (75)

Here the average is taken with respect to the thermal bath state at equilibrium. Thus, for an operator D which is a linear combination of creation and annihilation operators we get $\langle {e}^{D}\rangle ={e}^{\tfrac{\langle {D}^{2}\rangle }{2}}$, then considering this identity, we have

Equation (76)

In addition, using the definition of αn,j (t) and the Bose–Einstein distribution, we find that ${{tr}}_{B}[\rho (0){e}^{-{F}_{{n}_{j}^{{\prime} }{m}_{j}^{{\prime} }}(t)}]\,={\prod }_{n=0}^{\infty }\exp \{-\displaystyle \frac{1}{2}{\left({E}_{{m}_{j}^{{\prime} }}-{E}_{{n}_{j}^{{\prime} }}\right)}^{2}\displaystyle \frac{| {g}_{n,j}{| }^{2}}{{\omega }_{n,j}^{2}}(1-\cos ({\omega }_{n,j}t))\coth (\tfrac{{\omega }_{n,j}}{2{K}_{B}T})\}$, and definitively the exact solution of the reduced density matrix of equation (35) is given by

Equation (77)

Please wait… references are loading.
10.1088/1402-4896/ac0273