Brought to you by:
Paper

On the structure of the master equation for a two-level system coupled to a thermal bath

Published 18 March 2015 © 2015 IOP Publishing Ltd
, , Citation Inés de Vega 2015 J. Phys. A: Math. Theor. 48 145202 DOI 10.1088/1751-8113/48/14/145202

1751-8121/48/14/145202

Abstract

We derive a master equation from the exact stochastic Liouville–von-Neumann (SLN) equation (Stockburger and Grabert 2002 Phys. Rev. Lett. 88 170407). The latter depends on two correlated noises and describes exactly the dynamics of an oscillator (which can be either harmonic or present an anharmonicity) coupled to an environment at thermal equilibrium. The newly derived master equation is obtained by performing analytically the average over different noise trajectories. It is found to have a complex hierarchical structure that might be helpful to explain the convergence problems occurring when performing numerically the stochastic average of trajectories given by the SLN equation (Koch et al 2008 Phys. Rev. Lett. 100 230402, Koch 2010 PhD thesis Fakultät Mathematik und Naturwissenschaften der Technischen Universitat Dresden).

Export citation and abstract BibTeX RIS

Introduction

In recent decades, there has been an increasing interest in analyzing the dynamics of open quantum systems (OQS) coupled to an environment. Much effort has been devoted to the derivation of master equations that describe the dynamics of the reduced density matrix of the system [57].

An exact master equation can be derived for the case in which the OQS is a harmonic oscillator (or a collection of them) in a thermal bath [8, 9]. However, when the OQS contains an anharmonicity, giving rise for instance to dynamics restricted to a two state basis, no master equation has been derived up to now without making some approximation. Whereas different master equations have been derived within the so-called weak coupling approximation, i.e. assuming that the coupling between the system and the environment is small compared to other energy scales of the problem, the derivation of a master equation beyond such situation has been more elusive [10].

In the projection operator techniques, a projection superoperator $\mathcal{P}$ is defined such that $\mathcal{P}\rho $ captures the relevant part of the total density matrix. Then an approximated master equation is obtained for $\mathcal{P}\rho $ based on a perturbative expansion with respect to the coupling between the (relevant part of the) system and the environment. The standard choice is to consider a projection superoperator such that $\mathcal{P}\rho ={{\rho }_{S}}(t)\;\otimes \;{{\rho }_{B}}$, where ${{\rho }_{S}}(t)={\rm T}{{{\rm r}}_{B}}\rho (t)$ [11], and then follow the Nakajima–Zwanzig method [12, 13], that leads to an equation for $\mathcal{P}\rho $ which contains a time integration over the history of the system. Alternatively, one may follow the so-called time-convolutionless projection operator technique (TCL), which gives rise to an equation that is local in time [5, 14]. Another possibility is to consider that the relevant part of the dynamics can be described with a correlated system-environment state, rather than a tensor-product state ${{\rho }_{S}}(t)\;\otimes \;{{\rho }_{B}}$. In this regard, there is a class of projection superoperators that project the system into correlated system-environment states [14], so that then TCL technique can be applied to calculate the evolution of $\mathcal{P}\rho $, where $\mathcal{P}$ is this class of projection operators. This approach was considered in [15] and [16] by using two different types of correlated projector operators to derive two different types of master equations. Projection operator techniques they rely strongly on the specific choice of a suitable set of projection operators, an election that might depend on the particular situation and be difficult to make for some cases.

In this paper we propose a method to derive a master equation without the use of any approximation, and considering an arbitrary system Hamiltonian HS, which can in principle contain anharmonicities giving rise to a nonlinear spectrum. In particular, we analyse the so-called spin-boson model, in which the OQS is a two-level system. We base our derivation on the so-called stochastic Liouville–von-Neumann (SLN) equation [1, 2], which describes exactly the dynamics of an oscillator coupled to an environment in thermal equilibrium, according to the Hamiltonian

Equation (1)

where $\hat{x}$ is the system Hermitian coupling operator, ${{a}_{\lambda }}(a_{\lambda }^{\dagger })$ are the annihilation (creation) operators of the oscillator λ of the environment, and ${{\hat{g}}_{\lambda }}$ and ${{\omega }_{\lambda }}$ are respectively its coupling and frequency. The stochastic sampling in the SLN method often has convergence problems that have to be tacked in rather sophisticated ways [3, 4]. Indeed, the convergence of the SLN methods fails particularly at long times and for super-ohmic reservoirs. It would be interesting to have a master equation corresponding to the SLN equation, and hence avoid or at least have some insight into such convergence problems that appear in some cases.

This paper contains two main results: first, the derivation of a master equation without any approximation or assumption. In other words, a derivation that does not rely on a projection method, a Born approximation, Markov approximation, a high-temperature assumption, a perturbative approximation, or a rotating wave approximation. The derived master equation has a complex structure in terms of an infinite series of integral terms. Second, using an alternative derivation, we show that the reduced density matrix can also be obtained with a hierarchy of time-local coupled equations. In the case of an exponential correlation function this hierarchy can be simplified and acquire a form which is very similar (but not identical) to the hierarchy obtained by Tanimura [17, 18].

The SLN equation

In his seminal paper, Stockburger [1] (see also [2], [4] and references therein) derived a differential equation for the stochastic reduced density operator of an open quantum system $P_{\xi ,\nu }^{t}$, dependent on the noises $\xi (t)$ and $\nu (t)$, the so-called SLN equation

Equation (2)

Here, we have neglected a re-normalization factor which is not relevant for the present derivation. This equation, valid for environments in thermal equilibrium, allows us to compute different stochastic trajectories for ${{P}_{\xi ,\nu }}$, such that the reduced density operator can be obtained as a sum over them

Equation (3)

The two Gaussian noises in equation (2) must fulfil the following statistical properties

Equation (4)

where ${{\mathcal{M}}_{\xi ,\nu }}[\cdots ]$ is a Gaussian average over the two different noises, and $\theta (t)$ is the Heaviside step function. In the above definitions, ${{\alpha }_{R}}(t)$ and ${{\alpha }_{I}}(t)$ correspond respectively to the real and imaginary parts of the environment correlation function corresponding to the Hamiltonian (1) [5]

Equation (5)

where ${{\hat{g}}_{\lambda }}$ are the coupling constants appearing in (1), and $\beta =1/{{K}_{{\rm B}}}T$, whith kB the Boltzmann constant and T the environment temperature. The above equation (2) can be re-written as two stochastic equations for different stochastic wave vectors $|\psi _{t}^{1}\rangle $ and $|\psi _{t}^{2}\rangle $

Equation (6)

such that $P_{\xi ,\nu }^{t}=|\psi _{t}^{1}\rangle \langle \psi _{t}^{2}|$ [4], and therefore the reduced density operator can be obtained according to (3).

Re-expressing the noises in (2)

Let us define two time dependent quantities

Equation (7)

with

Equation (8)

and $z_{j}^{*}(t)={{({{z}_{j}}(t))}^{*}}$. The latter quantities come in terms of ${{z}_{j\lambda }}$, for j = 0,1, which are here considered as complex variables obeying the properties $\mathcal{M}[{{z}_{j\lambda }}]=0$, $\mathcal{M}[{{z}_{j\lambda }},{{z}_{l\lambda ^{\prime} }}]=0$, $\mathcal{M}[{{z}_{j\lambda }}z_{l\lambda ^{\prime} }^{*}]={{\delta }_{jl}}{{\delta }_{\lambda ,\lambda ^{\prime} }}$, with ${{\mathcal{M}}_{\xi ,\nu }}[\cdots ]=\mathcal{M}[\cdots ]=\int {{{\rm d}}^{2}}{{z}_{0}}{{{\rm e}}^{-|{{z}_{0}}{{|}^{2}}}}\int {{{\rm d}}^{2}}{{z}_{1}}{{{\rm e}}^{-|{{z}_{1}}{{|}^{2}}}}\cdots $. In this definition, the ${{z}_{j}}=\{{{z}_{j0}},{{z}_{j1}},\cdots {{z}_{j\lambda }},\cdots \}$ denotes the set of complex numbers corresponding to each of the λ harmonic oscillators [1921]. The average can also be written more compactly as $\mathcal{M}[\cdots ]=\int {\rm d}\mu ({{z}_{0}})\int {\rm d}\mu ({{z}_{1}})\cdots $ with

Equation (9)

Importantly, the above defined quantities obey the properties (4)

Equation (10)

In the following, one could randomly select a set of complex numbers zj in order to build the two quantities $\hat{\xi }(t)$ and $\hat{\nu }(t)$ according to (7), and repeat the procedure several times to generate several values for them. It is at this point that, because the zj are randomly sampled according to a Gaussian distribution, they can be considered as a set of complex Gaussian noises. In a similar way, the quantities $\hat{\xi }(t)$ and $\hat{\nu }(t)$ built with them become Gaussian coloured noises with properties (4) as $\xi (t)$ and $\nu (t)$. Hence, following a Monte Carlo method, one could re-capture the reduced density matrix as a sum running over M different trajectories of $P_{\xi ,\nu }^{t}$

Equation (11)

Here, we have re-written $P_{\xi ,\nu }^{t}$ as $P_{{{z}_{0}},{{z}_{1}}}^{t}$ to emphasize that the noises $\xi (t)$ and $\nu (t)$ for each trajectory are built according to (8), using a given value for the Gaussian noise sets z0 and z1. Also, the result is approximated because the number of stochastic trajectories M is finite.

In this manuscript, we propose following a different path. Using the fact that now the averages ${{\mathcal{M}}_{\xi ,\nu }}[\cdots ]$ are explicitly written as Gaussian integrals, we will carry on analytically (instead of numerically with a Monte Carlo sampling), the average ${{\mathcal{M}}_{\xi ,\nu }}[\cdots ]$. This permits us to obtain the evolution equation of the density matrix by considering analytically the average of (2).

We note that the particular values for the functions ${{g}_{\lambda }}$ and ${{f}_{\lambda }}$ is unimportant for the present derivation, for which the only relevant quantity is the correlation function (5). Note also that we can define ${{g}_{\lambda }}=(1/\sqrt{2}){{\hat{g}}_{\lambda }}(\sqrt{1+2N({{\omega }_{\lambda }})})$. In that way, the real part of the correlation function can be written in terms of $g_{\lambda }^{2}=\frac{1}{2}\hat{g}_{\lambda }^{2}{\rm coth} \left( {{\omega }_{\lambda }}\beta /2 \right)=\frac{1}{2}\hat{g}_{\lambda }^{2}(2N({{\omega }_{\lambda }})+1)$, with $N(\omega )={{[{\rm exp} (\omega \beta )-1]}^{-1}}$ the average thermal number of quanta in the mode ω.

Deriving a master equation

A time-convoluted master equation

In order to derive a master equation, we need to compute ${\rm d}\mathcal{M}[P]/{\rm d}t$, where for simplicity in the notation we have renamed $P={{P}_{\xi ,\nu }}$. With this aim, we take into account that ${\rm d}\mathcal{M}[P]/{\rm d}t=\mathcal{M}[{\rm d}P/{\rm d}t]$ and consider the SLN equation (2) for P. In order to have a more compact notation, we also re-express the equation (2) in the interaction picture with respect to the system and in a Liouvillian form

Equation (12)

where $\mathcal{L}(t)$ is a super-operator acting over the stochastic projector Pt flattened as a vector, denoted as Pvt. In detail, $\mathcal{L}(t)P_{t}^{v}={{\left( \frac{{\rm i}}{\hbar }\xi (t)[X(t),{{P}_{\xi ,\nu }}]+\frac{{\rm i}}{2}\nu (t)\{X(t),{{P}_{\xi ,\nu }}\} \right)}^{v}}$, where ${{\left( \cdots \right)}^{v}}$ denotes the vector form, and $X(t)={{{\rm e}}^{{\rm i}{{H}_{S}}t}}\hat{x}{{{\rm e}}^{-{\rm i}{{H}_{S}}t}}$. The equation above can be re-written as

Equation (13)

where $L_{j\lambda }^{0}(t)=\partial \mathcal{L}(t)/\partial {{z}_{j\lambda }}$, and $L_{j\lambda }^{*}(t)=\partial \mathcal{L}(t)/\partial z_{j\lambda }^{*}$ are new super-operators. Note that, because of the interaction picture, these quantities are rotated with respect to HS. In addition, they no longer depend on ${{z}_{j\lambda }}$ or $z_{j\lambda }^{*}$. The analytical average of the last equation leads to the evolution of the reduced density matrix of the system as

Equation (14)

Unfortunately this is an open equation, in the sense that the derivative of ${{\rho }_{s}}(t)$ does not depend only on ${{\rho }_{s}}$ as expected for a proper master equation, but it also depends on the unknown quantities $\mathcal{M}[{{z}_{j\lambda }}P_{t}^{v}]$ and $\mathcal{M}[z_{j\lambda }^{*}P_{t}^{v}]$. However, as noted above, we now know that $\mathcal{M}[\cdots ]=\int {\rm d}\mu ({{z}_{0}})\int {\rm d}\mu ({{z}_{1}})\cdots $, with ${\rm d}\mu ({{z}_{j}})$ given by (9). Then, we can re-express

Equation (15)

where the right-hand side of the expressions have been obtained by considering an integration by parts of the left-hand side. As one can see from these expressions, the evolution of Pvt is intimately related to the evolution of $\partial P_{t}^{v}/\partial z_{j\lambda }^{*}$ and $\partial P_{t}^{v}/\partial {{z}_{j\lambda }}$. Hence, the evolution equation for $\rho _{s}^{v}$ is written in the form

Equation (16)

This is still an open equation that depends on two unknown quantities $A_{\lambda }^{*}(t)$ and $A_{\lambda }^{0}(t)$. To proceed further, we need to re-express these quantities in terms of $\rho _{s}^{v}(t)$. The first step is to calculate their evolution equation, taking into account the consistency conditions $({\rm d}/{\rm d}t)(\partial P/\partial {{z}_{j\lambda }})=(\partial /\partial {{z}_{j\lambda }})({\rm d}P/{\rm d}t)$, and $({\rm d}/{\rm d}t)(\partial P/\partial z_{j\lambda }^{*})=(\partial /\partial z_{j\lambda }^{*})({\rm d}P/{\rm d}t)$ (see appendix for details), together with equation (13)

Equation (17)

If we now compute the average $\mathcal{M}[\cdots ]$ of the latter equation, we will find

Equation (18)

which depends on terms of the form

Equation (19)

For the sake of simplicity, we note that, in the quantities $A_{\lambda }^{\beta }$ and $B_{\lambda {{\lambda }^{\prime} }}^{\beta {{\beta }^{\prime} }}$ ($\beta ,\beta ^{\prime} =*,0$), we skip the indexes j and ${{j}^{\prime} }j$ respectively. To reach an even higher order, we need to compute the evolution of the four quantities $(\partial /\partial z_{j\lambda ^{\prime} }^{*})(\partial P_{t}^{v}/\partial {{z}_{j\lambda }})$, $(\partial /\partial z_{j\lambda ^{\prime} }^{*})(\partial P_{t}^{v}/\partial z_{j\lambda }^{*})$, $(\partial /\partial {{z}_{j\lambda ^{\prime} }})(\partial P_{t}^{v}/\partial {{z}_{j\lambda }})$, and $(\partial /\partial {{z}_{j\lambda ^{\prime} }})(\partial P_{t}^{v}/\partial z_{j\lambda }^{*})$. At this point, we consider the above introduced consistency condition, and the evolution equation (17), to get the following

Equation (20)

Computing the stochastic average of the latter equations gives rise to the following set of equations

Equation (21)

where the $C_{\lambda ^{\prime\prime} \lambda ^{\prime} \lambda }^{\alpha \beta \gamma }$, with $\alpha ,\beta ,\gamma =0,*$ are given by partial derivatives of the projector over $z_{{{j}^{{\prime\prime} }}{{\lambda }^{{\prime\prime} }}}^{\alpha }$, $z_{{{j}^{\prime} }\lambda ^{\prime} }^{\beta }$ and $z_{j\lambda }^{\gamma }$.

The equations (16), (18) and (21) form an open system of mutually dependent equations. Because of this structure, the evolution of $\rho _{S}^{v}(t)$ can be written in terms of a complex series of terms that only depend on $\rho _{S}^{v}$ at past times. The first two terms of the series can be obtained as follows:

  • (i)  
    First integrate analytically (18), obtaining a formal solution for $A_{\lambda }^{0}(t)$, and $A_{\lambda }^{*}(t)$, which comes in terms of different Bs
    Equation (22)
    Note that all initial states of A, B, C, etc are zero;
  • (ii)  
    In a similar way, compute the different Bs by analytically integrating (21). From this, an equation here labelled as (21b) is obtained, that expresses the Bs in terms of As and Cs;
  • (iii)  
    Insert the values of A given by (22) into (21b), which gives rise to an equation here labelled as (21c).
  • (iv)  
    Since (21c) will come again in terms of Bs, replace on its right-hand side the expression of B given by (21b), obtaining (21d).
  • (v)  
    In the resulting equation (21d) replace again the values of A given by (22), as well as the values of B given by (21b).

Equations (21a, b, c) are not written here for simplicity in the presentation, but are straight forwardly derivable following the above steps. The process shall be repeated until an equation for the As is obtained, which comes only in terms of $\rho _{s}^{v}$ at different times, and higher order terms $C_{\lambda ^{\prime\prime} \lambda ^{\prime} \lambda }^{\alpha \beta \gamma }$. Neglecting these higher order terms, the following master equation is obtained

Equation (23)

where $\xi ,\beta ,\eta ,\nu =0,*$, and $j,i,n,m=0,1$, and ${{\rho }^{v}}(t)$ is the reduced density matrix of the system flattened as a vector. Note that because of the restrictions of the sums, we are left with terms proportional to $\alpha _{jj}^{0*}\alpha _{jj}^{0*}$, $\alpha _{jj}^{0*}\alpha _{jj}^{*0}$, $\alpha _{jj}^{*0}\alpha _{jj}^{*0}$, or $\alpha _{jj}^{*0}\alpha _{jj}^{0*}$. We have also defined $\mathcal{U}_{jinm}^{\alpha ,\beta ,\gamma ,\eta }(t,s,{{s}^{\prime} },{{s}^{{\prime\prime} }})=\mathcal{L}_{j}^{\alpha }(t)\mathcal{L}_{i}^{\beta }(s)\mathcal{L}_{n}^{\gamma }({{s}^{\prime} })\mathcal{L}_{m}^{\eta }({{s}^{{\prime\prime} }})$.The super-operators appearing in this expression are $\mathcal{L}_{0}^{0}(t)=\mathcal{L}_{0}^{*}(t)={{\mathcal{L}}_{M}}(t)$, $\mathcal{L}_{1}^{0}(t)=2{{\mathcal{L}}_{M}}(t)$, and $\mathcal{L}_{1}^{*}(t)=\mathcal{L}_{M}^{c}(t)$, where

Equation (24)

As before, the quantity Av in the lhs is a vector with size ${{2}^{2N}}$ flattened from matrix A with size ${{2}^{N}}\times {{2}^{N}}$, where N is the OQS dimension. Finally, in this expression we have also settled

Equation (25)

Note also that $\alpha _{00}^{*0}(t,s)={{(\alpha _{00}^{0*}(t,s))}^{*}}$, and $\alpha _{11}^{*0}(t,s)={{(\alpha _{11}^{0*}(t,s))}^{*}}$, and that $\alpha _{jj}^{00}(t,s)=\alpha _{jj}^{**}(t,s)=0$ for any j.

The number of members of each class, $A_{\lambda }^{\alpha }$ (referred here as first order), $B_{\lambda ^{\prime} \lambda }^{\alpha \beta }$ (second order), $C_{\lambda ^{\prime\prime} \lambda ^{\prime} \lambda }^{\alpha \beta \gamma }$ (third order), etc, are $P\times {{2}^{k}}$, where k is the order number, and P is the number of complex noises, in our case P = 2 (see figure 1). Thus, the master equation (23) is written in terms of an infinite series of terms, wherein the first two terms are in the form of an integral over ${{\rho }_{s}}$ involving one correlation function (5), and the following eight terms are in the form of a triple integral over ${{\rho }_{s}}$ involving two correlation functions. The next terms will come in the form of five integrals involving three correlation functions, and so on and so forth. The derivation of the equation does not rely on any approximation, but the resulting equation unveils a very complex structure with an infinite hierarchy of terms. Hence, for practical uses, the hierarchy should be truncated, which would of course lead to a certain error and to an approximated result.

Figure 1.

Figure 1. Different levels of the hierarchy, corresponding to the average $\mathcal{M}[\cdots ]$ of different orders of the derivative of the stochastic projector Pt with respect to the quantities z0, $z_{0}^{*}$, z1, and $z_{1}^{*}$. The zero order derivative is given by the equation (16), while the first and second order are given by (18) and (21) respectively.

Standard image High-resolution image

For this reason, it is important to take into account the convergence properties of the series. If the system parameters are such that the relative weight of the terms involving a single integral is smaller than that of the terms involving a triple integral, then the case is not tractable because the equation cannot be truncated in a controlled way. In addition, it is worth noticing at this point that the above-described hierarchical structure does not appear because of the use of any expansion approximation, such as the weak coupling. Rather, it is a structure intrinsically attached to the OQS problem. Hence, for systems or parameter ranges where the hierarchy structure does not converge (i.e. successive terms are not smaller than the previous ones), one might conclude that these particular instances shall not be correctly dealt with master equations. Indeed, in these situations the dynamics of the open quantum system appear to alter the environment so dramatically that obtaining a closed equation within the Hilbert space of the system is not feasible in practice. In contrast, for these problems where the subsequent terms are less important, one can safely truncate the equation. This will be seen more clearly in the next section, particularly when analysing the case where the environment is characterized with an exponential correlation function.

Weak coupling and the Markov limit

For consistency, let us check that equation (23) coincides with the master equation up to the second order in the coupling parameter g derived from other approaches. Indeed, each correlation function $\alpha _{jj}^{\alpha \beta }$ will be of the order g2, so that if we just keep second order terms, the equation takes the form

Equation (26)

Replacing the values $\mathcal{L}_{0}^{0}(t)=\mathcal{L}_{0}^{*}(t)={{\mathcal{L}}_{M}}(t)$, $\mathcal{L}_{1}^{0}(t)=2{{\mathcal{L}}_{M}}(t)$, and $\mathcal{L}_{1}^{*}(t)=\mathcal{L}_{M}^{c}(t)$, we find

Equation (27)

with ${{\mathcal{L}}_{M}}(t)$ and $\mathcal{L}_{M}^{c}(t)$ defined in (24). Re-writing the resulting equation in its matrix form, it is identical to the time-convoluted master equation [5, 22] up to second order in g

Equation (28)

As noted above, the equation is convoluted because the reduced density matrix is within the integral, and not local in time as in the so-called convolution-less master equation [5]. However, taking into account that in the interaction picture with respect to the system ${{\rho }_{s}}(s)={{\rho }_{s}}(t)+\mathcal{O}({{g}^{2}})$, and that the ${{\rho }_{s}}(s)$ already appears in second order terms, one can write the latter equation as

Equation (29)

This is already a time local second order master equation, identical to the one derived following the TCL projection operator technique [5], and the second order stochastic Schrödinger equations [23, 24].

If one considers the Markov limit of the master equation (23), by considering $\alpha (t)=\Gamma \delta (t)$, the terms involving three integrals (and two correlation functions) vanish, while the terms involving a single integral give rise to the well-known Lindblad equation [25]

Equation (30)

A hierarchical structure

An alternative to deriving the dynamics of the reduced density operator of the system is to re-write (2) as

Equation (31)

where we have defined $\mathcal{E}={\rm i}{{L}^{-}}{{z}_{0t}}+{\rm i}{{L}^{-}}{{z}_{1t}}-{{L}^{+}}z_{1t}^{*}$, with the super-operators defined as ${{L}^{-}}{{\rho }^{v}}={{(\{\hat{x},\rho \})}^{v}}$, and ${{L}^{+}}{{\rho }^{v}}={{(\{\hat{x},\rho \})}^{v}}$. Importantly, z0t is now a real Gaussian coloured noise, such that $\mathcal{M}[{{z}_{0t}}{{z}_{0s}}]={{\alpha }_{R}}(t-s)$. In addition, $\mathcal{M}[{{z}_{1t}}z_{1s}^{*}]=-{{\alpha }_{I}}(t-s)\theta (t-s)$, which we simplify ${{\alpha }_{I}}(t-s)$, while $\mathcal{M}[z_{1t}^{*}{{z}_{1s}}]=-{{\alpha }_{I}}(t-s)\theta (s-t)$, here denoted as $\alpha _{I}^{*}(t-s)$. Notice that ${{\alpha }_{I}}(0)=\alpha _{I}^{*}(0)$. In order to compute the reduced density matrix, we consider the noise average of (31) as

Equation (32)

where we have defined ${{\rho }_{1}}=\mathcal{M}[{{P}_{1}}]=\mathcal{M}[\mathcal{E}P_{t}^{v}]$, or more in detail

Equation (33)

Here, we have performed the integration by using parts of the Gaussian integrals to find that in general

Equation (34)

a result that is known as the Novikov theorem (see for instance [28]). As before, we proceed further by computing the time evolution of ${{\rho }_{1}}$, which can be written as a sum of the time evolution of three different types of terms, namely: $\mathcal{M}[{{z}_{0t}}P_{t}^{v}]$, $\mathcal{M}[{{z}_{1t}}P_{t}^{v}]$, and $\mathcal{M}[z_{1t}^{*}P_{t}^{v}]$. In the above expression, the integral can be extended to infinity because the stochastic projector is only non-zero in the interval $[0,t]$. Considering this, the time derivative of the first term of (34), $\mathcal{M}[{{z}_{0t}}P_{t}^{v}]$, can be written as

Equation (35)

where we have considered equation (31) to describe $\mathcal{M}\left[ (\delta /\delta {{z}_{0s}})({\rm d}P_{t}^{v}/{\rm d}t) \right]$. The last term in (35) can be rewritten as

Equation (36)

From this, we can express (35) as

Equation (37)

where we have taken into account that $(\delta {{z}_{0t}}/\delta {{z}_{0s}})=\delta (t-s)$, and also that $\rho _{s}^{v}=\mathcal{M}[P_{t}^{v}]$. The time derivative of the two other terms, $\mathcal{M}[{{z}_{1t}}P_{t}^{v}]$ and $\mathcal{M}[z_{1t}^{*}P_{t}^{v}]$, contributing to ${{\rho }_{1}}$ can also be computed in a similar way, so that we can express

Equation (38)

where we have defined ${{V}_{0}}=-{{\alpha }_{R}}(0){{L}^{-}}{{L}^{-}}-{\rm i}{{\alpha }_{I}}(0)({{L}^{-}}{{L}^{+}}+{{L}^{+}}{{L}^{-}})$, and

Equation (39)

Also, we have defined the operator

Equation (40)

To rewrite ${{\rho }_{2}}$ in a different way, we shall consider the fact that ${{L}^{-}}{{L}^{+}}\equiv {{L}^{+}}{{L}^{-}}$, and that following integration by parts we have terms of the form

Equation (41)

Note that the terms in ${{\rho }_{2}}$ that depend on ${{\alpha }_{I}}(t-s)$ and $\alpha _{I}^{*}(t-s)$ can be written in a similar way. Considering this, and also the fact that in all these terms the integrals can be extended to infinity, we can re-write ${{\rho }_{2}}$ as

Equation (42)

After some straightforward calculations, it is found that the evolution of ${{\rho }_{2}}$ can be written similarly to (38)

Equation (43)

where we have defined a super-operator $V={\rm i}{{L}^{-}}{{\alpha }_{R}}(0)+{\rm i}{{L}^{-}}{{\alpha }_{I}}(0)-{{L}^{+}}{{\alpha }_{I}}(0)$. In this expression ${{\hat{\rho }}_{2}}$ is defined as the time derivative of only the correlation functions in ${{\rho }_{2}}$, according to (42). For instance, the first three terms of ${{\hat{\rho }}_{2}}$ can be written as

Equation (44)

Even though the hierarchy appears to become more and more complex, at any order $n\gt 1$, the terms can always be conveniently recast in a similar way, that leads to a general equation of the form

Equation (45)

The difficulty of solving this hierarchy numerically relies on the complexity of the terms ${{\hat{\rho }}_{n}}$, that in principle cannot be written in terms of the hierarchy terms ${{\rho }_{n}}$. One possibility to tackle this problem is to compute these terms with a parallel hierarchy, but this second hierarchy turns out to be coupled to yet another one and so on, and the resulting structure becomes too cumbersome to be dealt with reasonably. A way out of this problem is to consider a simple correlation function of the form $\alpha (t)=({{g}_{R}}+{\rm i}{{g}_{I}}){{{\rm e}}^{-\gamma t}}$. Then, ${{\hat{\rho }}_{n}}=-n\gamma {{\rho }_{n}}$, and the general term of the hierarchy (45) can be written as

Equation (46)

For $n\gt 1$, while for n = 1 we have $({\rm d}{{\rho }_{1}}/{\rm d}t)={{\mathcal{L}}_{S}}{{\rho }_{1}}-\gamma {{\rho }_{1}}+{{V}_{0}}\rho _{s}^{v}+{{\rho }_{2}}$. Interestingly, the latter expression is very similar to the hierarchy obtained by Tanimura in [17, 26] (see also [18] for a review) from a path integral representation. As in the Tanimura derivation, each matrix in the hierarchy (46) depends on the previous and the next matrix. However, the present result differs from the Tanimura one; however, in our case each term ${{\rho }_{n}}$ is coupled to the term ${{\rho }_{n+1}}$ without any preceding super-operator, while in the Tanimura case the link with ${{\rho }_{n+1}}$ appears multiplied by a super-operator. Interestingly, Tanimura shows in [17] that his expansion is non-perturbative, i.e. its different terms do not correspond to the terms found in a perturbative expansion.

In addition, following Tanimura, the hierarchy can be truncated with a terminator of the form ${{\rho }_{N}}(t)\approx {\rm exp} [{{\mathcal{L}}_{S}}t]{{\rho }_{N}}(0)$ for $N\gamma $ sufficiently large [18]. A similar result can be obtained in the present case. To show this, let us consider the Nth order term of the hierarchy, which can formally be integrated as

Here we have considered interaction picture, such that for any n, $\rho _{n}^{I}(t)={\rm exp} [{{\mathcal{L}}_{S}}t]{{\rho }_{n}}(t)$. For $N\gamma \gg 1$, we can approximate ${\rm exp} [-N\gamma (t-s)]\approx (1/(N\gamma ))\delta (t-s)$, which replaced in the above expression leads to $\rho _{N}^{I}(t)\approx \rho _{N}^{I}(0)$ for large $N\gamma $. In the non-interaction picture, this can be written as ${{\rho }_{N}}(t)\approx {\rm exp} [{{\mathcal{L}}_{S}}t]{{\rho }_{N}}(0)$, which brings the same result as in the Tanimura case.

As already discussed in the previous section, the above argumentation shows that only in certain situations, i.e. when $N\gamma \gg 1$, the hierarchy presented in this section can be truncated, and the corresponding reduced density operator obtained.

Numerical example

In the following, we consider the evolution of a two level system with Hamiltonian ${{H}_{S}}=({{\omega }_{s}}/2){{\sigma }_{z}}$, and coupling to an environment with a coupling operator $\hat{x}={{\sigma }_{x}}.$ The environment is assumed to be characterized with a correlation function of the form $\alpha (t)={{{\rm e}}^{-\gamma t}}$. With this example, it will be shown: first, that the derived hierarchy (46) corresponds to the result obtained with the SLN equation (2); second, that for non-Markovian situations ($g\gamma \gt 1$), several members of the hierarchy are necessary to correctly reproduce the non-exponential decay of the system.

Figure 2 displays a comparison between the result obtained from a reduced density matrix using the hierarchy (solid black curve), and the one obtained with the SLN equation (2), in this last case when considering a different number of trajectories M in the sum (11). Indeed, when M = 1000 the curve given by the SLN is almost indistinguishable from the one obtained with the hierarchy, at least up to times of the order of 15. For longer times the statistics of the SLN equation begin to be more demanding, and more trajectories would be needed to get to the correct averaged result.

Figure 2.

Figure 2. Comparison between the SLN equation (2) and the result obtained from a reduced density matrix using the hierarchy (solid black curve). Dotted blue and dotted–dashed orange correspond to the stochastic SLN equation considering the average with M = 500 and M = 1000 trajectories in (3) respectively. The curve for M = 1000 is almost indistinguishable from the one obtained with the hierarchy up to times of the order of 15. For longer times the statistics of the SLN equation begin to be more demanding, and more trajectories would be needed to get to the correct averaged result. We have considered g = 0.1, $\gamma =2$, ${{\omega }_{s}}=0.5$, and as initial condition ${{\rho }_{s}}(0)=|x\rangle \langle x|$, where $|x\rangle =(1/\sqrt{2})(|0\rangle +|1\rangle )$, where $|1\rangle $ and $|2\rangle $ is the basis of the two level system.

Standard image High-resolution image

In figure 3 we show the evolution of $\langle {{\sigma }_{z}}(t)\rangle $ obtained with the hierarchical equations (46) when considering different hierarchical levels $N=1,2,3$ and 4. It can be extracted from the figure that in a strongly non-Markovian situation (upper panel, with $g\gamma \gt 1$), there is a large difference between the results predicted by the hierarchy truncated at N = 1, and higher level truncations. While for N = 1 an exponential decay is observed, higher hierarchy levels unveil strong departures from this exponential form, which are due to the non-Markovian effects. The lower panel reproduces a situation in which, despite of the presence of a relative large coupling, the correlation time of the environment $1/\gamma $ is small as compared to the system evolution time scales $\sim 1/g$, so that the curves for all hierarchy levels practically show an exponential decay. For a decaying rate $\gamma \lt 1$, i.e. longer environment correlation times than the ones considered in the plots, the hierarchy performance is less optimal as there are situations in which it does not converge; at least at the relatively small N here considered.

Figure 3.

Figure 3. Evolution of $\langle {{\sigma }_{z}}(t)\rangle $ considering ${{\omega }_{s}}=0.5$, and $N=1,2,3$ (solid black, dotted green and dotted–dashed orange) levels in the hierarchy. The lower panel corresponds to the case where $\gamma =1.5$, while the lower one to $\gamma =10$. In both cases g = 0.5. The curves for N = 4 are also plotted in both panels but are almost indistinguishable from the N = 3 curve, indicating that the hierarchy has already converged at this level.

Standard image High-resolution image

Conclusions and perspectives

We have derived an evolution equation for the reduced density operator of the system without considering any approximation. Such an equation is written in terms of a hierarchy of terms that involve an increasing number of integrals containing the reduced density operator of the system. Using a second derivation, we have obtained an alternative form to calculate the reduced density matrix, based on a hierarchy of inter-related time local equations. For an exponential correlation function we show that this hierarchy acquires a very similar (although not identical) structure than the one derived in [17, 18] using the path integral representation. Interestingly, a similar hierarchical structure was also obtained in [27] by using a Heisenberg approach. In addition, the obtained hierarchy can be easily extended to deal with environment correlation functions that can be written as a sum of exponentials.

The complexity of the obtained hierarchical structures suggests that it may not be possible to deal with certain problems using a master equation approach or considering the dynamics in the reduced Hilbert space of the system. These problems may therefore have to be treated necessarily by considering an evolution equation for the total system (i.e. OQS and environment). In addition, the obtained hierarchy may be useful to gain some insight into the problem sampling of the SLN equation, and particularly on the sign problem that gives rise to a poor convergence for long times.

Finally, we note that the procedure here discussed, can also be used by considering as a starting point instead of (2), an SSE of the form [20, 27]

Equation (47)

where $G(z_{i}^{*}{{z}_{i+1}}|{{t}_{i}}{{t}_{i+1}})=\langle {{z}_{i}}|{{\mathcal{U}}_{I}}({{t}_{i}},{{t}_{i+1}})|{{z}_{i+1}}\rangle $ is the reduced propagator of the system. Here $|{{z}_{i+1}}\rangle $ represents the initial state of the environment at time ${{t}_{i+1}}$, and $|{{z}_{i}}\rangle $ its final state at time ti. The latter is the generalization to an arbitrary initial state of the stochastic Schrödinger equation derived in [28, 29] for an environment initially at zero temperature, i.e. ${{z}_{i+1}}=0$. Using (47) opens the possibility to build the structure of master equations for problems where the coupling operator is non-Hermitian, the OQS is composed of many particles, or a more general initial state is considered.

Acknowledgments

The author gratefully acknowledges D Alonso, and U Schollwöck for interesting discussions, and D Alonso, JI Cirac, A Ekert and U Schollwöck for encouragement and support. This project was financially supported by the Nanosystems Initiative Munich (NIM) (project No. 862050-2) and partially by the Spanish MICINN (Grant No. FIS2010-19998).

Appendix. Test of the consistency condition

Let us test the consistency condition with a very simple example. Let us assume the following evolution equation for the stochastic projector:

Equation (48)

where ${{L}_{\lambda }}(t)={{g}_{\lambda }}(t)L(t)$, and L(t) is a matrix within the Hilbert space of the system that does not depend on any noise quantity ${{z}_{\lambda }}$. A partial derivation of the latter equation with respect to ${{z}_{\lambda ^{\prime} }}$ is given as

Equation (49)

Integrating now (48), we find

Equation (50)

and deriving this solution with respect to ${{z}_{\lambda ^{\prime} }}$, we find

This equation, once derived with respect to t, gives rise to the same result as (49), proving that the time derivative and the partial derivative with respect to ${{z}_{\lambda }}$ commute, and therefore can be safely interchanged.

Please wait… references are loading.
10.1088/1751-8113/48/14/145202