Multiple-scale analysis of open quantum systems

D. N. Bernal-Garcíaa,b,∗, B. A. Rodríguezb, H. Vinck-Posadaa

aGrupo de Superconductividad y Nanotecnología, Departamento de Física, Universidad Nacional de Colombia, Ciudad Universitaria, K. 45 No. 26-85, Bogotá D.C., Colombia bGrupo de Física Atómica y Molecular, Instituto de Física, Facultad de Ciencias Exactas y Naturales, Universidad de Antioquia, Calle 70 No. 52-21, Medellín, Colombia

Abstract In this work, we present a multiple-scale perturbation technique suitable for the study of open quantum systems, which is easy to implement and in few iterative steps allows us to find excellent approximate solutions. For any time-local quantum , whether markovian or non-markovian, in Lindblad form or not, we give a general procedure to construct analytical approximations to the corresponding dynamical map and, consequently, to the temporal evolution of the . As a simple illustrative example of the implementation of the method, we study an atom-cavity system described by a dissipative Jaynes-Cummings model. Performing a multiple-scale analysis we obtain approximate analytical expressions for the strong and weak coupling regimes that allow us to identify characteristic time scales in the state of the physical system. Keywords: Open quantum systems, Time-local quantum master equations, Multiple-scale analysis, Time-dependent perturbation technique

1. Introduction use due to the complicated structure of the mathematical ex- pressions involved. Therefore, it is necessary to look for other The study of open quantum systems (OQS) has become approaches to solve a wider spectrum of OQS that otherwise mandatory when thinking about any technological application would be very difficult to study. based on quantum mechanics [1]. However, it is not limited to For this purpose, there are multiple more sophisticated an- applications, but has been crucial in important contributions to alytical and numerical methods, including perturbative expan- fundamental physics [2]. In fact, any area of physics interested sion of dissipative terms [9, 10], numerical renormalization group in the quantum description of a system that interacts with its [11, 12], density matrix renormalization group [13], effective environment needs the theory of open quantum systems. Such liouvillian superoperators [14, 15], Feshbach projection tech- is the case of quantum optics [3], quantum information [4], and nique [16], variational matrix product operators [17], Keldysh quantum measurement [5], just to give some examples. There- Green functions [18, 19], and Lindblad perturbation theory [20, fore, finding appropriate mathematical methods to solve the dy- 21]. These methods are either focused on obtaining solutions namics of OQS is as relevant as the study of the physical sys- to the steady state or are numerical techniques to solve the dy- tems themselves. namics of OQS. In any case, they do not provide approximate Traditionally, the dynamics and the steady state of an open analytical expressions that describe the dynamics of OQS in quantum system are calculated by means of exact diagonal- both the transient and steady states. To this end, it is natural ization of the Liouville superoperator (liouvillian) [6], or by to consider a time-dependent perturbation theory; however, a simulations with quantum trajectories [7]. However, both ap- usual perturbative expansion is often an insufficient approxima- proaches are very limited by the growth of the Hilbert space, tion to the dynamics of classical and quantum systems, since it and quickly become challenging numerical problems. The chal- inevitably leads to secular terms in the approximate solutions, arXiv:1805.12340v3 [quant-ph] 29 Apr 2019 lenge increases when the liouvillian is time-dependent, where i.e., terms that grow as powers of the time variable, thus limit- in general it is not possible to make an exact analytical diag- ing the range of validity of the approximate analytical expres- onalization to obtain the exponential solution to the quantum sions. Secular solutions appear in differential equations with master equation. In this case, it is common to use a Magnus se- constant coefficients, when the non-homogeneous term is itself ries [8], which is an analytical expression that allows express- a solution to the associated homogeneous differential equation, ing the exponential solution to a linear differential equation of as it happens in the regular time-dependent perturbation theory linear operators as a series expansion. Nevertheless, although [22]. In particular, in classical mechanics this type of asymp- it provides approximate analytical solutions, it is difficult to totic expansion provides secular solutions for systems known to evolve harmonically, and in quantum mechanics it gives secular terms that destroy the canonical commutation relations among ∗Corresponding author Email address: [email protected] (D. N. Bernal-García ) approximate operators [23]. As an alternative to the regular time-dependent perturbation

Preprint submitted to Elsevier April 30, 2019 theory in OQS, we propose a perturbative approach based on theory of open quantum systems can skip to the next section, the multiple-scale analysis (MSA) of time-local master equa- where the perturbative technique is described in detail. tions (TLMEs). MSA methods have been around for a long The study of OQS appears as a more realistic approach to time as techniques to avoid secular terms in the solution of or- the description of the quantum dynamics of systems that are in- dinary differential equations. In fact, the first of these was the teracting in a non-negligible way with an environment or reser- Poincaré-Von Zeipel method, which was developed at the end voir. Formally, system and reservoir constitute a universe that of the 19th century for the study of celestial dynamics [24]. is itself a closed quantum system with Hilbert space HT , and However, even though it provides approximate periodic solu- evolving unitarily under the total hamiltonian HT (t). There- tions without secular terms, the asymptotic expansions do not fore, the evolution from an initial time t0 to a posterior time t of converge and this limits their usability. Nevertheless, this and the total density matrix ρT describing the physical state of this other multiple-scale techniques were widely used, although the universe, will be described in general by the unitary operator mathematical justification for its operation was not well under- U(t, t0), such that (~ = 1), stood until the 1960s, when multiple-scale modeling was ex- † plained using coordinate transformations and invariant mani- ρT (t) = U(t, t0) ρT (t0) U (t, t0), (1) ( Z t ) folds [25, 26]. Since then, MSA has been used successfully in 0 0 the solution of linear and weakly nonlinear differential equa- U(t, t0) = T exp i HT (t ) dt ; (2) t tions describing the dynamics of classical systems [22, 27–29]; 0 and more recently, solving the dynamics of quantum systems, with T the Dyson time-ordering operator. such as the quantum anharmonic oscillator [30–32], a wide range However, OQS would be very difficult to study if all degrees of quantum optical systems [23, 33–35], and time-dependent of freedom are considered. Thus, a partial trace is carried out problems like the dynamical Casimir effect [36–39]. over the reservoir, reducing the problem to an efficient set of However, in the existing implementations of MSA tech- parameters and operators describing the system dynamics, niques, the solutions obtained must be subject to additional n † o constraints so that each term in the perturbative expansion is ρ(t) = TrR U(t, t0) ρT (t0) U (t, t0) . (3) physically consistent [40]. This limitation does not allow the construction of a general scheme, but for each physical sys- If the reservoir is in a steady state with a given spectral decom- P ED P tem a different implementation must be made. For example, position ρR = j P j r j r j , j P j = 1, and the system and its in the MSA solution of the Heisenberg equation of motion for environment are initially uncorrelated ρT (t0) = ρ(t0) ⊗ ρR, we the quantum anharmonic oscillator [32], the technique does not can explicitly trace over the reservoir and separate the evolution guarantee that the canonical commutation relations among the of the system, approximate creation and annihilation operators are satisfied, X E D † but in each iteration, the integration constants must be adjusted ρ(t) = P j hri| U(t, t0) r j ρ(t0) r j U (t, t0) |rii ; (4) according to this requirement, which implies a step-by-step de- i, j velopment that can hardly be generalized. Nevertheless, as we which can be written as, will show, it is possible to construct a multiple-scale perturba- X tion technique (MSPT) that is free of such restrictions, and still † ρ(t) = Ki j(t, t0) ρ(t0) Ki j(t, t0), (5) preserves the physical properties of the system at each step of i, j the perturbative iteration. Hence, the aim of this work is to ex- p E tend MSA to the study of OQS described by TLMEs. Our treat- where it was defined Ki j(t, t0) = P j hri| U(t, t0) r j . Eq. (5) ment is not limited to a specific physical system; on the con- is the Kraus decomposition of the system density matrix evolu- trary, it is applicable to a wide range of perturbations and OQS. tion [41], with Ki j(t, t0) the Kraus operators satisfying the time- Besides, it is not just a technique that avoids secular terms, but independent completeness relation [42], a tool to obtain approximate solutions of rapid convergence and X † which are stable with respect to initial conditions. Ki j(t, t0)Ki j(t, t0) = 1. (6) This paper is organized as follows: In Sec.2, we describe i, j time-local quantum master equations for the density matrix op- erator, which are the type of equations we focus on in this work. It is clear, that the operator sum representation of the time- In Sec.3, we describe the multiple-scale perturbation technique evolution in Eq. (5) is not unitary in general, however, it must for open quantum systems. In Sec.4, we give an illustrative ex- describe a valid physical time-evolution mapping a physical ρ t ρ t ample of the implementation of the method. Finally, in Sec.5, state ( 0) into a physical state ( )[43]; where a density ma- we conclude. trix describing a state must be positive semidefinite and trace preserving [44]. Therefore, the operator sum decomposition must represent a completely-positive trace-preserving (CPTP) 2. Time-local quantum master equations dynamical map Φ(t, t0)[45, 46], This section aims to provide a brief and clear definition X ρ t = Φ t, t ρ t = K t, t ρ t K† t, t . of time-local quantum master equations, with the intention of ( ) ( 0) ( 0) i j( 0) ( 0) i j( 0) (7) i, j making the document self-contained. Readers familiar with the 2 The complete positivity ensures that, for all t > t0 in the so- A detailed derivation of the Lindblad master equation in Eq. lution domain, the density matrix describing the physical state (12) can be found in Refs. [3, 54]. of the system is positive semidefinite. This is a consequence of On the other hand, when the Born-Markov approximation the representation theorem for quantum operations which states can not be made, we are left with a time-dependent generator that, given a density matrix that evolves according to the op- L(t), which has been shown that in general can be brought into erator sum representation in Eq. (5) with the operators satis- a canonical Lindblad form with time-dependent coefficients and fying the completeness relation in Eq. (6), then the dynami- operators [55, 56], Φ t, t cal map ( 0) describing its evolution is completely positive   [47, 48]. Meanwhile, the trace-preserving property, Tr{ρ(t)} = ρ˙ = −i H(t), ρ { } ( ) Tr ρ(t0) = 1, is a consequence of the completeness relation for X † 1  † †  + γn(t) On(t) ρ O (t) − O (t) On(t) ρ + ρ O (t) On(t) , Kraus operators in Eq. (6). n 2 n n In general, the time-evolution of the system density matrix n (14) shown in Eq. (7) can be described by a TLME,

The rates γn(t) can only be temporarily negative in order to pre- ρ˙(t) = L(t) ρ(t), (8) serve the complete positivity of the map, in addition, γn(t) ≥ 0 t ffi where the liouvillian L(t) is a linear non-hermitian superopera- for all in the solution domain, is a necessary and su cient condition for the divisibility of the quantum operation and, con- tor and the infinitesimal generator of the dynamical map Φ(t, t0) [49]. Therefore, it is satisfied, sequently, for the markovianity of the system it describes [44]. The above description is based on the assumption of an ini-  R t 0 0  L(t ) dt tially uncorrelated state between system and reservoir; however, ρ(t) = Φ(t, t ) ρ(t ) = T e t0 ρ(t ). (9) 0 0 0 in a more general scenario, a standard approach to the dynam- ics of OQS leads to the Nakajima-Zwanzig nonlocal quantum It is important to note that the dynamics described by TLMEs master equation, can be distinguished between markovian and non-markovian, Z t where markovianity is defined by means of the divisibility of 0 0 0 the dynamical map [44, 50, 51]; where divisibility is defined ρ˙(t) = K(t − t ) ρ(t ) dt , (15) t0 as, Φ(t, t0) = Φ(t, s) Φ(s, t0)(t ≥ s ≥ t0), for every (t, s, t0) in the solution domain. where K(t) is a memory kernel that accounts for the memory In many applications, the Born-Markov approximation is effects of the system, i.e., how the rate of change of the state justified and, consequently, to have master equations with time- at a given time t depends on the physical state at all previous independent liouvillians [3], times. The nonlocal master equation (15) represents a chal- lenge when looking for solutions, even approximate ones, nev- ρ˙(t) = L ρ(t). (10) ertheless, through the time convolution-less projection operator technique it is possible to bring a nonlocal master equation to ρ t = L(t1−t0) ρ t Thus, ( 1) e ( 0) and the evolution will be homoge- a time-local form,ρ ˙(t) = L(t) ρ(t); although, this approach of- t = t − t neous in time given by the single parameter, 1 0. Hence, ten leads to significant problems, such as the lack of complete the solution to Eq. (10) is given by, positivity of the corresponding dynamical map [48]. However, L t an alternative that leads to perfectly regular dynamics was pro- ρ(t) = Φ(t) ρ(0) = e ρ(0), (11) posed in Ref. [57], where it was shown that any solution to the nonlocal Eq. (15) always satisfies a TLME with a generator that with t0 = 0 and, therefore, the system dynamics is represented by a quantum dynamical semigroup, i.e., a single-parameter dy- depends explicitly on the initial time,ρ ˙(t) = L(t − t0) ρ(t). namical map Φ(t) satisfying the semigroup property Φ(t + s) = Note that as a consequence of the aforementioned, it is not Φ(t) Φ(s)(t, s ≥ 0) [52]. The most general type of markovian possible in general to reduce the dynamical map Φ(t, t0) to a and time-homogeneous master equation describing non-unitary single-parameter representation, i.e., to a quantum dynamical dynamics of the density matrix ρ(t), is the celebrated Gorini- semigroup. Therefore, here we focus on systems whose dynam- Kossakowski-Lindblad-Sudarshan (GKLS) master equation, or ical map exists and is CPTP, without further additional condi- simply, the master equation in Lindblad form [47, 53], tions. To simplify the notation, we assume that the initial time is always zero (t0 = 0) and then, Φ(t, t0 = 0) = Φ(t); although,   X ρ˙(t) = LGKLS ρ(t) = −i H, ρ(t) + γn D[On] ρ(t), (12) in general, we do not work with single-parameter dynamical n maps. † 1  † †  In summary, the description of the dynamics of OQS in D[On] ρ(t) = On ρ(t) O − O On ρ(t) + ρ(t) O On , (13) n 2 n n terms of TLMEs and CPTP dynamical maps is based in prin- ciple on the assumption of an initially uncorrelated state be- where H is the system’s hamiltonian, and γn are the param- tween system and reservoir. Nevertheless, by using the appro- eters associated with the different dissipative and decoherent priate techniques, more general OQS can also be described by processes described by the operators On. Besides, if γn > 0, TLMEs. On the other hand, the markovianity of the system is L t then the dynamical map Φ(t) = e GKLS is completely-positive. defined by means of the divisibility of the dynamical map; thus, 3 if the liouvillian L is time-independent, then, the dynamics is and separating according to the powers of the perturbation pa- trivially markovian; on the other hand, if L is time-dependent, rameter, the markovianity depends on whether the map is divisible or 0 (0) (0) not. However, the technique presented in this work, is suitable O(α ): D0% − L0% = 0, (20a)   to obtain approximate analytical solutions to TLMEs describing 1 (1) (1) (0) (0) O(α ): D0% − L0% = − D1% − L1(τ0)% , (20b) both markovian and non-markovian dynamics. 2 (2) (2) O(α ): D0% − L0% =   − D (1) − L (1) − D (0) 3. Multiple-scale perturbation technique 1% 1(τ0)% 2% , (20c) . As pointed out above, the key goal of this research is to . illustrate a perturbation technique based on the multiple-scale n (n) (n) O(α ): D0% − L0% = analysis of open quantum systems whose evolution is repre- Xn−2 sented by time-local quantum master equations,  (n−1) (n−1) (m) − D1% − L1(τ0)% − Dn−m% , m= ρ˙(t) = L(t) ρ(t). (16) 0 n ≥ 2; (20d) Such MSPT is essentially a derivative expansion method [22, 32], which promotes the time variable t to an infinite collection we can determine the time-evolution of the density matrix by means of the iterative solution of Eqs. (20). Hence, from the t = {τ0, τ1, τ2,... } of independent variables τn, each one a func- n zero-order equation (20a), performing a Dyson integration as tion of the perturbative parameter α such that τn = α t, n = 0, 1, 2,... , with 0 < α < 1. In consequence, the temporal shown in Appendix A, derivative transforms into an infinite sum provided by the chain n o %(0)(t) = eL0τ0 C (t ) %(0)(0), (21) rule, 0,1 1

where t1 follows the notation, tk = {τk, τk+1,... } (k ≥ 1), i.e., d X n ∂ = α Dn, Dn = . (17) t t ∂τ the new collection k includes all the time-scales from τk. Be- d n n sides, C0,1(t1) is a constant superoperator with respect to the τ0 The permitted values of the perturbative parameter α guarantee variable, whose dependence with respect to each time variable in tk will be determined order-by-order in the perturbation the- the positivity of the time scales τn and their shortening with respect to the growth of the discrete variable n, i.e., when- ory. Eq. (21) suggests that the description of the perturbative technique can also be performed by means of the dynamical ever n grows τn will describe increasingly rapid changes in t the physical quantity that is being resolved. In our implemen- map Φ( ), then if, n o tation of MSA to TLMEs, we introduce an extended density (0) (0) L0τ0 % (t) = Φ (t) ρ(0) = e C0,1(t1) ρ(0), (22) matrix ρ(t) = ρ(τ0, τ1, τ2,... ), considered an extension of the t true density matrix ρ(t), which is retrieved by restricting ρ( ): we can make describe the MSPT regardless of the initial con- ρ t = ρ t | n ( ) ( ) τn=α t. Expanding the density matrix asymptotically, ditions ρ(0). However, a description in terms of the extended we have, dynamical map Φ(t) will require to represent it as an asymp- X totic expansion as well, ρ(t) = αn%(n)(t). (18) n X Φ(t) = αnΦ(n)(T), (23) It is worth mentioning that only %(0) represents a valid density n matrix. In fact, it will be responsible of the unit-trace (all the where only Φ(0)(t) represents a valid dynamical map satisfying other terms of the expansion are traceless), and the satisfaction the initial condition Φ(0)(0) = 1, while Φ(n)(0) = 0 for n > 0. of initial conditions, %(0)(0) = ρ(0) (while %(n)(0) = 0 for n > 0). The time-evolution of the system density matrix will be given On the other hand, we can separate the liouvillian L(t) in two by ρ(t) = Φ(t) ρ(0), and therefore, parts: A solvable superoperator L0 and a perturbation L1(t), L t = L + α L t X n n such that ( ) 0 1( ); besides, in the extended time ρ(t) = α Φ( )(t) ρ(0), (24) L → L L representation we would have, (t) (τ0); thus, (τ0) = n L + α L (τ ). 0 1 0 %(n)(t) = Φ(n)(t) ρ(0). (25) Putting the above expressions together in the TLME, Besides, we can separate the unperturbed solution from the ex- ρ˙(t) = L(τ0) ρ(t), X X pansion, which is analogous to moving to an interaction picture, m+n (n) n (n) (n) L0τ0 α Dm % (t) = (L0 + αL1(τ0) ) α % (t); (19) Φ (t) = e An(t), and then, m,n n n o (n) L0τ0 % (t) = e An(t) ρ(0). (26)

4 Zero-order equation

D0 A0 = 0

⇓ (0,0) % (τ0)

First-order First-order equation solvability condition

D0 A1 = Ψ1 − Ψ1 = 0 h iτ0 ⇓ (1,0) (0,1) ⇓ % (τ0, τ1) % (τ0, τ1)

Second-order Second-order solvability condition equation

D A = Ψ Ψ2 = 0 0 2 − 2 h iτ0 ⇓ (0,2) ⇓ (2,0) % (τ0, τ1, τ2), % (τ0, τ1, τ2) (1,1) % (τ0, τ1, τ2)

.

nth-order solvability condition nth-order equation Ψn = 0 h iτ0 D0 An = Ψn − (0,n) ⇓ % (τ0, . . . , τn), ⇓ (1,n 1) % − (τ , . . . , τ ), %(n,0)(τ , ... , τ ) 0 n 0 n . . . (n 1,1) % − (τ0, . . . , τn)

.

Figure 1: Flowchart schematizing the solution algorithm. We solve in each step first the solvability condition associated with the equation to solve. The solvability conditions will intercalate the equations for the different orders, and must be fulfilled if we require absence of secular terms.

5 (1) (0,1) (1,0) In this way, in terms of the An = An(t), Eqs. (20) transform to, First-order: ρ (τ0, τ1) = % (τ0, τ1) + α % (τ0, τ1). 0 ρ(2) τ , τ , τ = O(α ): D0 A0 =0, (27a) Second-order: ( 0 1 2) 1  %(0,2)(τ , τ , τ ) + α %(1,1)(τ , τ , τ ) + α2 %(2,0)(τ , τ , τ ). O(α ): D0 A1 = − D1 A0 − F1,0(τ0)A0 , (27b) 0 1 2 0 1 2 0 1 2 2  (n) O(α ): D0 A2 = − D1 A1 − F1,0(τ0)A1 − D2 A0, (27c) nth-order: ρ (τ0, . . . , τn) = n . P m (m,n−m) . α % (τ0, . . . , τn), n ≥ 1. m=0 n  O(α ): D0 An = − D1 An−1 − F1,0(τ0)An−1 Therefore, the solution algorithm will go as follows (as schema- Xn−2 tized in Fig.1): From the zero-order equation, Eq. (27a), − D (0,0) n−m Am, % (τ0) is calculated; then, the non-homogeneous term Ψ1 in m=0 the first-order equation, Eq. (27b), defines the first order solv- n ≥ 2; (27d) h i ability condition, Ψ1 τ0 = 0, which improves the previous so- lution providing the functional dependence with respect to the −L L where, F (τ ) = e 0τ0 L (τ ) e 0τ0 . The solution proce- (0,1) 1,0 0 1 0 τ1 variable, % (τ0, τ1). Next, the first order equation is solved, dure will be iterative order to order, satisfying solvability con- (1,0) giving % (τ0, τ1). The approximate density matrix until this ditions before solving each order equation. Solvability condi- (1) (0,1) (1,0) point is, ρ (τ0, τ1) = % (τ0, τ1) + α % (τ0, τ1). In the same tions avoid the appearance of secular solutions, making zero the way, we can solve each order, improving the previous solution terms that resonate with the homogeneous equation. We con- through the solvability condition, and then adding a new term struct this conditions by making time averages with respect to to the asymptotic expansion. the τ0 variable and separating the terms that oscillate with zero On the other hand, when we return to the usual time variable frequency, which are the ones that resonate with the homoge- t, we will have, neous solution. Hence, the solution to Eq. (27a), which gives us Zero-order: ρ(0)(t) = %(0,0)(t) the τ0 dependence of A0, will be substituted into equation Eq. (27b), and here the solvability condition must be fulfilled before First-order: ρ(1)(t) = %(0,1)(t) + %(1,0)(t) solving the first-order partial differential equation. Thus, the , , , first-order equation (27b) can be written as D A = −Ψ , with Second-order: ρ(2)(t) = %(0 2)(t) + %(1 1)(t) + %(2 0)(t)  0 1 1 Ψ = D A − F (τ )A ; and the solvability condition will n 1 1 0 1,0 0 0 P − n ρ(n)(t) = %(m,n m)(t), n ≥ 1. be, hΨ1iτ = 0, which gives us the τ1 dependence of A0. After th-order: 0 m=0 this, the improved first-order equation must be solved. There- fore, the solvability conditions up to nth-order in the asymptotic Now, we show the general expressions obtained from the expansion are, solution to Eqs. (27) and (28) for any , with an arbitrary perturbation, described by a TLME. 1 O(α ): hΨ1iτ = D1 A0 − F1,0(τ0)A0 = 0, (28a) 0 τ0 Zero-order. From Eq. (27a), as mentioned above, we have, 2 O(α ): hΨ2iτ = D1 A1 − F1,0(τ0)A1 +D2 A0 τ = 0, (28b) 0 0 A (t) = C (t ). (30) . 0 0,1 1 . If we truncate to this order, the satisfaction of the initial condi- n (0) O(α ): hΨniτ = 0 tions force that, C0,1 = 1. Then, the zero-order density matrix Xn−2 will be, hD − D i 0 (0) (0,0) 1 An−1 F1,0(τ0)An−1 + n−m Am τ0 = , ρ (τ0) = % (τ0), (31) m=0 (0,0) K0τ0 n ≥ 2, (28c) % (τ0) = e ρ(0), K0 = L0. (32)

where the time-averaging is given by, First-order. The solvability conditions will provide the inte- gration constants to the next level of precision, i.e., the func- Z T 1 n tional dependence with respect to the next time variable. Then, hΨni = Ψn dτ , (29) τ0 0 τ Tn 0 from the solvability condition in Eq. (28a), we determine the 1 dependence of C0,1(t1), namely, the zero-order solution is im- being Tn the period of the solvability condition, which can ap- proved to a next level of precision. Now, looking after the first proach infinity if there is no well-defined period. solvability condition, we are going to separate the superopera- In terms of the density matrix, it is possible to recognize the τ tor F1,0( 0) in a part that oscillates with zero frequency, F1,0 τ0 , structure of the asymptotic sum up to a given level of precision. which corresponds to the temporal average of the signal in a pe- Thus, being ρ(n) the density matrix truncated to the nth-order, riod T1,0, and another, eF1,0(τ0), that contains all the terms that (m,n) and % the nth correction to the mth term of the asymptotic oscillate with nonzero frequencies, expansion. For each order in the perturbative expansion, we −K0τ0 K0τ0 have, F1,0(τ0) = e L1(τ0) e (33)

(0) (0,0) F1,0(τ0) = F1,0 + eF1,0(τ0); (34) Zero-order: ρ (τ0) = % (τ0). τ0 6 Z T 1 1,0 t = τ τ . In addition, the structure of the A2( ) superoperator is calcu- F1,0 τ0 F1,0( 0) d 0 (35) T1,0 0 lated from Eq. (27c), and then, until second-order,

K τ K τ K1τ1 t = 1 1 2 2 t , C0,1(t1) = e C0,2(t2), K1 = F1,0 τ . (36) A0( ) e e C0,3( 3) (52) 0 n o t K1τ1 K2τ2 t Furthermore, A1(t) is calculated from Eq. (27b), and we have A1( ) = B1,0(τ0) e e C0,3( 3) n o until first-order, K1τ1 K2τ2 + e B1,1(t1) e C0,3(t3) + C1,2(t2) , (53)

K1τ1 A0(t) = C0,1(t1) = e C0,2(t2), (37) A2(t) = B2,0(τ0)C0,1(t1) + B1,0(τ0)C1,1(t1)

A1(t) = B1,0(τ0)C0,1(t1) + C1,1(t1) + C2,1(t1),

K1τ1 K1τ1 K2τ2 = B1,0(τ0) e C0,2(t2) + C1,1(t1); (38) A2(t) = B2,0(τ0) e e C0,3(t3) n K1τ1 K2τ2 Z τ0 + B1,0(τ0) e B1,1(t1) e C0,3(t3) + C1,2(t2) τ = e τ0 τ0 . B1,0( 0) F1,0( 0) d 0 (39) t 0 + C2,1( 1); (54)

If we truncate to this order, the determination of the integration Z τ0 0 0 constants will be subject to the initial conditions, i.e., Φ(0) = 1. B2,0(τ0) = eF2,0(τ0) dτ0 . (55) (1) 0 Therefore, since B1,0(0) = 0 (in general, Bm,n(0) = 0), C0,2 = 1 (1) If we truncate to this order, C(2) = 1,C(2) = 0, and C(2) = 0. and C1,1 = 0. Then, the first-order density matrix will be, 0,3 1,2 2,1 Then, the second-order density matrix will be, ρ(1) τ , τ %(0,1) τ , τ %(1,0) τ , τ ( 0 1) = ( 0 1) + ( 0 1); (40) (2) (0,2) (1,1) ρ (τ0, τ1, τ2) =% (τ0, τ1, τ2) + % (τ0, τ1, τ2) (2,0) + % (τ0, τ1, τ2); (56) (0,1) K0τ0 K1τ1 % (τ0, τ1) = e e ρ(0), (41)

(1,0) K0τ0 K1τ1 % (τ0, τ1) = e B1,0(τ0) e ρ(0). (42) (0,2) K0τ0 K1τ1 K2τ2 % (τ0, τ1, τ2) = e e e ρ(0), (57) n Second-order. From the second-order solvability condition in (1,1) K0τ0 K1τ1 K2τ2 % (τ0, τ1, τ2) = e B1,0(τ0) e e Eq. (28b), the τ2 dependence of C0,2(t2) and the τ1 depen- o K1τ1 K2τ2 dence of C1,1(t1) are provided. Then, defining the new quantity + e B1,1(τ1) e ρ(0), (58) n F2,0(τ0), we have, (2,0) K0τ0 K1τ1 K2τ2 % (τ0, τ1, τ2) = e B2,0(τ0) e e K K o F , (τ ) = F , (τ )B , (τ ) − B , (τ ) F , (43) 1τ1 2τ2 2 0 0 1 0 0 1 0 0 1 0 0 1 0 τ0 + B1,0(τ0) e B1,1(τ1) e ρ(0). (59)

F2,0(τ0) = F2,0 + eF2,0(τ0); (44) τ0 nth-order. Similar to the previous steps, in the nth iteration, the Z T nth solvability condition, Eq. (28c), allows us to determine the 1 2,0 F2,0 = F2,0(τ0) dτ0 . (45) temporal dependence to a higher order of all the superoperators τ0 T 2,0 0 that appeared previously, and then the nth-order equation (27d) will allow finding the structure of the An(t) superoperator. On

−K1τ1 K1τ1 the other hand, we can find more general expressions if we de- F1,1(τ1) = e F2,0 τ e (46) 0 fine, τ = + e τ F1,1( 1) F1,1 τ1 F1,1( 1); (47) K0τ0 Cm,0(t) = e Am(t). (60) Z T 1 1,1 F1,1 = F1,1(τ1) dτ1 . (48) τ1 Consequently, we determine the general structure of A (t), T1,1 0 m Xm t t K2τ2 Am( ) = B(m−i),0(τ0)Ci,1( 1), B0,0(τn) = 1; (61) C0,2(t2) = e C0,3(t3), K2 = F1,1 , (49) τ1 i=0 K1τ1  C1,1(t1) = e B1,1(t1)C0,2(t2) + C1,2(t2) , n o K1τ1 K2τ2 m = e B1,1(t1) e C0,3(t3) + C1,2(t2) ; (50) X Knτn Cm,n(tn) = e B(m−i),n(τn)Ci,n+1(tn+1), Z τ1 i=0 τ = e τ0 τ0 . B1,1( 1) F1,1( 1) d 1 (51) τ = 1. 0 B0,n( n) (62)

Hence, the superoperators Bm,n(τn)(m > 0, n ≥ 0) and Kn (n > 0) can be calculated up to the desired order from the following

7 expression, which summarizes Eqs. (27) and (28), resonance (ωc ≈ ωa), if g is much smaller than the natural fre- quencies of the system (ωc, ωa), then, the rotating wave ap- n− X1 proximation (RWA) is justified and the interaction hamiltonian D τ t = 0 Bn−i,0( 0)Ci,1( 1) is, i=0 n−1  †  X g V = g a σ− + a σ+ . (66) F1,0(τ0) Bn−i−1,0(τ0)Ci,1(t1) i=0 Thus, the complete JC hamiltonian will be, Xn−1 Xm − D t ≥ Bn−i−1,0(τ0) n−m Ci,1( 1), n 1. (63) H = H0 + g V, m=0 i=0 † ωa  †  H = ωc a a + σz+ g a σ− + a σ+ . (67) It is worth mentioning, as is shown in Appendix B, that all su- 2 peroperators Kn (n ≥ 0) commute between them, [Kn+1 , Kn] = On the other hand, in presence of leakage of photons through 0. This means, that the zero term in the asymptotic expansion the cavity mirrors and continuous and incoherent pumping of of the dynamical map, the TLS, the master equation in Lindblad form is,

Φ(0) t = K0τ0 K1τ1 ... Knτn ...   ( ) e e e (64) ρ˙(t) = −i H, ρ(t) + κ D[a] ρ(t) + P D[σ+] ρ(t), (68) is an exponential factorization of commuting superoperators. where κ is the rate of leakage of photons out of the cavity and On the other hand, if L0 and L1 commute, the MSA solution P is the amplitude of the continuous pump, and explicitly (ρ = L t αL t will correspond to the exact solution, Φ(t) = e 0 e 1 ; which ρ(t)), is reached after solving the first-order solvability condition in   κ n † † † o the MSPT. ρ˙ = −i H, ρ + 2 a ρ a − a a ρ − ρ a a In the following section, we show a simple example of im- 2 P plementation of the MSPT, this with the intention of clarifying + {2 σ+ ρ σ− − σ− σ+ ρ − ρ σ− σ+}. (69) the ideas presented. 2 If there is low pumping in the system (P  g, κ), according to 4. Illustrative example: The dissipative Jaynes-Cummings the relationship between g and κ there are two clearly differen- model tiated operating regimes [60–62]. The first one, known as the strong coupling (SC) regime, is characterized by the fact that As a simple example of the implementation of the MSPT, let the interaction constant is bigger than the system dissipation us consider the Jaynes-Cummings (JC) model with dissipation, rate (g > κ). In the second one, known as the weak coupling which describes the quantum dynamics of a two-level system (WC) regime, the opposite occurs, i.e., the interaction constant (TLS) interacting with a quantized single-mode of the electro- is smaller than the dissipation rate of the system (g < κ). For magnetic field in a dissipative cavity. The system-reservoir in- the description of any of these regimes, we assume that the atom teraction occurs by means of the leakage of photons through can be in its ground |gi or excited |ei state, and the cavity pho- the cavity mirrors and the continuous and incoherent pumping tonic field can have zero |0ci or one photon |1ci. Therefore, the of the TLS, a situation common in semiconductor cavity quan- atom-cavity system is described by the bare states, |0i = |g, 0ci, tum electrodynamics [58]. In principle, this problem could be |1i = |e, 0ci, |2i = |g, 1ci. The system of equations that de- solved formally using the Glauber-Sudarshan P-representation scribes the quantum dynamics in resonance (ωc = ωa) is the of the corresponding TLME [59], though, this task might be following, challenging and not free of approximations. However, by per- forming a MSA we are able to obtain simple approximate an- ρ˙0,0 = − P ρ0,0 + κ ρ2,2, (70a)  alytical expressions for the strong and weak coupling regimes ρ˙1,1 = − ig ρ2,1 − ρ1,2 + P ρ0,0, (70b) that allow us to study the competition between pumping and  ρ˙2,2 = − ig ρ1,2 − ρ2,1 − κ ρ2,2, (70c) losses on the state of the physical system.  κ The JC hamiltonian has two parts, the first one contains the ρ˙1,2 = − ig ρ2,2 − ρ1,1 − ρ1,2; (70d) 2 information of the energy of the system, ∗ where, ρ j,i = ρi, j, andρ ˙0,0 = −(˙ρ1,1 + ρ˙2,2) is a consequence of † ωa H0 = ωc a a + σz, (65) the trace-preservation. We will consider that initially the atom 2 is in its excited state and there are no photons in the cavity field, where, a and a† are the annihilation and creation photonic op- i.e., ρ1,1(0) = 1 and all other entries of the density matrix will t = erators, σz is the third Pauli matrix, and ωc and ωa are respec- be zero when 0. tively the energies of the photonic mode (cavity) and the TLS (atom). The second part, has the information of the interac- 4.1. Strong coupling tion between the involved subsystems, which is characterized In the SC regime the perturbation will be associated with the by the radiation-matter interaction constant g. Therefore, near Lindblad terms for incoherent pumping and leakage of photons 8 (a) (b)

1.0 0.3 (0,0) %0,0 (t) (0,1) %0,0 (t) 0.2 (0,1) (1,0) %0,0 (t) + %0,0 (t) 0.5 Num.

0.1 Populations Ground state

0.0 0.0 0 1 2 3 4 0 2 4 6 g t g t

Figure 2: Comparison between multiple-scale analysis and numerical integration solutions for the strong coupling regime. (a) Population of the states |0i (green with pentagon markers), |1i (red with square markers), |2i (blue with circle markers); the markers show the numerical solution, while the continuous lines represent the approximate solution to first-order in the multiple-scale perturbation technique. (b) Convergence of the approximations to the numerical solution for the population of the ground state |0i; as before, the markers represent the numerical solution. For both plots, the parameters used were, ωc = ωa = 1000 g, κ = 0.1 g, P = 0.01 g.

(a) (b) 0.01 0.5

0.0 0.00

%(0,1)(t) + %(1,0)(t) 0.5 Num. 1,2 1,2 − (0,0) (0,1) (0,1) (0,2) (1,1) %1,2 (t) %1,2 (t) %1,2 (t) %1,2 (t) + %1,2 (t)

Imaginary part of 0.01 − 0 5 10 100 105 110 g t g t

Figure 3: Imaginary part of coherence in the strong coupling regime for, (a) short times, and, (b) long times. In both plots the circle markers represent the numerical solution to the system of ordinary differential equations. The parameters used were, ωc = ωa = 1000 g, κ = 0.1 g, P = 0.01 g. through the cavity mirrors, then,     L0 = − i H, ρ , (71) 0 0 0  ,   %(0 0) t  2 gt i gt  . SC ( ) = 0 cos ( ) 2 sin (2 ) (75) 0 0  i  L1 = κ D[a] ρ + P D[σ+] ρ, (72)  − 2  0 2 sin (2gt) sin (gt) where, κ = ακ0 and P = αP0, being α the perturbation parame- ter. Therefore, we will solve the master equation, First-order (SC). To first-order in the MSPT, it is necessary

to calculate F1,0(τ) and then its time-average F1,0 . For this {L L } τ0 ρ˙(t) = 0 + α 1 ρ(t). (73) purpose, the aforementioned superoperator was calculated and Afterwards, we will briefly describe the solution procedure up it was possible to recognize that its terms oscillate with a fre- Ω = g to second-order in the MSPT together with the approximate an- quency 1,0 ; thus, the temporal average was calculated T = π/g alytical expressions up to first-order in the MSPT. The shown with respect to the period 1,0 2 . Using this result, we calculate %(0,1)(t) and %(1,0)(t), analytical expressions were obtained using the general proce- SC SC dure developed in Sec.3, and they were used to obtain the (1) (0,1) (1,0) ρ (t) = % (t) + % (t), (76) curves in Figs.2 and3. SC SC SC

Zero-order (SC). As mentioned above, the zero-order solution solves the unperturbed problem,

(0) (0,0) ρSC (t) = %SC (t), (74) 9 (a) (b)

0.04 (0,2) (1,1) (2,0) %0,0 (t) + %0,0 (t) + %0,0 (t) %(0,2)(t) + %(1,1)(t) 0,0 0,0 0.020 Num.

0.03

(0,2) (1,1) (2,0) Ground state 0.019 %1,2 (t) + %1,2 (t) + %1,2 (t) (0,2) (1,1) %1,2 (t) + %1,2 (t) Num.

0.02 Imaginary part of coherence

0 200 400 0 200 400 κ t κ t Figure 4: Results of the multiple-scale perturbation technique in the weak coupling regime. (a) Population of the ground state. (b) Imaginary part of coherence. In both plots, the circle markers show the numerical solution. The parameters used were, ωc = ωa = 1000 κ, g = P = 0.01 κ.

equations in Eqs. (70), using the aforementioned parameters. κ   Fig. 2a presents the population in the transitory regime for the %(0,1) t = − −(2P+κ)t/2 , SC , ( ) 1 e (77a) three states considered, it shows that the first-order solution in 0 0 2P + κ the MSPT is in perfect agreement with the numerical solution (0,1) 1 n −(2P+κ)t/2 %SC (t) = 2P + κ e obtained. Besides, Fig. 3b shows how the MSPT approaches 1,1 2(2P + κ) the numerical solution for the population of the ground state. −κt/2 o +(2P + κ) e cos (2gt) , (77b) On the other hand, Fig.3 shows the imaginary part of coher- i ence between states |1i and |2i. In particular, Fig. 3a shows the %(0,1)(t) = e−κt/2 sin (2gt); (77c) SC1,2 2 perfect agreement of the first-order solvability condition solu- tion with the numerical one for very short times. In addition, Fig. 3b shows that for long times this agreement is not fulfilled (1,0) κ sin(2gt) − until the second-order solvability condition. % (t) = − e κt/2, (78a) SC0,0 4g κ sin (2gt) n   4.2. Weak coupling %(1,0)(t) = (2P + κ) 1 + e−Pt e−κt/2 SC1,1 8g(2P + κ) Although it is natural to think that the perturbative term will  o +4P 1 − e−(2P+κ)t/2 , (78b) always be associated with Lindblad dissipators, in some cases it is useful to include hamiltonian terms in the perturbation. Such iκ sin2(gt) n %(1,0)(t) = (2P + κ) e−(2P+κ)t/2 is the case of the WC regime, where the perturbation will be SC1,2 4g(2P + κ) associated with the hamiltonian interaction term and with the  o +4P 1 − e−(2P+κ)t/2 . (78c) Lindblad term for continuous and incoherent pumping,   Up to first-order order in the perturbative expansion, the contri- L0 = − i H0, ρ + κ D[a] ρ, (79) bution to the frequency from κ and P appears simultaneously, 0  0 L = − i g V, ρ + P D[σ ] ρ, (80) providing the characteristic time scales, 2/κ and 2/(2P + κ). On 1 + the other hand, the approximate solution for the populations, where, g = αg0 and P = αP0 being α the perturbation pa- up to this order is in very good agreement with the numerical rameter. The plots of the approximate solutions for the WC solution, as is shown in Fig.2. regime are shown in Fig.4. The parameters used in terms of the dissipative rate κ were, ω = ω = 1000 κ, g = P = 0.01 κ. Second-order (SC). To this order the time-averages were made c a Next, we briefly describe the solution procedure together with with respect to the periods, T = 2π/g and T = 2πi/P0. 2,0 1,1 the approximate analytical expressions up to second-order in Even though we do not show the approximate analytic expres- the MSPT. As before, the shown analytical expressions were sions up to this order, it is important to note that a further im- obtained using the general procedure developed in Sec.3, and provement of the coherences was made through the second- they were used to obtain the curves in Fig.4. order solvability condition, as plotted in Fig. 3b. In Fig.2 and Fig.3, we compare approximate and numer- Zero-order (WC). Unlike the previous case, the unperturbed ical solutions for the SC regime. The parameters used in units solution does not show time-evolution of the system, of the coupling constant g were, ωc = ωa = 1000 g, κ = 0.1 g, P = 0.01 g. The numerical solution was obtained by means of (0) (0,0) ρWC(t) = %WC (t), (81) the numerical integration of the system of ordinary differential 10   0 0 0 ,   %(0 0)(t) = 0 1 0 . (82) WC   0 0 0 0.004 First-order (WC). In this case the time-average of F1,0(τ0) is performed with respect to an imaginary period T1,0 = 4πi/κ in order to separate the zero-frequency terms. To first-order in the (ρnum, ρ(s1)) T SC SC MSPT, 0.002 (ρnum, ρ(1)) (1) (0,1) (1,0) T SC SC ρ (t) = % (t) + % (t), (83) (ρnum, ρ(s2)) WC WC WC T SC SC Trace distance SC   0 0 0 (0,1)   %WC (t) = 0 1 0 , (84) 0.000   0 50 100 150 0 0 0 g t Figure 5: Trace distances between the numerical solution and different orders   of the multiple-scale perturbation technique for the strong coupling regime. 0 0 0  The three quantities plotted oscillate with the frequency of the Rabi oscilla- ,  2ig  −κt/  %(1 0) t 0 0 1 − e 2  . tions; though, only the oscillations for T (ρnum, ρ(s2)) are shown, whilst for WC ( ) =  κ  (85) SC SC  2ig  −   T (ρnum, ρ(s1)) and T (ρnum, ρ(1) ) the curves follow the local maxima. The pa- 0 − 1 − e κt/2 0  SC SC SC SC κ rameters used here are the same as those used to obtain Figs.2 and3, ωc = ωa = 1000 g, κ = 0.1 g, P = 0.01 g. Second-order (WC). To second-order the time-averages were calculated with respect to the periods, T2,0 = 4πi/κ and T1,1 = 2πi/P0. The second-order approximate solution is, 0.002 (ρnum, ρ(s2)) (2) (0,2) (1,1) (2,0) T WC WC ρ (t) = % (t) + % (t) + % (t), (86) (ρnum, ρ(2)) WC WC WC WC T WC WC

  0 0 0 ,   %(0 2)(t) = 0 1 0 ; (87) 0.001 WC   0 0 0 Trace distance WC 4g2   %(1,1) (t) = 1 − e−Pt , (88a) WC0,0 0.000 Pκ 0 400 800 4g2   %(1,1) t = − − −Pt , κ t WC1,1 ( ) 1 e (88b) Pκ Figure 6: Trace distances between the numerical integration solution and dif- 2ig   %(1,1) (t) = 1 − e−κt/2 ; (88c) ferent orders of the multiple-scale analysis in the weak coupling regime. The WC1,2 κ parameters used here are the same as those used to obtain Figs.4, ωc = ωa = 1000 κ, g = P = 0.01 κ.

4g2   %(2,0) (t) = − 3 − 4 e−κt/2 + e−κt , (89a) 4.3. Error estimation WC0,0 κ2 In the quantum dynamics shown above, the truncation cri- g2   (2,0) 8 −κt/2 terion was based on the qualitative comparison of the results %WC (t) = 1 − e , (89b) 1,1 κ2 obtained for each order of the perturbative expansion with the 8ig3   %(2,0) (t) = − 1 − e−Pt − e−κt/2 + e−(2P+κ)t/2 . (89c) exact numerical dynamics. The asymptotic series were trun- WC1,2 Pκ2 cated when a good correspondence was found for short and long times according to the natural time scales of each prob- From comparison between Figs.2 and3, and Fig.4; we lem. Thus, the regimes of applicability of the method were not can see that the results for the SC regime are in better agree- clear, since the errors introduced by the truncation process were ment with the numerical calculations than those for the WC not understood quantitatively. Although, it is evident that ignor- regime. This is because in the latter case we have the pertur- ing higher order contributions will introduce significant errors bation of a purely damped system, which makes it difficult to in the approximation, especially, for long times, and perturba- introduce new time-scales that actually change the system dy- tions αL comparable with the unperturbed liouvillian of the namics. Then, a better approximation for the WC regime would 1 system L . However, attempting to find an analytical estima- require going beyond in the perturbation expansion. 0 tion of errors essentially requires calculating the exact general evolution of the system density matrix in all regimes. Neverthe- less, using a combination of numerical and phenomenological approaches, here we provide a simple error estimation for the 11 strong and weak coupling regimes, based on the results shown On the other hand, Fig.6 shows the following trace dis- in Figs.2–4, thus quantifying the reliability of the found ap- tances for the WC regime, proximate analytical expressions.     T ρnum t , ρ(s2) t , T ρnum t , ρ(2) t To estimate the error and, therefore, arriving to more robust WC ( ) WC ( ) WC ( ) WC( ) ; conclusions about accuracy and convergence to each order in where, the perturbative expansion, we use the trace distance between (s2) (0,2) (1,1) the quantum dynamics calculated numerically and the different ρWC (t) = %WC (t) + %WC (t), orders in the perturbation theory for each regime. The trace (2) (0,2) (1,1) (2,0) ρWC(t) = %WC (t) + %WC (t) + %WC (t). distance between two density matrices ρ1 and ρ2, is defined as, In this case, the trace distances for zero-order, first solvability ( q ) X 1 2 1 condition, and first-order, are exactly the same and asymptot- T (ρ1, ρ2) = Tr (ρ1 − ρ2) = |λi|, (90) 2 2 i ically approximately equal to 0.04. In this case the dynamics is much more regular and, as before, from a phenomenological where the λi are the eigenvalues of the operator (ρ1 − ρ2). description we can conclude that trace distances below 0.001 Figs.5 and6 show the trace distance for the strong and correspond to very good perturbative approximations. weak coupling regimes respectively, where we used the same sets of parameters as those in Figs.2–4. In Fig.5, we study the time-evolution (extended towards long times) of the trace 5. Conclusions distance between the numerically calculated quantum dynamics To conclude, we have presented a multiple-scale perturba- and three different solutions in the MSPT. Hence, we plotted, tion technique that allows us to find excellent approximate so-       lutions to time-local master equations describing open quan- T ρnum t , ρ(s1) t , T ρnum t , ρ(1) t , T ρnum t , ρ(s2) t SC ( ) SC ( ) SC ( ) SC ( ) SC ( ) SC ( ) ; tum systems. The technique provides the time-evolution of the corresponding dynamical map and, consequently, the time- with, evolution of the system density matrix for arbitrary initial con- ρ(s1)(t) = %(0,1)(t), ditions, allowing us to identify in each order the characteristic SC SC time scales involved in the problem. The presented method is (1) (0,1) (1,0) ρSC (t) = %SC (t) + %SC (t), easy to implement using the general expressions derived in this (s2) (0,2) (1,1) paper, and is proper for a broad range of perturbations and open ρSC (t) = %SC (t) + %SC (t); quantum systems. Moreover, as is shown in the illustrative ex- where, as schematized in Fig.1, the solvability condition so- ample, the method encourages separating the problem accord- lutions are intermediate orders in the multiple-scale perturba- ing to the physical regimes that are being described, facilitating tive expansion. The three aforementioned trace distances oscil- the study of open quantum system from physical criteria. Be- late with the frequency of the Rabi oscillations shown in Fig. sides, if the solvable liouvillian and the perturbation commute, num (s2) 2a. However, only the oscillations of T (ρSC , ρSC ) are shown, the presented technique reaches the exact solution in a single num (s1) num (1) whilst for T (ρSC , ρSC ) and T (ρSC , ρSC ) the curves follow the step in the perturbation theory. The description in terms of the local maxima, since otherwise, the high frequency oscillations dynamical map is suitable to obtain analytical expressions for would make the data very difficult to visualize. physical observables, such as the energy density spectrum, and It is noteworthy that while the trace distance for short times high-order correlation functions; in a similar way, it is appropri- decreases from the first solvability condition to the first-order ate to calculate quantifiers and witnesses of entanglement, such solution; for long times, it increases, and needs the second- as concurrence and negativity. order solvability condition to decrease again. This behavior was Finally, it is worth noting that the developed technique is evident in Figs.3, where for short times the imaginary part of a useful tool for the study of the dynamics of time-dependent (0,1) quantum systems, both open and hamiltonian. This, as conse- the coherence %1,2 (t) already corresponded with the exact dy- namics; while for long times it differed by a constant factor, quence of the time-averaging procedure in the multiple-scale however, the first-order solution added imprecision that had to analysis, which allows us to deal with time-dependent param- be corrected by the second-order solvability condition. Besides, eters in a surprisingly simple way and, therefore, to find general although not shown in Fig.5, the second-order solution reduces analytical expressions for time-dependent quantum systems that both the error for short and long times; though, the reduction is would otherwise be very difficult to calculate. In particular, the more significant for short times. This irregular behavior of the method is appropriate for systems from nonstationary cavity error for short and long times to each order in the perturba- quantum electrodynamics, such as those related with the dy- tive expansion, makes a quantitative estimation very difficult; namical Casimir effect. nevertheless, as the trace distance describes very well what is observed for the quantum dynamics, we can conclude from a Acknowledgments phenomenological analysis that trace distances below 0.001 ev- idence excellent agreement of the approximation with the exact D.N.B.-G. and H.V.-P. acknowledge partial financial sup- numerical results. port from COLCIENCIAS under the projects, “Emisión en sis- temas de qubits superconductores acoplados a la radiación”, 12 0 00 code 110171249692, CT 293-2016, HERMES 31361; “Con- with τ0 > τ0 . We continue in this way infinitely until the inte- trol dinámico de la emisión en sistemas de qubits acoplados gration interval is so small that the integral is null and we get, con cavidades no-estacionarias”, HERMES 41611; and “Elec- X 1 trodinámica cuántica de cavidades no estacionarias”, HERMES Φ(0)(t) = (L τ )n C (t ) (A.5) n 0 0 0,1 1 43351. D.N.B.-G. acknowledges funding from COLCIENCIAS n ! through “Convocatoria No. 727 de 2015 – Doctorados Na- (0) L0τ0 cionales 2015”. B.A.R. acknowledges financial support from Φ (t) = e C0,1(t1), (A.6) Universidad de Antioquia under Project No. 2014-989 (CODI- where there is an implicit time ordering associated with the fact UdeA). B.A.R. and D.N.B.-G. gratefully acknowledge support that τ0 > τ1 > ... > τn, which is guaranteed by the perturbative from “Proyecto de sostenibilidad del Grupo de Física Atómica condition 0 < α < 1. Finally, y Molecular”. n o (0) L0τ0 % (t) = e C0,1(t1) ρ(0). (A.7) Appendix A. Solution to the zero-order equation Appendix B. Commutation relations between the Kn su- It is worth noting, that given the nature of the mathemati- peroperators cal objects involved in the zero-order equation, Eq. (20a), the integration of the differential equation must be done in such a Up to first-order in the MSPT, we have from Eq. (33) the way that the temporal order is respected. Therefore, we follow definition of the superoperator F1,0(τ0), a procedure similar to the one used for the construction of the −K τ K τ τ = 0 0 L τ 0 0 , K = . Dyson series. From Eq. (20a), we have, F1,0( 0) e 1( 0) e 1 F1,0 τ0 (B.1)

(0) (0) D0 % (t) − L0 % (t) = 0. (A.1) It is clear that the equation of motion for F1,0(τ0) is given by,

(0) (0)   Being % (t) = Φ (t) ρ(0), we have then for the zero-order D0 F1,0(τ0) = F1,0(τ0), K0 , F1,0(0) = L1(0). (B.2) dynamical map, Taking the time-average with respect to τ0, we have, (0) (0) D0 Φ (t) − L0 Φ (t) = 0. (A.2) h i D = , K = K , K , 0 F1,0 τ0 F1,0 τ0 0 [ 1 0 ] (B.3) Integrating with respect to τ0,

Z T Z τ0 (0) 0 1,0 ∂Φ (τ0, τ1,... ) 0 1 D0 F1,0 = D0 F1,0(τ0) dτ0 0 dτ0 = τ0 T1,0 0 ∂τ0 0 Z τ0 1   L Φ(0)(τ0 , τ ,... ) dτ0 = F1,0(T1,0) − F1,0(0) = 0, (B.4) 0 0 1 0 T1,0 0

(0) (0) 0 since F1,0 is periodic with period T1,0, and therefore, Φ (t) − Φ (τ0, τ1,... ) = 0 τ0=0 Z τ0 [K1 , K0 ] = 0. (B.5) (0) 0 0 L0 Φ (τ0, τ1,... ) dτ0 , 0 Similarly, up to the (n + 1)th-order in the MSPT (n ≥ 1), (0) 0 Φ τ , τ ,... = t −Knτn Knτn where we label ( 0 1 ) τ0 =0 C0,1( 1). The indexing τ = − , 0 F1,n( n) e F2,n 1 τn−1 e of the C0,1(t1) superoperator is such that the first index is asso- Kn+ = F ,n ; (B.6) ciated with the term of the expansion that is being determined, 1 1 τn Φ(0) in this case; and the second index is associated with the   minimum-labeled time-scale on which the superoperator de- D τ = τ , K , = − . n F1,n( n) F1,n( n) n F1,n(0) F2,n 1 τn−1 (B.7) pends, here τ1. Hence, we have Analogously to the above, we take the time-average in a period Z τ0 0 0 (0) (0) T1,n of the superoperator F1,n, Φ (t) = C0,1(t1) + L0 Φ (τ0, τ1,... ) dτ0 , (A.3) 0 h i Dn F ,n = F ,n , Kn , Dn F ,n = 0; (B.8) then, we can solve this iteratively by replacing the integral form 1 τn 1 τn 1 τn (0) 0 of Φ (τ0, τ1,... ) in Eq. (A.3), and therefore, (0) L Φ (t) = C0,1(t1) + 0 τ0 C0,1(T1) [Kn+1 , Kn ] = 0, n ≥ 0. (B.9) 0 Z τ0 Z τ0 (0) 00 00 0 + L0 Φ (τ0 , τ1,... ) dτ0 dτ0 , (A.4) This shows that the zeroth term in the asymptotic expansion of 0 0 the dynamical map, Φ(0)(t) = eK0τ0 eK1τ1 ... eKnτn ... , is an exponential factorization of commuting superoperators. 13 References [24] W. Scherer, Quantum averaging. I. Poincare-von Zeipel is Rayleigh Schrodinger, J. Phys. A: Math. Gen. 27 (24) (1994) 8231–46. doi: [1] A. Kavokin, J. J. Baumberg, G. Malpuech, F. P. Laussy, Microcavities, 10.1088/0305-4470/27/24/028. Oxford University Press, Oxford, 2007. doi:10.1093/acprof:oso/ [25] G. Sandri, A new method of expansion in mathematical physics - I, Nuovo 9780199228942.001.0001. Cim. 36 (1) (1965) 67. doi:10.1007/BF02750660. [2] Á. Rivas, S. F. Huelga, Open Quantum Systems, SpringerBriefs in [26] R. Ramnath, G. Sandri, A generalized multiple scales approach to a class Physics, Springer-Verlag Berlin Heidelberg, Berlin, Heidelberg, 2012. of linear differential equations, J. Math. Anal. Appl. 28 (2) (1969) 339– doi:10.1007/978-3-642-23354-8. 64. doi:10.1016/0022-247X(69)90034-1. [3] D. Walls, G. J. Milburn, Quantum Optics, Springer-Verlag Berlin Heidel- [27] J. K. Kevorkian, J. D. Cole, Multiple Scale and Singular Perturbation berg, 2008. doi:10.1007/978-3-540-28574-8. Methods, Vol. 114 of Applied Mathematical Sciences, Springer-Verlag [4] M. A. Nielsen, I. L. Chuang, Quantum Computation and Quantum Infor- New York, 1996. doi:10.1007/978-1-4612-3968-0. mation, Cambridge University Press, Cambridge, 2010. doi:10.1017/ [28] A. H. Nayfeh, Perturbation Methods, Wiley-VCH Verlag GmbH, Wein- CBO9780511976667. heim, Germany, 2004. doi:10.1002/9783527617609. [5] H. M. Wiseman, G. J. Milburn, Quantum Measurement and Con- [29] S. H. Strogatz, Nonlinear dynamics and chaos: with applications to trol, Cambridge University Press, Cambridge, 2009. doi:10.1017/ physics, biology, chemistry, and engineering, Westview Press, Boulder, CBO9780511813948. CO, 2014. [6] N. Quesada, H. Vinck-Posada, B. A. Rodríguez, Density operator of a [30] C. M. Bender, L. M. A. Bettencourt, Multiple-Scale Analysis of the system pumped with polaritons: a Jaynes–Cummings-like approach, J. Quantum Anharmonic Oscillator., Phys. Rev. Lett. 77 (20) (1996) 4114. Phys. Condens. Matter 23 (2011) 025301. doi:10.1088/0953-8984/ doi:10.1103/PhysRevLett.77.4114. 23/2/025301. [31] C. M. Bender, L. M. A. Bettencourt, Multiple-scale analysis of quantum [7] H. J. Carmichael, Statistical Methods in Quantum Optics 2, Theoreti- systems, Phys. Rev. D 54 (12) (1996) 7710. doi:10.1103/PhysRevD. cal and Mathematical Physics, Springer-Verlag Berlin Heidelberg, 2008. 54.7710. doi:10.1007/978-3-540-71320-3. [32] G. Auberson, M. Capdequi Peyranère, Quantum anharmonic oscillator in [8] W. Magnus, On the exponential solution of differential equations for a the Heisenberg picture and multiple scale techniques, Phys. Rev. A 65 linear operator, Commun. Pure Appl. Math. 7 (4) (1954) 649–73. doi: (2002) 032120. doi:10.1103/PhysRevA.65.032120. 10.1002/cpa.3160070404. [33] M. W. Janowicz, J. M. a. Ashbourn, Dynamics of the four-level Λ system [9] X. X. Yi, C. Li, J. C. Su, Perturbative expansion for the master equation in a two-mode cavity, Phys. Rev. A 55 (3) (1997) 2348. doi:10.1103/ and its applications, Phys. Rev. A 62 (2000) 013819. doi:10.1103/ PhysRevA.55.2348. PhysRevA.62.013819. [34] M. Janowicz, Suppression of the Rabi oscillations in a cavity partially [10] C. H. Fleming, N. I. Cummings, Accuracy of perturbative master equa- filled with a dielectric having a time-dependent refractive index, Phys. tions, Phys. Rev. E 83 (2011) 031117. doi:10.1103/PhysRevE.83. Rev. A 57 (6) (1998) 5016. doi:10.1103/PhysRevA.57.5016. 031117. [35] M. Janowicz, Evolution of wave fields and atom-field interactions in a [11] F. B. Anders, A. Schiller, Real-time dynamics in quantum-impurity cavity with one oscillating mirror, Phys. Rev. A 57 (6) (1998) 4784. doi: systems: A time-dependent numerical renormalization-group approach, 10.1103/PhysRevA.57.4784. Phys. Rev. Lett. 95 (2005) 196801. doi:10.1103/PhysRevLett.95. [36] D. A. R. Dalvit, F. D. Mazzitelli, Creation of photons in an oscillating 196801. cavity with two moving mirrors, Phys. Rev. A 59 (4) (1999) 3049. doi: [12] R. Bulla, T. A. Costi, T. Pruschke, Numerical renormalization group 10.1103/PhysRevA.59.3049. method for quantum impurity systems, Rev. Mod. Phys. 80 (2008) 395– [37] M. Crocce, D. A. R. Dalvit, F. D. Mazzitelli, Resonant photon creation in 450. doi:10.1103/RevModPhys.80.395. a three-dimensional oscillating cavity, Phys. Rev. A 64 (1) (2001) 013808. [13] M. J. Hartmann, J. Prior, S. R. Clark, M. B. Plenio, Density matrix renor- doi:10.1103/PhysRevA.64.013808. malization group in the heisenberg picture, Phys. Rev. Lett. 102 (2009) [38] M. Crocce, D. A. R. Dalvit, F. D. Mazzitelli, Quantum electromagnetic 057202. doi:10.1103/PhysRevLett.102.057202. field in a three-dimensional oscillating cavity, Phys. Rev. A 66 (3) (2002) [14] F. Reiter, A. S. Sørensen, Effective operator formalism for open quantum 033811. doi:10.1103/PhysRevA.66.033811. systems, Phys. Rev. A 85 (2012) 032111. doi:10.1103/PhysRevA. [39] A. V. Dodonov, V. V. Dodonov, Approximate analytical results on the 85.032111. cavity dynamical Casimir effect in the presence of a two-level atom, Phys. [15] E. M. Kessler, Generalized schrieffer-wolff formalism for dissipative sys- Rev. A 85 (1) (2012) 015805. doi:10.1103/PhysRevA.85.015805. tems, Phys. Rev. A 86 (2012) 012126. doi:10.1103/PhysRevA.86. [40] P. B. Kahn, Y. Zarmi, Consistent Application of the Method of Multiple- 012126. Time Scales to Nonlinear Systems, in: C. W. Wang (Ed.), Nonlinear Phe- [16] D. Chrusci´ nski,´ A. Kossakowski, Feshbach projection formalism for open nomena Research Perspectives, Nova Science Publishers, Inc, New York, quantum systems, Phys. Rev. Lett. 111 (2013) 050402. doi:10.1103/ 2007, Ch. 4, pp. 103–30. PhysRevLett.111.050402. [41] K. Kraus, States, Effects, and Operations: Fundamental Notions of Quan- [17] J. Cui, J. I. Cirac, M. C. Bañuls, Variational matrix product operators tum Theory, Vol. 190 of Lecture Notes in Physics, Springer-Verlag Berlin for the steady state of dissipative quantum systems, Phys. Rev. Lett. 114 Heidelberg, Berlin, Heidelberg, 1983. doi:10.1007/3-540-12732-1. (2015) 220601. doi:10.1103/PhysRevLett.114.220601. [42] G. Schaller, Open Quantum Systems Far from Equilibrium, Vol. 881 of [18] A. Kamenev, Field Theory of Non-Equilibrium Systems, Cambridge Uni- Lecture Notes in Physics, Springer International Publishing, 2014. doi: versity Press, Cambridge, 2011. doi:10.1017/CBO9781139003667. 10.1007/978-3-319-03877-3. [19] L. M. Sieberer, M. Buchhold, S. Diehl, Keldysh field theory for driven [43] D. Bruß, G. Leuchs (Eds.), Lectures on Quantum Information, Wiley- open quantum systems, Rep. Prog. Phys. 79 (9) (2016) 096001. doi: VCH Verlag GmbH, Weinheim, Germany, 2006. doi:10.1002/ 10.1088/0034-4885/79/9/096001. 9783527618637. [20] A. C. Y. Li, F. Petruccione, J. Koch, Perturbative approach to Markovian [44] H.-P. Breuer, E.-M. Laine, J. Piilo, B. Vacchini, Colloquium : Non- open quantum systems, Sci. Rep. 4 (1) (2015) 4887. doi:10.1038/ Markovian dynamics in open quantum systems, Rev. Mod. Phys. 88 (2) srep04887. (2016) 021002. doi:10.1103/RevModPhys.88.021002. [21] A. C. Y. Li, F. Petruccione, J. Koch, Resummation for nonequilibrium [45] K. Kraus, General state changes in quantum theory, Ann. Phys. 64 (2) perturbation theory and application to open quantum lattices, Phys. Rev. (1971) 311–35. doi:10.1016/0003-4916(71)90108-4. X 6 (2016) 021037. doi:10.1103/PhysRevX.6.021037. [46] M. D. Choi, Completely positive linear maps on complex matrices, Linear [22] C. M. Bender, S. A. Orszag, Advanced Mathematical Methods Algebra Appl. 10 (3) (1975) 285–90. doi:10.1016/0024-3795(75) for Scientists and Engineers I: Asymptotic Methods and Perturba- 90075-0. tion Theory, Springer-Verlag New York, 1999. doi:10.1007/ [47] V. Gorini, A. Kossakowski, E. C. G. Sudarshan, Completely positive dy- 978-1-4757-3069-2. namical semigroups of N-level systems, J. Math. Phys. 17 (5) (1976) 821. [23] M. Janowicz, Method of multiple scales in quantum optics, Phys. Rep. doi:10.1063/1.522979. 375 (5) (2003) 327–410. doi:10.1016/S0370-1573(02)00551-3. [48] H.-P. Breuer, F. Petruccione, The Theory of Open Quantum Systems,

14 Oxford University Press, Oxford, 2007. doi:10.1093/acprof:oso/ 9780199213900.001.0001. [49] E. Andersson, J. D. Cresser, M. J. W. Hall, Finding the Kraus decompo- sition from a master equation and vice versa, J. Mod. Opt. 54 (12) (2007) 1695–716. doi:10.1080/09500340701352581. [50] Á. Rivas, S. F. Huelga, M. B. Plenio, Entanglement and non-Markovianity of quantum evolutions, Phys. Rev. Lett. 105 (5) (2010) 1–4. doi:10. 1103/PhysRevLett.105.050403. [51] Á. Rivas, S. F. Huelga, M. B. Plenio, Quantum non-Markovianity: Char- acterization, quantification and detection, Rep. Prog. Phys. 77 (9). doi: 10.1088/0034-4885/77/9/094001. [52] R. Alicki, K. Lendi, Quantum Dynamical Semigroups and Applications, Vol. 717 of Lecture Notes in Physics, Springer-Verlag Berlin Heidelberg, 2007. doi:10.1007/3-540-70861-8. [53] G. Lindblad, On the generators of quantum dynamical semigroups, Com- mun. Math. Phys. 48 (2) (1976) 119–30. doi:10.1007/BF01608499. [54] H. J. Carmichael, Statistical Methods in Quantum Optics 1, Theoreti- cal and Mathematical Physics, Springer-Verlag Berlin Heidelberg, 1999. doi:10.1007/978-3-662-03875-8. [55] M. J. W. Hall, J. D. Cresser, L. Li, E. Andersson, Canonical form of master equations and characterization of non-Markovianity, Phys. Rev. A 89 (4) (2014) 042120. doi:10.1103/PhysRevA.89.042120. [56] M. R. Hush, I. Lesanovsky, J. P. Garrahan, Generic map from non- Lindblad to Lindblad master equations, Phys. Rev. A 91 (3) (2015) 032113. doi:10.1103/PhysRevA.91.032113. [57] D. Chrusci´ nski,´ A. Kossakowski, Non-Markovian Quantum Dynamics: Local versus Nonlocal, Phys. Rev. Lett. 104 (7) (2010) 070406. doi: 10.1103/PhysRevLett.104.070406. [58] G. Khitrova, H. M. Gibbs, M. Kira, S. W. Koch, A. Scherer, Vacuum Rabi splitting in semiconductors, Nat. Phys. 2 (2) (2006) 81–90. doi: 10.1038/nphys227. [59] C. W. Gardiner, P. Zoller, Quantum noise: a handbook of Markovian and non-Markovian quantum stochastic methods with applications to quan- tum optics, Springer-Verlag Berlin Heidelberg, 2004. [60] F. P. Laussy, E. del Valle, C. Tejedor, Luminescence spectra of quantum dots in microcavities. i. bosons, Phys. Rev. B 79 (2009) 235325. doi: 10.1103/PhysRevB.79.235325. [61] C. A. Vera, N. Quesada M, H. Vinck-Posada, B. A. Rodríguez, Charac- terization of dynamical regimes and entanglement sudden death in a mi- crocavity quantum dot system, J. Phys.: Condens. Matter 21 (39) (2009) 395603. doi:10.1088/0953-8984/21/39/395603. [62] P. Lodahl, S. Mahmoodian, S. Stobbe, Interfacing single photons and single quantum dots with photonic nanostructures, Rev. Mod. Phys. 87 (2015) 347–400. doi:10.1103/RevModPhys.87.347.

15