Unified Gorini–Kossakowski–Lindblad–Sudarshan quantum beyond the secular approximation

Anton Trushechkin1, 2 1Steklov Mathematical Institute of Russian Academy of Sciences, Moscow 119991, Russia 2National University of Science and Technology MISIS, Moscow 119049, Russia∗ (Dated: July 27, 2021) Derivation of a quantum master equation for a system weakly coupled to a bath which takes into account nonsecular effects, but nevertheless has the mathematically correct Gorini–Kossakowski– Lindblad–Sudarshan form (in particular, it preserves positivity of the density operator) and also satisfies the standard thermodynamic properties is a known long-standing problem in theory of open quantum systems. The nonsecular terms are important when some energy levels of the system or their differences (Bohr frequencies) are nearly degenerate. We provide a fully rigorous derivation of such equation based on a formalization of the weak-coupling limit for the general case.

I. INTRODUCTION particular, does not preserve positivity of the density op- erator, thus leading to unphysical predictions. Also, this Quantum master equations are at the heart of theory of equation does not have the mentioned thermodynamic open quantum systems [1–3]. They describe the dynam- properties. ics of the reduced density operator of a system interact- Derivation of a mathematically correct quantum mas- ing with the environment (“bath”) and are widely used ter equation which takes into account nonsecular effects in quantum optics, condensed matter physics, charge and is actively studied. Several heuristic approaches turning energy transfer in molecular systems, bio-chemical pro- the Redfield equation into an equation of the GKLS form cesses [4], quantum thermodynamics [5,6], etc. The without the (full) secular approximation have been pro- Redfield and Davies quantum master equations are well- posed. One possibility is the time coarse graining [22–27]. known microscopically derived equations for a system Another method is a partial secular approximation fol- weakly coupled to a bath and are crucial for understand- lowed by the approximation of slow variation of the spec- ing many physical phenomena. tral density [28–31]. Also, in certain cases, the so-called The Davies quantum master equation [7,8] is derived local approach is considered as an alternative to the sec- in a mathematically rigorous way and has the Gorini– ular approximation and is largely debated [32–40]. But, Kossakowski–Lindblad–Sudarshan (GKLS) form [9–11], like the Redfield equation, these equations do not satisfy which guarantees that the corresponding dynamics of all the mentioned thermodynamic properties. Moreover, the reduced density operator is well defined. The Davies it has been explicitly shown that the local master equa- equation also satisfies a number of properties important tion violates the second law of thermodynamics [41]. for thermodynamics: stationarity of the Gibbs state, the detailed balance condition, a covariance law (related to In this paper, we derive a unified master equation for the first law of thermodynamics [12–14]), non-negativity the weak-coupling regime in a mathematically rigorous of the entropy production (i.e., the second law of ther- and systematic way, which leads to the GKLS form and modynamics) [1,5,6, 15–17]. all the desired thermodynamic properties. The derivation However, this equation assumes that all distinct Bohr is based on a rigorous formalization of the weak-coupling frequencies (differences between the energy levels) of the limit for the general (nonsecular) case. The unified equa- system are well separated from each other. In other tion has a simple and intuitive structure similar to that words, all differences between distinct Bohr frequencies of the Davies equation. are much higher than the dissipation rates (the secular arXiv:2103.12042v3 [quant-ph] 25 Jul 2021 approximation). This assumption is not satisfied if some Interestingly, this equation coincides with the refined energy levels or Bohr frequencies are nearly degenerate (thermodynamically consistent) form of the local mas- (but not exactly degenerate), which is often the case in ter equation when the local approach is expected to be physical systems. valid. Thus, we rigorously justify the correct form of the The Redfield master equation [18] does not adopt the local master equation (popular due to its simplicity) and secular approximation and thus is more general. It is complete the results of Ref. [42]. widely used in various physical applications, the role of Note that the general idea was also proposed in Ref. [1] the nonsecular terms is studied, e.g., in Refs. [19–21]. and master equations for a particular system were rig- But, unfortunately, it is not of the GKLS form and, in orously derived in Ref. [43]. General properties of the dynamics of particular models of open quantum systems with nearly degenerate spectrum were also rigorously es- tablished in Refs. [44, 45]. Here we derive a master equa- ∗ [email protected] tion for the general situation. 2

II. REDFIELD EQUATION AND SECULAR are bath correlation functions. APPROXIMATION The matrix γαβ(ω, ω0) [with two double indices i = (α, ω0) and j = (β, ω)] is, in general, not positive semidef- Consider the system-bath Hamiltonian inite, so, the equation is not of the GKLS form and de- fines the dynamics which violates positivity. i(ω0 ω)τ/λ2 H = HS + HB + λHI , (1) As λ 0, the exponents e − for ω = ω0 rapidly oscillate→ and the corresponding terms can be6 ne- where HS is the isolated system Hamiltonian, HB is the glected (the secular approximation). After this, the equa- isolated bath Hamiltonian, H = A B is the I α α ⊗ α tion becomes of the first standard GKLS form [1,2] since, interaction Hamiltonian (with Aα being system opera- P for each ω, the matrix γαβ(ω, ω) γαβ(ω) (with the in- tors and Bα being bath operators), and λ is a formal dices α and β) is well known to be≡ positive semidefinite. small dimensionless parameter. Let HS have a purely We will refer to this equation as the secular or the Davies discrete spectrum: HS = j εjPj, where εj are the dis- master equation. tinct eigenvalues and Pj are the corresponding eigenpro- P However, as pointed out above, the secular approxima- jectors. The differences εj0 εj are called the Bohr fre- tion is not always valid. If so, the formal mathematical quencies. Denote the set of− all (positive, negative, and F limit in the present form does not correspond to physics. zero) Bohr frequencies. There is no limit in a concrete physical system, but all Let the initial system-bath state be ρ σ , where ρ 0 ⊗ B 0 physical quantities have concrete values. The limit λ 0 is the initial state of the system and σB is a reference is just a mathematical expression of the fact that the→ dis- iHB t iHB t state of the bath such that e− σBe = σB for all sipative dynamics caused by the coupling to the bath is t (e.g., a thermal state). The dynamics of the reduced much slower than all other time scales, but this is not density operator of the system is given by the case if two different Bohr frequencies ω and ω0 are iHt iHt close to each other. Obviously, nearly degenerate energy ρ(t) = Tr e− (ρ σ )e , B 0 ⊗ B levels is a particular case. where TrB is the partial trace over the bath. The density operator in the interaction picture isρ ˜(t) = iH t iH t III. UNIFIED MASTER EQUATION e S ρ(t)e− S . The derivation of the quantum master equation is based on the idea of the separation of different time A. Derivation scales. In the rigorous derivation, this intuition is formal- ized by the Bogolyubov–van Hove limit: λ 0, t , If we claim that, for a given physical system, a differ- 2 → → ∞ i(ω0 ω)τ/λ2 λ t = τ = const (see Refs. [3,7,8, 46]). The standard ence ω ω0 is small and, hence, the term e − “physical” derivation [2,3] leads to the Redfield equation is not rapidly− oscillating (with respect to the time scale (we express it in a GKLS-like form [27]): of the dissipation), then, in the formal derivation, the difference ω0 ω should be treated as infinitesimal and, d − 2 ρ˜(τ) = i[HLS(τ), ρ˜(τ)] + γαβ(ω, ω0) moreover, of the order of λ . This should be explicitly dt − ω,ω0 α,β formalized in the mathematical language of the limit. X∈F X 0 τ 1 Let us express the system Hamiltonian HS as i(ω ω) 2 e − λ A ρ˜(τ)A† 0 A† 0 A , ρ˜(τ) , (2) × βω αω − 2 αω βω   H = H(0) + λ2δH , (7) where  S S S

i(ω0 ω) τ (0) λ2 where [H , δHS] = 0 and all nearly degenerate Bohr HLS(τ) = Sαβ(ω, ω0)e − Aαω† 0 Aβω, (3) S (0) ω,ω0 α,β frequencies in H are exactly degenerate in H . In X∈F X S S other words, all distinct Bohr frequencies of H(0) are well (the subindex LS stands for the Lamb shift), S separated. Proportionality of the remaining part to λ2 mathematically expresses the fact that some oscillations Aαω = Pj0 AαPj, [HS,Aαω] = ωAαω, (4) i(ω0 ω)τ/λ2 − e − occur on the same time scale as the dissipa- j,j0 : ε ε 0 =ω Xj − j tion. (0) γ (ω, ω0) = Γ (ω) + Γ∗ (ω0), αβ αβ βα Due to commutativity and since HS may be more (0) 1 degenerate than HS, its spectral decomposition is H = S (ω, ω0) = Γ (ω) Γ∗ (ω0) S αβ 2i αβ − βα (0) (0) (0) k εk Pk , where εk are the distinct eigenvalues and ∞  (0) iωs each eigenprojector P is either one of P or a sum of Γαβ(ω) = ds e Cαβ(s) (5) P k j Z0 several Pj. (0) (0) (we assume that the last integral converges). Here Denote the set of the Bohr frequencies of HS . Then, eachF Bohr frequency ω of the original system iHB s iHB s 2 Cαβ(s) = Tr[e− Bα† e BβσB] (6) Hamiltonian HS can be expressed as ω = ω+λ δω, where 3

(0) ω and δω is a Bohr frequency of δHS. In other (i) The dissipator is constructed as if the system ∈ F (0)D words, the set of Bohr frequencies of HS is divided into F Hamiltonian was HS , with the secular approximation disjoint subsets (clusters) of the Bohr frequencies cen- (0) ω with respect to H . tered around ω (0). TheF difference between any pair S (ii) The Lamb-shift Hamiltonian H is as for the Red- of Bohr frequencies∈ F from the same cluster is proportional LS 2 field equation with the secular approximation with re- to λ . Physically, the Bohr frequencies from different (0) clusters are well separated, while those from the same spect to HS [compare Eqs. (3) and (12)]. cluster are not: If we want to describe a concrete physical system, a question about the value of λ arises. Formally, in order 2 ω0 ω = ω ω0 + λ (δω0 δω). (8) to apply the proposed analysis to a concrete physical sys- − − − (0) tem, we should express HS HS and the spectral density 2 − So, the exponent exp[i(ω0 ω)τ/λ ] is rapidly oscillating of the bath (indicating the strength of the system-bath − (as λ 0) if and only if ω and ω0 belong to different coupling) as products of a small dimensionless parameter → 2 clusters (ω = ω0). λ and energy quantities of the same order as the zeroth- Using this,6 let us drop the rapidly oscillating terms (0) 2 order system energies εk . In this case, λ is the actual from the Redfield equation (2) (i.e., apply the secular ratio of the scale of the small parameters to the scale of (0) approximation with respect to HS , which is a partial the large parameters (both are of energy dimensionality). secular approximation with respect to HS). Also, let us So, λ is defined up to an order of magnitude. perform the limit λ 0 in the arguments of γ , i.e., Moreover, λ can be treated to be incorporated into the → αβ γαβ(ω, ω0) γαβ(ω, ω) = γαβ(ω) for ω, ω0 ω. Then Hamiltonian (as it is often assumed in the physical litera- → ∈ F 2 we arrive at the following master equation: ture). Indeed, if we denote λ δHS = δHS0 and λHI = HI0 , 2 2 then λ HLS and λ in Eq. (10) can be substituted by d D H0 and 0, where the latter expressions are derived for ρ˜(τ) = i[HLS(τ), ρ˜(τ)] LS D dτ − the interaction Hamiltonian H0 . The system Hamilto- 0 I i(ω ω)τ (0) + e − γ (ω) αβ nian is then expressed as HS = HS + δHS0 . So, in prac- 0 ω (0)ω,ω ω α,β tice, for a given system, we should express its Hamilto- ∈FX X∈F X 1 nian as a sum of a reference part H(0), where all nearly A ρ˜(τ)A† 0 A† 0 A , ρ˜(τ) , (9) S × βω αω − 2 αω βω degenerate Bohr frequencies are exactly degenerate, and   a small perturbation δH , and then apply Eq. (10) (for- iHS τ iHS τ  S0 HLS(τ) = e HLSe− (HLS is given below). In the mally, with λ = 1 since λ is implicitly incorporated into Schr¨odingerpicture and in the original time scale t, the the Hamiltonian now). An explicit separation of a small equation takes the form factor λ was required only in a formal derivation. Let us summarize the physical conditions of validity of d 2 2 ρ(t) = i[HS + λ HLS, ρ(t)] + λ [ρ(t)] [ρ(t)], master equation (10): dt − D ≡ L (10) (i) The usual weak system-bath coupling condition [1– 4, 47]. Roughly, it can be expressed as Γ(ω) Ω for 1 all Bohr frequencies ω, where Ω is a characteristic| |  de- ρ = γ (ω) A ρA† A† A , ρ , D αβ βω αω − 2 αω βω cay rate of the bath correlation functions (6). For the ω (0) α,β ∈FX X    Drude–Lorentz spectral density, Ω is an explicit parame- (11) ter, see below. This condition means that the dissipative dynamics of the system is much slower than the bath HLS = Sαβ(ω, ω0)Aαω† 0 Aβω, (12) relaxation. As a consequence, the system-bath state is 0 ω (0) ω,ω ω α,β ∈FX X∈F X always close to ρ(t) σB, which leads to a Markovian dissipative dynamics⊗ for the system. where (0) (ii) Secular approximation with respect to H )S, i.e., i(ω0 ω)t (0) (0) oscillations e − are much faster than the dissipative Aαω = Aαω = Pk0 AαPk , (13) dynamics. Formally, Γ(ω) ω0 ω for all different ω ω ω 0 (0) (0) ∈F k,k : εk ε 0 =ω | |  | − | X X− k and ω0. (0) (iii) The functions Γαβ(ω) do not change significantly [HS ,Aαω] = ωAαω. (14) − within the clusters of Bohr frequencies emerging due to A rigorous result (a theorem) is given in AppendixA. the perturbative part δHS0 of the system Hamiltonian. Since the matrix γαβ(ω) is positive-semidefinite for an Formally, arbitrary ω, Eqs. (9) and (10) are of the first standard Re[Γ0 (ω)] ∆ω Re[Γαβ(ω)] , GKLS form. | αβ |  | | Im[Γ0 (ω)] ∆ω Im[Γ (ω)] , Equations (9) and (10) are different expressions of | αβ |  | αβ | the unified quantum master equation of weak-coupling where limit type. A simple algorithm of its construction in the ∆ω = max ω ω . Schr¨odingerpicture is as follows: ω ω | − | ∈F 4

B. Comparison with the secular master equation equation coincides with the master equation obtained in this limit, but with the refined Lamb-shift Hamiltonian. Let us describe explicitly the terms neglected in the We can simplify the Lamb-shift Hamiltonian HLS (12) secular (Davies) master equation and taken into account if we replace Sαβ(ω, ω0) by Sαβ(ω, ω) (i.e., perform the limit λ 0 in the arguments of the functions S , anal- in the presented unified master equation. We will use → αβ the common terms “populations” and “coherences” for, ogously to γαβ). In other words, both the dissipator and the Lamb-shift Hamiltonian would be constructed respectively, the diagonal and off-diagonal elements of (0) (0) the in some energy eigenbasis (not unique as if the system Hamiltonian was HS . If HS = 0, the if some levels are degenerate). If HS ei = εi ei and corresponding master equation is the well-known mas- H e = ε e , we say that the | i e ρ|e i cor- ter equation for the singular coupling limit. But in some S | ji j | ji h i| | ji responds to the Bohr frequency εj εi. cases (see AppendixC), solutions of this master equation The secular master equation describes− (i) transfer be- significantly deviate from the exact dynamics, while the tween populations, (ii) decay of coherences, (iii) trans- proposed master equation with the refined Lamb-shift fer between coherences corresponding to equal Bohr fre- Hamiltonian gives good results. So, we have obtained an quencies, and (iv) transfer between populations and co- improved version of the singular coupling limit master herences corresponding to the zero Bohr frequency (i.e., equation. coherences inside the eigensubspaces of HS). From the other side, we could keep the Lamb-shift The unified master equation describes the same pro- Hamiltonian from the Redfield equation HLS (i.e., with- out even partial secular approximation). The equation cesses with the following corrections: (iii0) transfer be- tween coherences corresponding to close Bohr frequen- would be still of the GKLS form, but without the de- sired thermodynamic properties. cies and (iv0) transfer between populations and coher- ences corresponding to Bohr frequencies which are equal or close to zero (i.e., coherences inside the eigensubspaces (0) E. Properties of HS ). This is illustrated on Fig.1 for the example given be- low. The Davies generator is well known to be covariant with respect to the unitary group eiHS t [15]. This simpli- fies the structure of the dynamics and the steady states. C. Comparison with other non-secular GKLS Moreover, this property is related to the total (system master equations and bath) energy conservation (hence, to the first law of thermodynamics) and for the resource theory of co- herence [12–14]. The unified generator (10) shares this Equations similar to the unified quantum master equa- iH(0)t tion (but not exactly the same) appeared in the frame- property, but with respect to the unitary group e S : work of the partial secular approximation [28–31, 48]. Let iH(0)t iH(0)t iH(0)t iH(0)t us compare our equation with the similar ones. Group- e− S ( ρ)e S = e− S ρ e S (15) ing the Bohr frequencies and taking the function γ in the L L centers of the clusters makes the equation simpler in com-  (0) This relation is satisfied in view of Eq. (14) and since HS parison with those in Refs. [28–30]. Also, due to this, the (0) (0) commutes with both H and H (H = P H P obtained equation has the desired properties important S LS S k k S k and the same is true for HLS). Note that HLS commutes for thermodynamics; see below. (0) P In contrast to one of equations in Ref. [48] (derived in with HS , but, in general, not with HS. more detail in Ref. [31]), which also adopts the clustering If the bath is thermal with the inverse temperature of the Bohr frequencies and Γ taken in the centers of β (not to be confused with the subindex), then the the clusters, the free dynamics in Eq. (9) (manifested Kubo–Martin–Schwinger (KMS) condition γαβ( ω) = βω − in the exponents) is defined by the original Hamiltonian e− γβα(ω) guarantees the stationarity of the thermal (0) (Gibbs) state with respect to the bare system Hamil- HS (not HS ). Also, the functions Sαβ in the Lamb-shift (0) (0) (0) βHS βHS Hamiltonian have their original arguments. tonian HS : ρβ = 0, where ρβ = e− / Tr e− . The same propertiesL guarantee that the quantum dynam- t ical semigroup eL satisfy the detailed balance property D. Particular cases. Refined Lamb-shift [1, 15, 16] with respect to ρβ. Hamiltonian Remark 1. Note that the true system steady state is ex- pected to be not the Gibbs state with respect to HS βHS If δHS = 0, then the unified master equation is re- (i.e., proportional to e− ), but the so-called mean force β(HS +λHI ) duced to the Davies master equation. In contrast, the Gibbs state, i.e., proportional to TrB e− . In (0) case H = 0 (i.e., all Bohr frequencies are small) is the zeroth order with respect to λ, it coincides with ρβ S (0) known to be equivalent to the so-called singular coupling (the Gibbs state with respect to HS ). However, the first limit [1–3, 46, 49, 50]. In this case, the unified master nontrivial correction (proportional to λ2) does not coin- 5 5

"11 0.10 0.5 cide with the Gibbs state with respect to HS, but takes D 0.08 > 0.4 01 e into account system-bath steady-state correlations [51]. !12 !1 !2 !12 ! ! »

0.06 r 0.3 » distance

"10 10

0.04 e 0.2

!0 <

Let the system weakly interact with several baths with @ "01 0.02 0.1 Trace the inverse temperatures βn and the corresponding gener- Re 0.00 0.0 2 !1 !2 ! ! 0 500 1000 1500 2000 0 500 1000 1500 2000 ators n, so that = i[HS, ]+λ n n. We can con- L L − · L Time fs Time fs sider the quantity of entropy production according to the "00 P Figure 2. Comparison of calculations of the dynamics of the general formalism [15]. It is a sum of the increase of the FigureFigure 1. 1. Scheme Scheme of of energy energy levels levels and and Bohr Bohr frequencies frequencies of two of two-qubit system according to the Redfield equation (blue von Neumann entropy of the system S(ρ) = Tr ρ ln ρ weaklytwo weakly interacting interacting qubits. qubits. Left: The Left: secular The secularquantum quantum master H L H L − dash-dotted line), secular (Davies) master equation (green and the entropy flows from the system to the baths: equationmaster equation treats all treats di↵erent all distinct energy levels Bohr andfrequencies Bohr frequen- as well dotted line), and the unified approach (red solid line). Left: ciesseparated. as well Right: separated Nearly and degenerate thus neglects Bohr the frequencies transfer between consti- the coherences marked by the red and blue lines, the transfer Trace distance to the numerically exact solution of HEOM. tute clusters. Namely, the frequencies ω1 and ω2 constitute a d 2 between populations and the coherences between the two mid- Right: The dynamics of the coherence that gives the main S(ρ(t)) λ βn Tr HS n[ρ(t)] . cluster with the center at ω = (ω1 +ω2)/2 and the frequencies dt − L dle almost degenerate levels, and the transfer between these contribution to the error for the secular master equation. The n 0 and ±ω0 constitute a cluster with the center at 0. X  coherences. Right: In the dissipator of the unified master solution of HEOM is shown by the black solid line. The pa- equation, almost degenerate energy levels or Bohr frequencies rameters are given in the main text. We see the breakdown of 4 Up to terms of the order O(λ ) (i.e., of a higher order are treated as exactly degenerate, while in the Lamb shift the secular master equation, while the unified master equation The eigenvalues and eigenvectors of HS are as follows: with respect to the second-order master equation), HS Hamiltonian they are still treated separately (with the par- correctly describes the dynamics. here can be substituted by H(0). Then the expression is tial secular approximation). This allows one to correctly de- S ε = + E2 + J 2, e = cos θ 11 + sin θ 00 , non-negative according to the analysis in Ref. [15] since scribe11 the mentioned12 transfers| 11i in the framework| i of completely| i positive dynamics. This is a result of a rigorous derivation. According to Eq. (13), A = A + A , A = ρ = 0. We cannot calculate the fourth-order con- q j! j!1 j!2 j0 n βn ε = E2 + J 2, e = cos θ 00 sin θ 11 , L 00 12 00 Aj0 + Aj! + Aj, ! ,whereAj! , Aj! , Aj! , Aj! , tribution to the entropy production using the second- − | i | i − | i 0 0 1 2 12 0 q 2 2 and Aj0 are the jump operators corresponding to indi- order master equation. For the same reason, all other "ε10 = + ∆E2 + J 2,, e10 = cos 'ϕ 0110 +sin 'ϕ 1001 ,, thermodynamic properties are also satisfied with respect 01 | 01 i | i | i vidual Bohr frequencies defined by Eq. (4) and are easily 2 2 | i | i | i (0) ε01 = p∆E + J , e01 = cos ϕ 01 1sin ϕ 10 J, to the zeroth-order system Hamiltonian H . In other where E12p= E1 + E2, E = E1 E2, ✓ = arctan , calculated. In particular, the x terms in the operators S − | i | i −2 | Ei12 words, resolution of small energy differences in the weak- and ' = p1 arctan J . The Bohr frequencies1 are shownJ Aj contribute to the frequencies !1 and !2,whilethez where E122 = E1 +EE2, ∆E = E1 E2, θ = 2 arctan E , − 12 terms contribute to the frequencies 0, !0, and !12. coupling regime requires higher order corrections to the onand Fig.ϕ =1.1 Wearctan considerJ . the The case Bohr of frequencies large E1 and areE2 shown,but 2 ∆E The Lamb shift Hamiltonian is just the Redfield Lamb dissipator. smallon Fig.1E. Weand considerJ, so that the case the energy of large levelsE and 01E and, but 10 1 2 shift Hamiltonian (3), where the terms with ! and ! aresmall almost ∆E degenerateand J, so that and close the energy to zero. levels Also 01 the and Bohr 10 0 from di↵erent clusters are neglected. frequenciesare almost! degenerate1 and !2 almost and close coincide. to zero. Also the Bohr Calculations. An example of calculations is presented frequenciesSo, we chooseω1 and theω following2 almost decompositioncoincide. (7): IV. EXAMPLE: TWO WEAKLY INTERACTING So, we choose the following decomposition (7): on Fig. 2. We have chosen the same Drude–Lorentz spec- QUBITS HS = "11P11 + "00P00 +("10P10 + "01P01) tral density for all three baths (see Supplemental ma- 2 terial for details), (!) g (k) 2(!(k) !) dk = HS =="ε1111PP1111 ++"ε0000PP0000 ++ (ε10("P0 10P10++ε01"P0 01P)01), j j 10 01 2 2 J ⌘ | 1| 1 2 2⌘⌦!/[⇡(! + ⌦ )] with ⌘ =1cm and ⌦ = 100 fs Consider a system of two weakly interacting qubits = ε11P11 + ε00P00 + λ2(ε100 P10 + ε010 P01), 1 R where Pab = eab eab . Here is a ratio of a charac- (⌦ 53.08 cm ). The temperatures of the baths are with the Hamiltonian | ih | ⇡ teristic large energy (i.e., "11) to2 a characteristic small T1 = 300 K, T2 = 400 K, and T0 = 350 K. The where Pab = eab eab . Here λ is a ratio of a charac- energy (e.g., "|10 ori h spectral| densities of the baths), so parameters of the system Hamiltonian are as follows: (1) (2) (1) (2) teristic large energy (i.e., ε11) to a characteristic small HS = E1σz + E2σz + Jσx σx , (16) 2 1 1 that "100 ,01 = "10,01/ are of the order of "11 and "00. E = E = 50 cm , J =2cm . These parameters of energy (e.g., ε10 or spectral densities of the baths), so 1 2 As we mentioned above,2 practically this is equivalent to the baths and the system Hamiltonian are mainly taken that ε100 ,01 = ε10,01/λ are of the order of ε11 and ε00. where σz = 1 1 0 0 and σx = 1 0 + 0 1 are putting "0 = "10,01, dealing with as with a formal from Ref. [42] as model parameters for excitation energy | i h | − | i h | | i h | | i h | With such10,01 decomposition, we have five clusters of Bohr the Pauli matrices, the superindex denotes a qubit, E1 small parameter, and putting =1intheend. transfer in a molecular dimer. Let us specify the inter- ≥ frequencies: the clusters +ω = ω1, ω2 and ω = E2 > 0. Let qubits interact with three bosonic ther- F { } F− Withω1, suchω2 with scaling, the thecenters Bohr at frequenciesω = ε11 are, the grouped cluster action Hamiltonian: jx = 0 and jz = 1 for j =1, 2, mal baths (with different temperatures). The interaction {− − } ± ± as follows:= ω , 0 thewith clusters the center+! at= 0,! and1, ! the2 and clusters! with= 0x = 1 and 1z = 0. The initial state is ⇢(0) = 01 01 . 0 0 Hamiltonian is HI = HI,0 +HI,1 +HI,2, HI,j = Aj Bj. F {± } F { } F | ih | single!1, elements!2 withω the= centersω12 and! = ω "=11,theclusterω12 . On Fig. 2, we compare solutions according to di↵erent (j) (j) ⊗ { } F 12 { } ± F− ±12 {− } Here Aj = κjxσx + κjzσz , where κjx and κjz are real 0Now= we!0 have, 0 with all information the center to at construct 0, and the the clusters unified approximate mater equations with the numerically exact (0) F {± } numbers and, for the unified notations, we put σ = withmaser single equation; elements see Appendix!12 = !B12 forand details.!12 = !12 . non-perturbative method of the hierarchical equation of z,x F { } F { } (1) (2) ThenAn example= of2 calculationsand the is dissipator presented incorrespond- Fig.2. We σz,x +σz,x. Further, Bj = dk [g (k)aj(k)+gj(k)a†(k)], j=0 j j motion (HEOM) in the high-temperature approximation, j j have chosenD the sameD Drude–Lorentz spectralD density for ing to the jth bath takes the form j = j!12 + j!+ j0, according to Ref. [42]. We see the breakdown of the sec- where aj(k)[a†(k)] is an annihilation [creation] operator D D D D j R whereall three baths,P ular master equation, while the unified master equation for the kth mode of the jth bath and g (k) are complex- j correctly describes the dynamics. valued functions. So, bath 0 is a common bath interact- 2 1 2ηΩω (ω()⇢)= g(!()k)Aδ(⇢ωA(k†) ω) dkA†=A , ⇢ , (17) Global vs local approaches. The question about the ing with both qubits and baths 1 and 2 are individual JDjj! ≡ j| j | j! j!− 2 j! πj(!ω2 + Ω2) baths for the corresponding qubits. The model has been Z ⇣ ⌘ applicability of local and global approaches is largely de- j ! 1 + j(!1 )e A1 † ⇢Aj! Aj!A† , ⇢ ,1 bated in the literature [23–30]. Our rigorous derivation taken from Ref. [37]. In general, a two-qubit system is with η = 1 cm− and Ω− j=! 100 fs (Ω2 53.j08! cm− ). ≈ is based on the global approach. However, in some cases, nontrivial (in particular, has nontrivial thermodynamic The temperatures of the⇣ baths are T1= 300 K, T⌘2 = ! is either !12 or !, and it coincides with the local one. Let, in the system Hamil- properties) and often used as a benchmark for compari- 400 K, and T0 = 350 K. The parameters of the system 1 tonian (17), E1 + E2 J and, hence, ✓ 0. This corre- son of various descriptions of dy- Hamiltonian are as follows: E1 = E1 2 =2 50 cm− , J = 1 (⇢)=j(0) A ⇢A A , ⇢ . ⇡ namics [25, 34, 36, 41, 42]. 2 cm− .D Thesej0 parametersj0 of thej0 baths2 j0 and the system spond to the rotating-wave approximation in the system ⇣ ⌘ 6

0.10 0.5

% is exactly the case where the local approach is shown 0.08 # 0.4 01

e to be better than the secular global one [32–35, 37, 40]. $

0.06 Ρ 0.3 $ However, the same factor γ (ω) in the dissipators of the

10 j

0.04 e 0.2 !

# two local baths instead of different factors γj(2E1) and 0.02 0.1 Trace distance Re γ (2E ) in the “naive” local approach make the obtained 0.00 0.0 j 2 0 500 1000 1500 2000 0 500 1000 1500 2000 local equation consistent with the second law of thermo- Time !fs" Time !fs" dynamics. Figure 2. Comparison of calculations of the dynamics of Thus, if the intersite couplings are small, the local ap- the two-qubit system according to the Redfield equation proach can be used, but the arguments of the dissipation (blue dash-dotted line), secular (Davies) master equation coefficients γ should be the same for the sites with close (green dotted line), and the unified approach (red solid line). Left: Trace distance to the numerically exact solution of the local energies. If the difference between the local ener- HEOM. Right: The dynamics of the coherence that gives the gies (∆E in our example) is large such that the secular main contribution to the error for the secular master equa- approximation can be used, but the intersite couplings tion. The solution of the HEOM is shown by the black solid are still small, then the local approach can be justified as line. The parameters are given in the text. an approximation to the secular or unified (i.e., global) quantum master equation [42].

Hamiltonian are mainly taken from Ref. [52] as model parameters for excitation energy transfer in a molecular B. Unified vs Redfield master equation dimer. Let us specify the interaction Hamiltonian: κjx = 0, and κjz = 1 for j = 1, 2, κ0x = 1 and κ0z = 0. The Our numerical results (see Fig.2 and also Fig.3 in initial state is ρ(0) = 01 01 . | i h | AppendixC) suggest that the Redfield equation is, in In Fig.2, we compare solutions according to differ- most cases, more precise that the unified GKLS master ent master equations with the numerically exact non- equation. This agrees with the results of Ref. [48]. So, perturbative method of the hierarchical equation of mo- the choice between the Redfield equation and the uni- tion (HEOM) in the high-temperature approximation, fied GKLS master equation depends on our purposes. according to Ref. [52]. We see the breakdown of the sec- If our priority is the numerical precision and we do not ular master equation, while the unified master equation care the possible small violation of positivity, then the correctly describes the dynamics. Redfield equation is preferable. If we want to have an It is worthwhile to recall that the HEOM is a numer- equation with good theoretical properties, with the ab- ically exact but computationally expensive method (in sence of nonphysical predictions, and with a reasonable terms of both time and memory), especially for the low- error, then the proposed unified GKLS master equation temperature case. Quantum master equations are much of weak-coupling limit type is a good candidate. simpler to solve numerically. Also, a relatively simple Also, a “hybrid” variant can be used to use the unified structure of the master equation gives more insights into GKLS master equation for the initial short-time period mechanisms of various quantum dynamical phenomena. (where the Redfield equation can violate positivity) and A two-qubit system interacting with high-temperature then to switch to the Redfield equation. Other concate- reservoirs was chosen since it is a simple (but nontriv- nation schemes for open quantum systems where different ial) model which can be solved also by the HEOM; i.e., descriptions are used for the initial short-time period and approximate descriptions can be compared with the nu- for further time scales are considered in Refs. [53, 54]. merically exact one.

C. Arbitrary scaling of the Bohr frequency spacing V. DISCUSSION Instead of Eq. (7), an arbitrary scaling of the Bohr A. Global vs local approach frequency spacing can be considered:

The local and global approaches to open quantum dy- (0) ν HS = HS + λ δHS, ν > 0. (18) namics are largely debated in the literature [32–42]. On the one hand, the unified approach is global, i.e., adopts Rigorous results for particular models were obtained in the eigenvalues and eigenprojectors of the whole system Refs. [43–45]. However, as these results suggest, a kind of Hamiltonian HS rather than of the Hamiltonian of nonin- a dynamical phase transition occurs exactly in the case teracting sites (corresponding to J = 0 in our example). ν = 2. This can be understood also in terms of the On the other hand, in our example, J E1 + E2, so, present analysis. Namely, ν < 2 means that the oscilla- we can neglect θ and approximately put e = 00 and iλν (ω0 ω)τ/λ2 | 00i | i tions e − in Eq. (2) are much faster than the e11 = 11 . This corresponds to the rotating-wave ap- dissipative dynamics. Hence, the full secular approxima- | i | i (0) proximation in the system Hamiltonian. Thus, HS cor- tion can be applied, which leads to the Davies master responds to J = 0 and leads to a local dissipator. This equation. 7

If, on the contrary, ν > 2, then these oscillations are canonical anticommutation relations (CAR) with an in- much slower than the dissipative dynamics. This leads to finite number of degrees of freedom. Denote the cor- emergence of the two time scales described in Refs. [43– responding Hilbert space by and the single-particle HB 45]: a relaxation to a manifold of quasistationary states Hilbert space by (1). For each f (1), there is a HB ∈ HB (stationary if δHS = 0) and a slower relaxation to a fi- bounded operator a(f) acting on B with an antilinear nal stationary state. Formal derivation for this scaling dependence on f which satisfies theH anticommutation re- according to Sec.IIIA again leads to the unified master lations equation. Namely, the secular approximation with re- (0) a(f)a†(g) + a†(g)a(f) = f g , spect to H and the limit λ 0 in the arguments of h | i S → γαβ are still valid. So, the unified master equation can a(f)a(g) + a(g)a(f) = 0. be applied irrespective of whether some oscillations occur The single-particle free evolution is given by ft = on the same time scale as the dissipative dynamics or on ith e− f, where h is the single-particle Hamiltonian (on a larger scale. (1) B ). There is a Hamiltonian HB on B and a cyclic Hvector Ω such that H Ω = 0 andH | i ∈ HB B | i VI. CONCLUSIONS iH t iH t iht e B a(f)e− B = a(e f).

(1) 2 d The unified approach allows one to derive the correct If, for example, B = L (R ), then the anticommuta- quantum master equation for a specific physical systems tion relations canH be rewritten in terms of the operator- (for a specific structure of energy levels) in a rigorous valued distributions: and systematic way. The unified master equation has a(k)a†(k0) + a†(k0)a(k) = δ(k k0), the GKLS form (hence, preserves positivity) and all the − desired properties important for thermodynamics: sta- a(k)a(k0) + a(k0)a(k) = 0, tionarity of the Gibbs state, the detailed balance condi- k, k d. The distribution and operator pictures are tion, the covariance law related to the first law of thermo- 0 R related∈ to each other by the formula dynamics, and non-negativity of the entropy production (the second law of thermodynamics). Thus, the unified a(f) = f(k)a(k) dk. quantum master equation can be used in a wide range of d physical applications. ZR If, further, hf(k) = ω(k)f(k) for a real-valued function itω(k) ω(k), then ft(k) = e− f(k) and ACKNOWLEDGMENTS HB = ω(k)a†(k)a(k) dk. d I am grateful to Marco Cattaneo, Dariusz Chru´sci´nski, ZR Luis A. Correa, Camille Lombard Latune, Angel´ Rivas, Writing for the expectation with respect to Ω , we Thomas Schulte-Herbr¨uggen,Ilya Sinayskiy, and Alexan- have h·i | i der Teretenkov for fruitful discussions and some biblio- a†(g ) a†(g )a(f ) a(f ) = δ det f R g graphic references. This work was funded by Russian h m ··· 1 1 ··· n i nm {h i| | ji} Federation represented by the Ministry of Science and (A2) (the quasi-free state property), where R is the defin- Higher Education (Grant No. 075-15-2020-788). (1) ing operator on B . Since HB Ω = 0, we have iht iht H | i e Re− = R. For the equilibrium state at the inverse temperature β and the chemical potential µ, Appendix A: Rigorous result β(h µ) 1 R = (e − + 1)− . We consider a quantum system specified by an either The full Hamiltonian is given by finite- or infinite-dimensional Hilbert space and a HS (λ) (λ) Hamiltonian of the form H = HS + HB + λHI ,

H H(λ) = H(0) + λ2δH , (A1) where the interaction Hamiltonian HI is given by a fi- S ≡ S S S nite sum H = A B with A and B being I α α ⊗ α α α (0) bounded operators in S and B, respectively. We as- where the operators HS and δHS are commuting self- P H H sume that each Bα is a linear combination of a(gα) and adjoint operators with purely discrete spectra. Hence, (1) the spectrum of the operator H(λ) is also purely discrete. a†(gα) for some gα B . In particular, Bα = 0. De- S iHB∈t H iHB t h i note Cαβ(t) = e Bα† e− Bβ the correlation func- Also we assume that the operator δHS is bounded. h i Like in the Davies’s paper [7], for simplicity, we con- tions and let sider the case when the bath is fermionic. Namely, the ∞ C (t) (1 + t)δ dt < (A3) bath is described by a quasi-free representation of the | αβ | ∞ Z0 8 for some δ > 0. is an operator on S. The evolution of the open quantum system is given by The norm of theB integrand in Eq. (A6) is upper bounded by the integrable function (λ) iH(λ)t iH(λ)t t ρ = TrB e− (ρ σB)e T ⊗ 4 A A C (s) ρ k α βk · | αβ | · k k for an arbitrary trace-class operator ρ on S, where σB = α,β Ω Ω . H X | i h | Remark 2. Let us stress that Ω is the vacuum vector independent from λ (the factor 4 comes from the two | i (λ) in a Fock space only in the case of the zero temperature. commutators in I and I resulting in four terms). (0) L L (λ) (0) 2 it 0 it 0 iλ tδ S Otherwise, this is a cyclic vector in the representation of Since HS and δHS commute, e L = e L e L . 2 the CAR algebra. In the physical literature, the notation Since δH is bounded, eiλ tδ S 1 0 as λ 0. So, βH βH S L e− B / Tr e− B is used for a thermal state instead of by the Lebesgue’s dominatedk convergence− k → theorem,→ it is Ω Ω , but, strictly speaking, the former is not a genuine easy to show that (λ) (0) 0 as λ 0. Then (the | i h | βH density operator because the trace of e− B is ill defined proof is completelykK analogous−K k to → that of→ Theorem 1.2 of in the case of an infinite number of degrees of freedom. Ref. [8]), As a remedy, in the physical literature, one often con- siders a finite number N of the bath modes and then ( i (λ)+λ2 (λ))t ( i (λ)+λ2 (0))t lim sup e − LS K e − LS K = 0. tends this number to infinity. But, strictly speaking, this λ 0 0 λ2t τ − → ≤ ≤ 1 may cause issues with the order of the limit N (A7) → ∞ and the Bogolyubov–van Hove limit λ 0, t , It is interesting to note that the generator i (λ) + λ2t = τ = const. → → ∞ − LS λ2 (λ) differs from the usual Redfield generator, which So, in the rigorous derivation, an infinite number of K is i (λ) + λ2 (λ), where degrees of freedom is considered from the beginning, and − LS R a thermal state can be understood only in a “generalized” ∞ sense: as a functional on the CAR algebra. The state (λ) = Tr (λ)( s)(ρ σ ) ds. R − B LI LI − ⊗ B represented by a cyclic vector Ω with property (A2) is Z0 a formalization of this situation.| i Thus, physically, σ is   B The generator (λ) is similar to the Redfield generator. an arbitrary stationary state of the bath with property In particular, itK is also non-GKLS and does not preserve (A2). In particular, it can be associated with the thermal positivity. Its explicit form also can be described by state with an arbitrary temperature. Eqs. (2) and (3) if we replace γ(ω, ω0) and S(ω, ω0) there Theorem. Under the described assumptions, by γ(ω0, ω) and S(ω0, ω). The secular approximation with respect to the refer- (λ) (λ) 2 2 (0) ( i[HS +λ HLS, ]+λ )t lim sup t ρ e − · D ρ = 0 ence system Hamiltonian HS (which corresponds to a λ 0 0 λ2t τ kT − k → ≤ ≤ 1 partial secular approximation with respect to the original (A4) system Hamiltonian H(λ)) can be expressed as for any τ > 0 and a trace-class operator ρ on . S 1 HS (λ) T Here and HLS H are given in Eqs. (11) and (12) (0) (0) LS sec 1 i u i u and Ddenotes the≡ trace norm. = lim e LS e− LS du, (A8) k · k A T 2T T A →∞ Z− Proof. Let us introduce the operators (λ) = [H(λ) + 0 S for an arbitrary operator on the Banach space S. (λ) L (λ) A B HB, ], I = [HI , ], and I (s) = [HI (s), ] act- According to Theorem 1.4 of Ref. [8], ing on· theL Banach· space L of the trace-class· opera- B (λ) (λ) (0) 2 (0) 2 sec ( i S +λ )t ( i S +λ )t tors on S B, and the operators S = [HS , ] lim sup e − L A ρ e − L A ρ = 0 H ⊗ H L · λ 0 0 λ2t τ k − k and δ S = [δHS, ] acting on the Banach space S → ≤ ≤ 1 L · (λ) B (A9) of the trace-class operators on S. Here HI (s) = (λ) (λ) H for all ρ S. i(HS +HB )s i(HS +HB )s ∈ B e HI e− . Obviously, the secular approximation applied to (0) According to Theorems 1.2 and 1.3 of Ref. [8], Theo- and (0) gives the same result i ˜ + , where K is R − LLS D D rems 3.1–3.5 of Ref. [7] (with minor modifications noted as required [i.e., given by Eq. (11)] and ˜LS = [H˜LS, ]. in Ref. [49]), and assumptions of our theorem, Here, in the notations of the main text, L ·

(λ) (λ) 2 (λ) ( i S +λ )t ˜ lim sup t e − L K = 0, (A5) HLS = Sαβ(ω)Aα† ωAβω, (A10) λ 0 0 λ2t τ T − → ≤ ≤ 1 ω (0) α,β ∈FX X where Sαβ(ω) Sαβ(ω, ω). ≡ (0) ∞ (λ) The term δ is already secular since H and δH (λ)ρ = Tr (s) (ρ σ ) ds (A6) LS S S K − B LI LI ⊗ B commute. Hence, the secular approximation applied to Z0   9

(0) (0) iδ S + and iδ S + gives the same result. but our numerical results suggest that the Redfield gener- −Since,L dueK to Eq. (A9−),L the correspondingR semigroups are ator is more precise than the “nonsecular Davies” gener- close to the same third semigroup, they are close to each ator i (λ) + λ2 (λ). Also, our unified GKLS generator − LS K other in the same sense: i (λ) iλ2 (λ) + λ2 is more precise than the GKSL − LS − LLS D (λ) 2 (0) (λ) 2 (0) generator with the same , but the Lamb-shift Hamilto- ( i S +λ )t ( i S +λ )t lim sup e − L K ρ e − L R ρ = 0 nian constructed from theD operator (λ) instead of (λ) λ 0 0 λ2t τ k − k → ≤ ≤ 1 K R (A11) [i.e., with S(ω, ω0) in Eq. (12) replaced by S(ω0, ω)]. Also, mathematically, it would be more natural to deal for all ρ S. ∈ B (λ) (λ) (λ) with the generator with the same and the simplified The decomposition of = i LS + into the D R(λ) − L(λ) D Lamb-shift Hamiltonian (A10), i.e., with the generator Lamb-shift Hamiltonian = [H , ] and the dissi- obtained by the application of the secular approximation LLS LS · (λ) (λ) to (0) or, equivalently, (0). Such possibility was dis- pator [cf. Eq. (2), where HLS HLS because HLS R K D (λ) ≡ cussed in the main text: In this case, both the dissipator depends on the spectrum of H ] can be expressed in S and the Lamb-shift Hamiltonian are constructed as if the the general form as follows: (0) system Hamiltonian was HS . Of course, (λ) ( i[H(λ)+λ2H˜ , ]+λ2 )t lim sup ρ e − S LS · D ρ = 0 (λ) ∞ t 1 λ 0 2 kT − k ρ = Tr [H H ( s) H ( s)H , ρ σ ] ds, → 0 λ t τ1 LLS 2i B I I − − I − I ⊗ B ≤ ≤ Z0 (A14) is also true as a consequence of Eqs. (A5), (A7), and (A9) (so, the proof is easier). However, in Appendix D, we give ∞ (λ) an example where such generator gives significant error, (λ)ρ = Tr H (ρ σ )H ( s) D B I ⊗ B I − while the proposed generator with a refined Lamb-shift Z0 h(λ) Hamiltonian gives good results. + H ( s)(ρ σB)HI I − ⊗ Remark 4. Expressions (A4) and (A14) assert that the 1 (λ) (λ) quantum dynamical semigroups approach the exact re- HI HI ( s) + HI ( s)HI , ρ σB ds. − 2 − − ⊗ duced dynamics in the limit λ 0 on arbitrarily long but  i → Analogously to the proof of Eq. (A7), we can prove finite time segments [0, τ1] (in the rescaled time). This restriction is important if we study the long-time limit that 2 limτ ρ(τ/λ ). Recently [55], for the standard weak- →∞ (λ) ( i (λ) iλ2 +λ2 (λ))t coupling limit (i.e., leading to the secular master equa- lim sup e − LS − LLS D tion), it was shown that the quantum dynamical semi- λ 0 0 λ2t τ → ≤ ≤ 1 group approaches the exact reduced dynamics uniformly (λ) (λ) ( i iλ2 +λ2 (0))t e − LS − LLS D = 0. (A12) on the whole time half-line τ [0, ). Probably, the − same result can be proved also∈ for our∞ limiting regime

(λ) (A1). Finally, due to Eq. (A9) applied to = iδ S i LS + (0), we obtain A − L − L But, nevertheless, we know (see the same paper D Ref. [55] for a review of the rigorous results about

(λ) this) that, if the bath is in thermal equilibrium and ( i (λ) iλ2 +λ2 (0))t lim sup e − LS − LLS D ρ certain conditions are met, ρ(t) tends to the state λ 0 2 → 0 λ t τ1 βH(λ) βH(λ) ≤ ≤ TrB e− / Tr e− as t for an arbitrary ini- ( i (λ) iλ2 (λ)+λ2 )t → ∞ e − LS − LLS D ρ = 0, (A13) tial state ρ(0). As λ 0, this stationary state tends to − βH(0) βH→(0) ρβ = e− S / Tr e− S . As we know from the main (λ) (λ) (λ) for all ρ S, where LS = [HLS , ] and HLS is ob- text, the unified master equation correctly predicts the ∈ B L · (λ) stationarity of this state. Under the same conditions, this tained by dropping the nonsecular terms from HLS and is given by Eq. (12). stationary state is unique. A combination of Eqs. (A5), (A7), (A11), (A12), and (A13) gives the required Eq. (A4). Appendix B: Details of calculations in the example Remark 3. The proof is a bit cumbersome because we wanted to have the Lamb-shift Hamiltonian from the The unified quantum master equation for the consid- Redfield equation (up to the secular approximation with ered example has the form (0) 2 respect to HS ). However, the Davies method gives an- (λ) 2 (λ) ρ˙ = i[HS, ρ] + jρ, other generator i S + λ , which is different from − L − L K j=0 the Redfield one i (λ) + λ2 (λ). Both generators give X − LS R (j) the same result after the secular approximation and the where j = i[HLS , ] + j is the generator corre- Davies method does not say which one is more precise, spondingL to the− interaction· D with the jth bath. Here 10

j = jω12 + jω + j0, [recall that S(ω) S(ω, ω)]. D D D D For the baths j≡= 0, 1, 2, we choose the same Drude– 1 Lorentz spectral density j(ω)(17). The correlation ρ = γ (ω) A ρA† A† A , ρ J Djω j jω jω − 2 jω jω function of the jth reservoir can be expressed as   βj ω  1 + γj(ω)e− Ajω† ρAjω AjωAjω† , ρ , C (s) = B (s)B − 2 j h j ji    ∞ βω ω is either ω or ω, and = j(ω) coth cos ωs i sin ωs dω. (B1) 12 J 2 − Z0     1 2 iH s iH s ρ = γj(0) A ρA A , ρ . Here B (s) = e B B e− B and denotes the expec- Dj0 j0 j0 − 2 j0 j j h·i   tation with respect to the thermal state with the inverse  temperature β. The jump operator Ajω12 corresponds to an individual Bohr frequency of HS and defined by Eq. (4): Ajω = We adopt the high-temperature approximation βΩ 12 1 (j) 1. For example, for the used value Ω 53.08 cm− κ P σz P . Let us discuss the operators A and A zj 00 11 jω j0 and the minimal considered temperature ≈T = 300 K, we corresponding to clusters of Bohr frequencies of HS, or, have βΩ 0.24. We have used that β = 1/kBT , where equivalently, to individual Bohr frequencies of ≈ 1 k 0.734 cm− /K is the Boltzmann constant. In this B ≈ (0) case, coth(βω/2) in Eq. (B1) can be approximated as H = ε (P P ) + 0 (P + P ). S 11 11 − 00 · 01 10 2/(βω) and As we see, the eigenprojectors of H(0) are P (0) = P , S + 11 2 Ωs (0) (0) Cj(s) ηΩ i e− , (B2) P = P00, and P0 = P01 + P10. ≈ βΩ − −According to Eq. (13),  

(0) (0) (0) (0) ∞ iωs ηΩ 2 Ajω = P AjP0 + P0 AjP+ Γj(ω) = Cj(s)e ds i . (B3) − 0 ≈ Ω iω βΩ − = P00Aj(P01 + P10) + (P01 + P10)AjP11, Z −   (j) (j) So, we have obtained the functions Γj(ω) which define = κxj[P00σ (P01 + P10) + (P01 + P10)σ P11], x x both the dissipation rates γ (ω) = 2 Re Γ (ω) and the (0) (0) (0) (0) (0) (0) j j Aj0 = P0 AjP0 + P+ AjP+ + P AjP Lamb shifts of the energy levels Sj(ω, ω0) = [Γj(ω) − − − (j) Γj∗(ω0)]/2i. = κzj[(P01 + P10)σz (P01 + P10) (j) (j) + P00σz P00 + P11σz P11], Appendix C: Importance of the refined Lamb-shift Alternatively, these operators can be calculated as sums Hamiltonian of the operators corresponding to the individual Bohr frequencies of HS from the corresponding clusters: In this section, we will see that the use of the refined Lamb-shift Hamiltonian (12) instead of the simplified one A = A + A , jω jω1 jω2 (A10) is important in some cases. This is surprising since Aj0 = Aj0 + Ajω0 + Aj, ω0 , the Lamb shift is often believed to be insignificant for the − dissipation processes at all. Namely, we will consider the where case when there is a nontrivial manifold of states station- ary for the dissipator, but only one of them commutes A = κ (P σ(j)P + P σ(j)P ), jω1 xj 00 x 10 01 x 11 with the Hamiltonian. So, the Hamiltonian part and, in (j) (j) Ajω2 = κxj(P00σx P01 + P10σx P11), particular, the refined Lamb-shift Hamiltonian are cru- cial for the dynamics toward the unique stationary state Aj0 = PµAjPµ, in a manifold of states that are stationary for the dis- µ=00,01,10,11 X sipator only. The existence of two different time scales A = A† = P A P . jω0 j, ω0 01 j 10 (dynamics toward a quasi-stationary manifold and dy- − namics toward the unique final equilibrium) for an open Further, quantum system with nearly degenerate energy levels is rigorously established in Refs. [44, 45]. (j) Let us consider the example from the main text and H = S(ω12)Ajω† Ajω12 + S( ω12)Ajω12 Ajω† LS 12 − 12 further simplify it: Let two identical qubits interact with + [S(ω, ω0)Ajω† 0 Ajω + S( ω, ω0)Ajω0 Ajω† ] each other and with two local dephasing baths. Namely, 0 − − ω,ω ω1,ω2 consider two identical interacting qubits with the Hamil- ∈{X } tonian + S(ω, ω0)Ajω† 0 Ajω 0 ω,ω 0, ω0 (1) (2) (1) (2) ∈{X± } HS = E(σz + σz ) + Jσx σx , 11 coupled to two thermal baths of harmonic oscillators. So, and the full Hamiltonian is ρ = γ(0)(ZρZ ρ), (C4) D − H = HS + HB,1 + HB,2 + HI,1 + HI,2, 2 (we have used that Z = 1). Here γ(ω) = 2 Re[Γ1(ω) + where Γ2(ω)] and S(ω) = Im[Γ1(ω) + Γ2(ω)] . The “simplified” version of the unified master equation (due to a simpler form of the Lamb-shift Hamiltonian) is HB,j = ω(k)aj†(k)aj(k) dk, d ZR ρ˙(t) = iλ2[H + H˜ , ρ(t)] + λ2 [ρ(t)], (C5) − S LS D a (k)[a†(k)] is an annihilation [creation] operator for the j j where kth mode of the jth bath and ω(k) is a non-negative real- e valued function of k. The interaction Hamiltonians are H˜ = S(0)( + + + ). (C6) (j) LS HI,j = σz Bj, j = 1, 2, where | i h | |−i h−| ⊗ A comparison of performances of master equations (C2) and (C5) with the numerically exact result of the B = dk [g (k)a (k) + g (k)a†(k)] j j j j j hierarchical equations of motion (HEOM) (also with the Z high-temperature approximation [52]) is shown on Fig.3. and gj(k) are complex-valued functions. That is, in the We see that the secular master equation highly overesti- notations of the main text, we put κjx = 0, j = 0, 1, 2, mates the rate of decoherence, while the simplified uni- κ0z = 0, κ1z = κ2z = 1, and E1 = E2 = E. fied master equation significantly significantly underesti- Then, the reduced system dynamics in the subspace mates it. The Redfield equation and the unified master spanned by 01 , 10 is independent from that in the equation have good agreement with the numerically ex- {| i | i} subspace spanned by 00 , 11 . We choose the initial act result. state as ρ(0) = 01 01{|. So,i | thei} reduced system dynam- Remark 5. Note that, in this case, the simplified version ics is confined inside| i h the| subspace spanned by 01 , 10 of the unified master equation (C5) coincides with the and, thus, is reduced to the two-dimensional{| case.i | Leti} master equation derived in the singular coupling limit [1– us restrict our consideration to this subspace. Then, the 3, 46, 49, 50]. So, we have obtained an improved version system Hamiltonian is (C2) of this master equation, which, as we see, is crucial in some cases. HS = J( 01 10 + 10 01 ). (C1) | i h | | i h | Let us make some analytic explanations of these nu- Its eigenvaluese are J with the corresponding eigenvec- merical results. From the expression of the dissipator tors = ( 01 ±10 )/√2. The interaction Hamiltoni- (C4), we can see that a peculiarity of this system is that |±i | i ± | i j ans are then HI,j = ( 1) Z Bj, where the initial state ρ(0) = 01 01 as well as any state di- − ⊗ agonal in the “local” basis| i h 01| , 10 is stationary for {| i | i} Z = 01e 01 10 10 = + + + . the dissipator. However, only the state I/2, where I is | i h | − | i h | | i h−| |−i h | the identity operator [i.e., diagonal in both the local ba- We again choose the Drude–Lorentz spectral density (17) sis and the eigenbasis (global basis)], is stationary for 1 for both baths with the coupling strength η = 1 cm− and both the dissipator and the Hamiltonian part. Due to 1 1 the cut-off frequency Ω− = 100 fs (Ω 53.08 cm− ). this, the Hamiltonian part and, in particular, the Lamb- ≈ The temperatures of the baths are T1 = T2 = 300 K. The shift Hamiltonian are crucial for the dynamics toward the 1 parameter J of the system Hamiltonian is J = 2 cm− . unique stationary state in the manifold of states station- Now we should choose a proper decomposition of the ary for the dissipator. system Hamiltonian into the reference and perturbation But, as we see from Eq. (C6), the simplified approach parts. Since J is small with respect to the bath relaxation (the standard singular coupling limit) eliminates the rate Ω and comparable to the dissipation constant η, we Lamb-shift Hamiltonian: Since only differences between treat the whole system Hamiltonian (C1) as small (i.e., the energy levels matter, the addition of the same quan- the free dynamics takes place on the same time scale as tity S(0) to the levels has no impact. 2 (0) the dissipation): HS λ HS. In other words, HS = 0 The initial populations of the eigenstates of HS are ≡ equal: ρ(0) = 1/2. The equal populations are sta- and δHS = HS. As we mentioned in the main text, h±| |±i this limiting regimee is equivalente to the singular couplinge tionary since they correspond to the same energy level of limit.f e the reference system Hamiltonian (equal to zero). So, The unified master equation has the form the dynamics manifests itself only in the decoherence in the eigenbasis of HS. Denote x(t) = + ρ(t) and 2 2 h | |−i ρ˙(t) = iλ [HS + HLS, ρ(t)] + λ [ρ(t)], (C2) y(t) = ρ(t) + . Then, the unified approach gives such − D equationsh−| for the| i coherences: where e x˙ = i[2J + S(2J) S( 2J)]x + γ(0)(y x), (C7a) − − − − H = S(2J) + + + S( 2J) (C3) y˙ = +i[2J + S(2J) S( 2J)]y γ(0)(y x). (C7b) LS | i h | − |−i h−| − − − − 12

0.5 dissipation). The same fact can be expressed in the language of 0.4 eigenvalues of the system (C7). Indeed, the smallest

% eigenvalue in magnitude is $%

$ 0.3 Ρ $ γ(0) + γ(0)2 [2J + S(2J) S( 2J)]2. !"

# 0.2 − − − −

Re p If γ(0) is greater than the second term under the square 0.1 root, then it can be approximated by

0.0 [2J + S(2J) S( 2J)]2/2γ(0). 0 10 000 20 000 30 000 40 000 − − − Time !fs" Anyway, the neglection of the terms S( 2J) decreases Figure 3. Comparison of calculations of the dynamics of the ± two-qubit system according to the numerically exact HEOM this eigenvalue whenever the signs of J and S(2J) S( 2J) coincide. Moreover, if S(2J) S( 2J) > 2J−, method (black solid line), the Redfield equation (blue dash- − − − dotted line), secular (Davies) master equation (green dotted such neglect may cause a large error. line), the unified master equation (red solid line), and the sim- Recall that both J and S(ω) are small. In the formal plified unified master equation (with the simplified Lamb-shift theory, both are multiplied by λ2, so, S(2J) S( 2J) Hamiltonian), which, in this case, coincides with the known is of the order λ4. Hence, in the limit λ 0,− the− effect singular coupling master equation (purple dashed line). The of this term disappears. This is confirmed→ by the nu- plot shows the dynamics of an off-diagonal element (in the merical results if we multiply J and η by an infinitesimal eigenbasis of the system Hamiltonian). The secular master dimensionless parameter λ2. Thus, Eq. (A14) is true. equation largely overestimates the rate of decoherence, while However, for the concrete chosen parameters, the sim- the simplified unified master equation significantly underesti- mates it. The Redfield equation and the unified master equa- plified equation (the standard master equation for the tion have good agreement with the numerically exact result. singular coupling limit) gives an incorrect prediction for the decoherence rate, while the equation with the refined Lamb-shift Hamiltonian provides good results. The secular approach treats the energy levels as well sep- Summarizing, the refined Lamb-shift Hamiltonian is arated and thus neglects the coherence–coherence trans- important for the case when there exists a nontrivial fer between + ρ(t) and ρ(t) + . Namely, it ne- manifold of states stationary for the dissipator, but only glects the termsh | +γ|−i(0)y andh−| +γ(0)|x iin the right-hand one of them is stationary also for the Hamiltonian part sides of Eqs. (C7a) and (C7b), respectively. Doing so, (i.e., commutes with the Hamiltonian). In this case, the the secular approximation overestimates the decoherence Hamiltonian part is crucial for relaxation to this unique rate. state. If we are not in such situation, a simplified Lamb- The simplified unified approach corresponds to the re- shift Hamiltonian also can be used. Note that the con- placement of S( 2J) by S(0). So, the terms with S van- sidered example is practically important because this is a ish. If the signs± of J and S(2J) S( 2J) coincide, such model of excitation energy transfer in a molecular dimer neglect decreases the difference− of the− rotation rates for [52]. Finally, it is worthwhile to note that a bit higher the coherences (in the complex plane) and thus decreases precision of the Redfield equation in comparison with the the decoherence rate because, as we noticed above, the unified GKLS master equation seen on Fig.3 is achieved initial state is stationary for the dissipator and the dy- at the cost of small violation of positivity on the initial namics is caused by the unitary rotation (followed by the short time (not seen on the plot).

[1] R. Alicki and K. Lendi, Quantum Dynamical Semigroups [6] R. Alicki and R. Kosloff, Introduction to quantum ther- and Applications (Springer, Berlin, 2007). modynamics: History and prospects, in Thermodynamics [2] H.-P. Breuer and F. Petruccione, The Theory of Open in the Quantum Regime (Springer, Berlin, 2018), pp. 1– Quantum Systems (Oxford University Press, Oxford, UK, 33. 2002). [7] E. B. Davies, Markovian master equations, Commun. [3] A. Rivas and S. F. Huelga, Open Quantum Systems: An Math. Phys. 39, 91–110 (1974). introduction (Springer, Berlin, 2012). [8] E. B. Davies, Markovian master equations. II, Math. [4] V. May and O. K¨uhn, Charge and Energy Transfer Dy- Ann. 219, 147–158 (1976). namics in Molecular Systems (Wiley-VCH, Weinheim, [9] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Com- Germany, 2011). pletely positive dynamical semigroups of N-level systems, [5] R. Kosloff, Quantum thermodynamics: A dynamical J. Math. Phys. 17, 821–825 (1976). viewpoint, Entropy 15, 2100–2128 (2013). [10] G. Lindblad, On the generators of quantum dynamical 13

semigroups, Commun. Math. Phys. 48, 119–130 (1976). 115109 (2020). [11] D. Chru´sci´nskiand S. Pascazio, A brief history of the [31] C. L. Latune, I. Sinayskiy, and F. Petruccione, Negative GKLS equation, Open Sys. Inf. Dyn. 24, 1740001 (2017). contributions to entropy production induced by quantum [12] M. Lostaglio, K. Korzekwa, D. Jennings, and T. Rudolph, coherences, Phys. Rev. A 102, 042220 (2020). Quantum coherence, time-translation symmetry, and [32] H. Wichterich, M. J. Henrich, H.-P. Breuer, J. Gemmer, thermodynamics, Phys. Rev. X 5, 021001 (2015). and M. Michel, Modeling heat transport through com- [13] I. Marvian and R. W. Spekkens, How to quantify coher- pletely positive maps, Phys. Rev. E 76, 031115 (2007). ence: Distinguishing speakable and unspeakable notions, [33] A.´ Rivas, A. D. K. Plato, S. F. Huelga, and M. B. Plenio, Phys. Rev. A 94, 052324 (2016). Markovian master equations: a critical study, New J. [14] R. Dann and R. Kosloff, Open system dynamics from Phys. 12, 113032 (2010). thermodynamic compatibility, Phys. Rev. Res. 3, 023006 [34] P. P. Hofer, M. Perarnau-Llobet, L. D. M. Miranda, (2021). G. Haack, R. Silva, J. B. Brask, and N. Brunner, Marko- [15] H. Spohn and J. Lebowitz, Irreversible thermodynam- vian master equations for quantum thermal machines: ics for quantum systems weakly coupled to thermal local versus global approach, New J. Phys. 19, 123037 reservoirs, in Advances in Chemical Physics, edited by (2017). S. A. Rice (Wiley Online Library, 1978), Vol. 38, pp. 109– [35] J. O. Gonz´alez, L. A. Correa, G. Nocerino, J. P. Palao, 142. D. Alonso, and G. Adesso, Testing the validity of the “lo- [16] V. Gorini, A. Frigerio, M. Verri, A. Kossakowski, and cal” and “global” GKLS master equations on an exactly E. C. G. Sudarshan, Properties of quantum Markovian solvable model, Open Syst. Inf. Dyn. 24, 1740010 (2017). master equations, Rep. Math. Phys. 13, 149–173 (1978). [36] G. L. De¸cordi and A. Vidiella-Barranco, Two coupled [17] H. Spohn, Entropy production for quantum dynamical qubits interacting with a thermal bath: A comparative semigroups, J. Math. Phys. 19, 1227–1230 (1978). study of different models, Opt. Commun. 387, 366–376 [18] A. G. Redfield, The theory of relaxation processes, Adv. (2017). Magn. Opt. Reson. 1, 1–32 (1965). [37] M. Cattaneo, G. L. Giorgi, S. Maniscalco, and [19] A. Ishizaki and G. R. Fleming, On the adequacy of the R. Zambrini, Local versus global master equation Redfield equation and related approaches to the study with common and separate baths: superiority of the of quantum dynamics in electronic energy transfer,J. global approach in partial secular approximation, New Chem. Phys. 130, 234110 (2009). J. Phys. 21, 113045 (2019). [20] A. Dodin, T. V. Tscherbul, and P. Brumer, Quantum dy- [38] A. I. Trubilko and A. M. Basharov, The effective Hamilto- namics of incoherently driven V-type systems: Analytic nian method in the thermodynamics of two resonantly in- solutions beyond the secular approximation, J. Chem. teracting quantum oscillators, J. Exp. Theor. Phys. 129, Phys. 144, 244108 (2016). 339–348 (2019). [21] P. R. Eastham, P. Kirton, H. M. Cammack, B. W. Lovett, [39] A. I. Trubilko and A. M. Basharov, Theory of relaxation and J. Keeling, Bath-induced coherence and the secular and pumping of quantum oscillator non-resonantly cou- approximation, Phys. Rev. A 94, 012110 (2016). pled with the other oscillator, Phys. Scr. 95, 045106 [22] R. Alicki, Master equations for a damped nonlinear oscil- (2020). lator and the validity of the Markovian approximation, [40] S. Scali, J. Anders, and L. A. Correa, Local master equa- Phys. Rev. A 40, 4077–4081 (1989). tions bypass the secular approximation, Quantum 5, 451 [23] G. Schaller and T. Brandes, Preservation of positivity by (2021). dynamical coarse graining, Phys. Rev. A 78, 022106 [41] A. Levy and R. Kosloff, The local approach to quantum (2008). transport may violate the second law of thermodynamics, [24] C. Majenz, T. Albash, H.-P. Breuer, and D. A. Lidar, EPL 107, 20004 (2014). Coarse graining can beat the rotating-wave approxima- [42] A. Trushechkin and I. Volovich, Perturbative treatment tion in quantum Markovian master equations, Phys. Rev. of inter-site couplings in the local description of open A 88, 012103 (2013). quantum networks, EPL 113, 30005 (2016). [25] J. D. Cresser and C. Facer, Coarse-graining in the deriva- [43] E. B. Davies, A model of atomic radiation, Ann. Inst. H. tion of Markovian master equations and its significance Poincar´e.Physique th´eorique 28, 91–110 (1978). in quantum thermodynamics, arXiv:1710.09939. [44] M. Merkli and H. Song, Overlapping resonances in open [26] A.´ Rivas, Refined weak-coupling limit: Coherence, entan- quantum systems, Ann. Henri Poincar´e 16, 1397–1427 glement, and non-Markovianity, Phys. Rev. A 95, 042104 (2014). (2017). [45] M. Merkli, H. Song, and G. P. Berman, Multiscale dy- [27] D. Farina and V. Giovannetti, Open-quantum-system dy- namics of open three-level quantum systems with two namics: Recovering positivity of the Redfield equation quasi-degenerate levels, J. Phys. A 48, 275304 (2015). via the partial secular approximation, Phys. Rev. A 100, [46] H. Spohn, Kinetic equations from Hamiltonian dynamics: 012107 (2019). Markovian limits, Rev. Mod. Phys. 53, 569–615 (1980). [28] K. Ptaszy´nski and M. Esposito, Thermodynamics of [47] A. S. Trushechkin, Higher-order corrections to the Red- quantum information flows, Phys. Rev. Lett. 122, 150603 field equation with respect to the system-bath cou- (2019). pling based on the hierarchical equations of motion, [29] G. McCauley, B. Cruikshank, D. I. Bondar, and Lobachevskii J. Math. 40, 1606–1618 (2019). K. Jacobs, Accurate Lindblad-form master equation [48] R. Hartmann and W. T. Strunz, Accuracy assessment of for weakly damped quantum systems across all regimes, perturbative master equations: Embracing nonpositivity, npj Quantum Inf. 6, 74 (2020). Phys. Rev. A 101, 012103 (2020). [30] F. Nathan and M. S. Rudner, Universal Lindblad equa- [49] P. F. Palmer, The singular coupling and weak coupling tion for open quantum systems, Phys. Rev. B 102, 14

limits, J. Math. Phys. 18, 527–529 (1977). proach, J. Chem. Phys. 130, 234111 (2009). [50] L. Accardi, A. Frigerio, and Y. G. Lu, On the relation be- [53] Y. C. Cheng and R. J. Silbey, Markovian approximation tween the singular and the weak coupling limits, Acta in the relaxation of open quantum systems, J. Phys. Appl. Math. 26, 197–208 (1992). Chem. B 109, 21399–21405 (2005). [51] J. D. Cresser and J. Anders, Weak and ultrastrong cou- [54] A. E. Teretenkov, Non-perturbative effects in corrections pling limits of the quantum mean force Gibbs state, to quantum master equations arising in Bogolubov–van arXiv:2104.12606. Hove limit, J. Phys. A 54, 265302 (2021). [52] A. Ishizaki and G. R. Fleming, Unified treatment of quan- [55] M. Merkli, Quantum Markovian master equations: Res- tum coherent and incoherent hopping dynamics in elec- onance theory shows validity for all time scales, Ann. tronic energy transfer: Reduced hierarchy equation ap- Phys. 412, 167996 (2020).