This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Markovian and non-Markovian master equations versus an exactly solvable model of a qubit in a cavity

Zihan Xia    Juan Garcia-Nila Center for Quantum Information Science & Technology Department of Electrical & Computer Engineering    Daniel A. Lidar Center for Quantum Information Science & Technology Department of Electrical & Computer Engineering Department of Physics & Astronomy Department of Chemistry
University of Southern California, Los Angeles, CA 90089, USA
Abstract

Quantum master equations are commonly used to model the dynamics of open quantum systems, but their accuracy is rarely compared with the analytical solution of exactly solvable models. In this work, we perform such a comparison for the damped Jaynes-Cummings model of a qubit in a leaky cavity, for which an analytical solution is available in the one-excitation subspace. We consider the non-Markovian time-convolutionless master equation up to the second (Redfield) and fourth orders as well as three types of Markovian master equations: the coarse-grained, cumulant, and standard rotating-wave approximation (RWA) Lindblad equations. We compare the exact solution to these master equations for three different spectral densities: impulse, Ohmic, and triangular. We demonstrate that the coarse-grained master equation outperforms the standard RWA-based Lindblad master equation for weak coupling or high qubit frequency (relative to the spectral density high-frequency cutoff ωc\omega_{c}), where the Markovian approximation is valid. In the presence of non-Markovian effects characterized by oscillatory, non-decaying behavior, the TCL approximation closely matches the exact solution for short evolution times (in units of ωc1\omega_{c}^{-1}) even outside the regime of validity of the Markovian approximations. For long evolution times, all master equations perform poorly, as quantified in terms of the trace-norm distance from the exact solution. The fourth-order time-convolutionless master equation achieves the top performance in all cases. Our results highlight the need for reliable approximation methods to describe open-system quantum dynamics beyond the short-time limit.

I Introduction

The study of open quantum systems presents both conceptual and technical challenges due to the complexity and high dimensionality of the environment, or bath. Exact analytical solutions describing the joint system-bath evolution are rarely attainable, necessitating the development of approximation methods to capture the reduced system dynamics [1, 2, 3]. To address this challenge, various approaches have been developed to derive master equations that describe the system’s evolution using the reduced density matrix, the best known of which is the Markovian Lindblad [or Gorini-Kossakowski-Lindblad-Sudharshan (GKLS)] equation [4, 5]. Numerous other master equations have been derived, some of which include non-Markovian effects. In some cases, rigorous error bounds have been derived that quantify the deviation between the solutions of known master equations and the exact solution [6]. However, these tend to be rather loose. Hence, it is desirable to compare the predictions of various master equations to non-trivial examples of exactly solvable open-system problems. This has been done, e.g., for the central spin model [7, 8].

This is the goal of the present work, where we study the damped Jaynes-Cummings model of a qubit inside a leaky cavity [9]. In this model, the qubit system interacts with the cavity electromagnetic field through the dipole approximation [10]. The qubit decay rate can be associated with experimentally measurable parameters such as the dipole moment and the energy gap [11]. Experimental proposals for simulating the spin-boson model with an Ohmic spectral density using superconducting circuits have been previously discussed [12, 13, 14]. We solve this model analytically assuming the zero temperature limit and the 11-excitation subspace of the joint qubit-cavity system, where the cavity is populated by at most a single photon. This is similar to previous studies assuming a Lorentzian spectral density [15, 16], but we do this for three different spectral densities: impulse, Ohmic, and triangular (formally defined below). These choices are motivated by experiments involving condensed matter systems such as superconducting qubits interacting with bosonic modes [17], rather than the original quantum-optical setting of an atom in a cavity that inspired the damped Jaynes-Cummings model. We then compare the exact solution to a number of different Markovian and non-Markovian master equations. We find that in the weak-coupling limit and for large qubit frequencies relative to the spectral density cutoff, where the Markovian approximation holds, the Lindblad equation derived using the coarse-graining approach [18] is more accurate than the standard rotating wave approximation based Lindblad equation. In the non-Markovian regime, the time-convolutionless master equation proves to be accurate in approximating the exact solution for relatively short evolution times. All the approximation methods we consider struggle at long evolution times.

This paper is structured as follows. In Section II, we introduce the damped Jaynes-Cummings model and derive the exact solutions for all three spectral densities. In Section III, we explore various Markovian approximation methods in the context of the damped Jaynes-Cummings model. We begin with the coarse-grained Lindblad equation (CG-LE) in Section III.1, then the cumulant LE (C-LE) in Section III.2, and finally the commonly used rotating-wave approximation (RWA)-based LE (RWA-LE, Section III.3). We apply the time-convolutionless (TCL) approach to second (TCL2, also known as the Redfield equation) and fourth orders (TCL4) in Section III.4. Section IV is the heart of this work, where we present comparisons between the exact solutions and the various approximation methods. This includes a comparison of the exact solution with Markovian and TCL approximations for the Ohmic spectral density, a comparison of the exact solution with the Markovian CG-LE and RWA-LE, and finally a comparison with the non-Markovian TCL models for the impulse and triangular spectral densities. We summarize our findings and present our conclusions in Section V. A variety of technical details that complement the main text are presented in the Appendices.

Readers who are already familiar with the different types of master equations may choose to skip Section III. All our key analytical results are conveniently accessible via Table 1, which provides the corresponding equation numbers. Readers who are interested primarily in the results of the comparison between the exact model results and the various master equations may choose to skip ahead to Section IV and focus on the graphs presented there.

II Exact Dynamics

II.1 General open system setup

The total Hamiltonian of the system and the bath is given by

H=H0+HSB,\displaystyle H=H_{0}+H_{SB}\ , (1)

where H0=HSIB+ISHBH_{0}=H_{S}\otimes I_{B}+I_{S}\otimes H_{B} with HSH_{S} and HBH_{B} being the pure system and bath Hamiltonians, respectively, II the identity operator.

We move to the interaction picture, where all operators transform according to:

XX~(t)=eiH0tXeiH0t.\displaystyle X\mapsto\tilde{X}(t)=e^{iH_{0}t}Xe^{-iH_{0}t}\ . (2)

The dynamics of the total system in the interaction picture are governed by the Liouville-von Neumann equation:

dρ~SBdt=i[H~SB,ρ~SB],\displaystyle\frac{d\tilde{\rho}_{SB}}{dt}=-i[\tilde{H}_{SB},\tilde{\rho}_{SB}]\ , (3)

where ρ~SB\tilde{\rho}_{SB} is the density matrix of the total system acting on the Hilbert space SB=SB\mathcal{H}_{SB}=\mathcal{H}_{S}\otimes\mathcal{H}_{B}. The joint system-bath state is thus given by:

ρ~SB(t)=U~(t)ρ~SB(0)U~(t),\displaystyle\tilde{\rho}_{SB}(t)=\tilde{U}(t)\tilde{\rho}_{SB}(0)\tilde{U}^{\dagger}(t)\ , (4)

where the unitary evolution operator is:

U~(t)=T+exp(i0tH~SB(t)𝑑t).\displaystyle\tilde{U}(t)=T_{+}\exp(-i\int_{0}^{t}\tilde{H}_{SB}(t^{\prime})dt^{\prime})\ . (5)

and T+T_{+} denotes Dyson time-ordering.

The solution can equivalently be expressed as a Dyson series by integrating and iterating Eq. 3:

ρ~SB(t)=ρSB(0)+n=1(i)n0t𝑑t10t1𝑑t2\displaystyle\tilde{\rho}_{SB}(t)=\rho_{SB}(0)+\sum_{n=1}^{\infty}(-i)^{n}\int_{0}^{t}dt_{1}\int_{0}^{t_{1}}dt_{2}\cdots (6)
0tn1dtn[H~SB(t1),[H~SB(t2),[H~SB(tn),ρSB(0)]].].\displaystyle\ \ \int_{0}^{t_{n-1}}dt_{n}[\tilde{H}_{SB}(t_{1}),[\tilde{H}_{SB}(t_{2}),...[\tilde{H}_{SB}(t_{n}),\rho_{SB}(0)]]....]\ .

The state of the system is given by the reduced density operator:

ρ~(t)=TrB[ρ~SB(t)],\displaystyle\tilde{\rho}(t)=\mathrm{Tr}_{B}[\tilde{\rho}_{SB}(t)]\ , (7)

where TrB\mathrm{Tr}_{B} denotes the partial trace over the bath state.

II.2 Model of a qubit in a leaky cavity

We analyze the dynamics of a single qubit inside a leaky cavity, coupled to a bosonic bath at zero temperature. By working in the single-excitation subspace this model becomes analytically solvable, and we closely follow the solution method of Refs. [10, 15] (see also Ref. [2]). However, we consider different bath spectral densities.

The system Hamiltonian, HSH_{S}, can be expressed as

HS=Ω0|11|=Ω0σ+σ.\displaystyle H_{S}=\Omega_{0}|{1}\rangle\!\langle 1|=\Omega_{0}\sigma_{+}\sigma_{-}\ . (8)

Here, σ+=|10|\sigma_{+}=|{1}\rangle\!\langle 0| and σ=|01|\sigma_{-}=|{0}\rangle\!\langle 1| are the raising and lowering operators for the qubit, respectively. The qubit ground state is |0\ket{0} with energy 0 and its excited state is |1\ket{1} with energy Ω0\Omega_{0}. The bath Hamiltonian, HBH_{B}, is given by:

HB=kωkbkbk=kωknk.\displaystyle H_{B}=\sum_{k}\omega_{k}b_{k}^{\dagger}b_{k}=\sum_{k}\omega_{k}n_{k}\ . (9)

Here, bkb_{k} and bkb_{k}^{\dagger} represent the annihilation and creation operators for the bosonic modes and nkn_{k} is the number operator for mode kk with energy ωk\omega_{k} (we set =1\hbar=1 throughout). The qubit-cavity interaction Hamiltonian is

HSB=σ+B+σB,\displaystyle H_{SB}=\sigma_{+}\otimes B+\sigma_{-}\otimes B^{\dagger}\ , (10)

where B=kgkbkB=\sum_{k}g_{k}b_{k}. Here the gkg_{k}’s are coupling constants with dimensions of energy. Introducing complex phases in the couplings can model chirality  [19].

In the interaction picture, the interaction Hamiltonian HSBH_{SB} becomes:

H~SB(t)\displaystyle\tilde{H}_{SB}(t) =σ+(t)B(t)+σ(t)B(t)\displaystyle=\sigma_{+}(t)\otimes B(t)+\sigma_{-}(t)\otimes B^{\dagger}(t) (11a)
σ±(t)\displaystyle\sigma_{\pm}(t) =e±iΩ0tσ±,B(t)=keiωktgkbk.\displaystyle=e^{\pm i\Omega_{0}t}\sigma_{\pm}\ ,\quad B(t)=\sum_{k}e^{-i\omega_{k}t}g_{k}b_{k}\ . (11b)

This model is not analytically solvable in general, but it is when we make the assumption that the cavity supports at most one photon. We thus consider the initial joint system-bath state to be:

|ϕ(0)=c0(0)|ψ0+c1(0)|ψ1+k𝔠k(0)|φk,\displaystyle\ket{\phi(0)}=c_{0}(0)\ket{\psi_{0}}+c_{1}(0)\ket{\psi_{1}}+\sum_{k}\mathfrak{c}_{k}(0)\ket{\varphi_{k}}\ , (12)

where

|ψ0\displaystyle\ket{\psi_{0}} =|0S|vB\displaystyle=\ket{0}_{S}\otimes\ket{v}_{B} (13a)
|ψ1\displaystyle\ket{\psi_{1}} =|1S|vB\displaystyle=\ket{1}_{S}\otimes\ket{v}_{B} (13b)
|φk\displaystyle\ket{\varphi_{k}} =|0S|kB.\displaystyle=\ket{0}_{S}\otimes\ket{k}_{B}\ . (13c)

Here |vB\ket{v}_{B} denotes the vacuum state of the cavity, and |kB=bk|vB=|01,,0k1,1k,0k+1,\ket{k}_{B}=b_{k}^{\dagger}\ket{v}_{B}=\ket{0_{1},\cdots,0_{k-1},1_{k},0_{k+1},\cdots} denotes the state with one photon in mode kk. The subspace spanned by {|ψ0,|ψ1,|φk}\{\ket{\psi_{0}},\ket{\psi_{1}},\ket{\varphi_{k}}\} is referred to as the 11-excitation subspace, and is conserved under the Hamiltonian in Eq. 1. I.e., the joint state remains in the following form for all time tt:

|ϕ(t)=c0(t)|ψ0+c1(t)|ψ1+k𝔠k(t)|φk,\displaystyle\ket{\phi(t)}=c_{0}(t)\ket{\psi_{0}}+c_{1}(t)\ket{\psi_{1}}+\sum_{k}\mathfrak{c}_{k}(t)\ket{\varphi_{k}}\ , (14)

subject to the normalization condition:

|c0(t)|2+|c1(t)|2+k|𝔠k(t)|2=1.\displaystyle|c_{0}(t)|^{2}+|c_{1}(t)|^{2}+\sum_{k}|\mathfrak{c}_{k}(t)|^{2}=1\ . (15)

The problem remains solvable with a more general bath state assuming interaction with a continuous-mode laser field [20].

We assume that initially there are no photons in the cavity [10], hence:

𝔠k(0)=0.\displaystyle\mathfrak{c}_{k}(0)=0\ . (16)

We introduce a spectral density J(ω)J(\omega) via:

k|gk|2eiωkt=0𝑑ωJ(ω)eiωt.\displaystyle\sum_{k}|g_{k}|^{2}e^{-i\omega_{k}t}=\int_{0}^{\infty}d\omega J(\omega)e^{-i\omega t}\ . (17)

The continuum of bath spectral modes is a necessary condition for irreversibility; a discrete spectrum necessarily results in recurrences.

To complete the model specification, we consider three different bath spectral densities: an impulse function centered at the cutoff frequency ωc\omega_{c}, an Ohmic function with the same cutoff frequency, and a triangular spectral density with a sharp cutoff, namely:

J1(ω)\displaystyle J_{1}(\omega) =|g|2δ(ωωc)\displaystyle=|g|^{2}\delta(\omega-\omega_{c}) (18a)
J2(ω)\displaystyle J_{2}(\omega) =ηωeω/ωc\displaystyle=\eta\omega e^{-\omega/\omega_{c}} (18b)
J3(ω)\displaystyle J_{3}(\omega) =ηωΘ(ωcω),\displaystyle=\eta\omega\Theta(\omega_{c}-\omega)\ , (18c)

where the Heaviside function obeys Θ(x)=0\Theta(x)=0 for x<0x<0 and Θ(x)=1\Theta(x)=1 for x0x\geq 0. In the impulse spectral density J1(ω)J_{1}(\omega), the bath has a singular response at the cutoff frequency, characterized by a Dirac delta function. The Ohmic spectral density J2(ω)J_{2}(\omega) is ubiquitous in the study of the spin-boson problem [21]. The dimensionless parameter η\eta in the Ohmic spectral density serves as a measure of the coupling strength between the bath and the system, while the ratio of the qubit frequency to the cutoff frequency measures the ratio of the photonic energy gap, which is the energy between the ground state and the excited state, and the number of frequency modes before reaching the cutoff frequency. The triangular spectral density J3(ω)J_{3}(\omega) is a sharp-cutoff approximation to the Ohmic spectral density, which we introduce to simplify analytical calculations. The case of a Lorentzian spectral density was studied in Ref. [2].

II.3 Exact Solution

Considering the entire qubit-cavity system as closed, its evolution is governed by the Schrödinger equation, which can be expressed as [see Appendix A]:

it|ϕ(t)\displaystyle i\partial_{t}\ket{\phi(t)} =kgk𝔠k(t)ei(Ω0ωk)t|ψ1\displaystyle=\sum_{k}g_{k}\mathfrak{c}_{k}(t)e^{i(\Omega_{0}-\omega_{k})t}\ket{\psi_{1}}
+kgkc1(t)ei(Ω0ωk)t|φk.\displaystyle+\sum_{k}g_{k}^{*}c_{1}(t)e^{-i(\Omega_{0}-\omega_{k})t}\ket{\varphi_{k}}\ . (19)

Multiplying this equation by ψ1|\bra{\psi_{1}} or φk|\bra{\varphi_{k}}, we obtain the following set of differential equations for the amplitudes:

c˙0(t)\displaystyle\dot{c}_{0}(t) =0\displaystyle=0 (20a)
c˙1(t)\displaystyle\dot{c}_{1}(t) =ikgk𝔠k(t)ei(Ω0ωk)t\displaystyle=-i\sum_{k}g_{k}\mathfrak{c}_{k}(t)e^{i(\Omega_{0}-\omega_{k})t} (20b)
𝔠˙k(t)\displaystyle\dot{\mathfrak{c}}_{k}(t) =igkc1(t)ei(Ω0ωk)t\displaystyle=-ig_{k}^{*}c_{1}(t)e^{-i(\Omega_{0}-\omega_{k})t} (20c)

Integrating, we arrive at:

c0(t)\displaystyle c_{0}(t) =c0(0)\displaystyle=c_{0}(0) (21a)
𝔠k(t)\displaystyle\mathfrak{c}_{k}(t) =i0t𝑑tgkc1(t)ei(Ω0ωk)t\displaystyle=-i\int_{0}^{t}dt^{\prime}g_{k}^{*}c_{1}(t^{\prime})e^{-i(\Omega_{0}-\omega_{k})t^{\prime}} (21b)

Let us define the memory kernel f(t)f(t) as the Fourier transform of the spectral density, shifted by the qubit frequency Ω0\Omega_{0}:

f(t)=0𝑑ωJ(ω)ei(Ω0ω)t.\displaystyle f(t)=\int_{0}^{\infty}d\omega J(\omega)e^{i(\Omega_{0}-\omega)t}\ . (22)

Substituting Eq. 21b into Eq. 20b, we obtain:

c˙1(t)=0t𝑑tf(tt)c1(t),\displaystyle\dot{c}_{1}(t)=-\int_{0}^{t}dt^{\prime}f(t-t^{\prime})c_{1}(t^{\prime})\ , (23)

which can be solved via a Laplace transform since the RHS is a convolution. Denoting the Laplace transform Lap\operatorname{Lap} of a general function g(t)g(t) by g^(s)\hat{g}(s), and recalling that Lap[g˙(t)]=sg^(s)g(0)\operatorname{Lap}[\dot{g}(t)]=s\hat{g}(s)-g(0), we have:

c^1(s)=c1(0)s+f^(s),\displaystyle\hat{c}_{1}(s)=\frac{c_{1}(0)}{s+\hat{f}(s)}\ , (24)

and c1(t)c_{1}(t) is then found via the inverse Laplace transform of Eq. 24.

Next, in order to determine the interaction picture system state by tracing out the bath state, we can utilize Eq. 14 and find:

ρ~(t)\displaystyle\tilde{\rho}(t) =TrB(|ϕ(t)ϕ(t)|)=(1|c1|2c0c1(t)c0c1(t)|c1|2),\displaystyle=\mathrm{Tr}_{B}\left(|{\phi(t)}\rangle\!\langle\phi(t)|\right)=\left(\begin{array}[]{cc}1-|c_{1}|^{2}&c_{0}c_{1}^{*}(t)\\ c_{0}^{*}c_{1}(t)&|c_{1}|^{2}\end{array}\right)\ , (27)

so that:

ρ~˙(t)=(t|c1|2c0c˙1(t)c0c˙1(t)t|c1|2).\displaystyle\dot{\tilde{\rho}}(t)=\left(\begin{array}[]{cc}-\partial_{t}|c_{1}|^{2}&c_{0}\dot{c}_{1}^{*}(t)\\ c_{0}^{*}\dot{c}_{1}(t)&\partial_{t}|c_{1}|^{2}\end{array}\right)\ . (30)

At this point, it is straightforward to verify that the dynamics are time-local,

ρ~˙\displaystyle\dot{\tilde{\rho}} =𝒦S(t)ρ~,\displaystyle=\mathcal{K}_{S}(t)\tilde{\rho}\ , (31)

with a generator given by:

𝒦S(t)ρ~(t)\displaystyle\mathcal{K}_{S}(t)\tilde{\rho}(t) =i2𝒮(t)[σ+σ,ρ~(t)]\displaystyle=-\frac{i}{2}\mathcal{S}(t)[\sigma_{+}\sigma_{-},\tilde{\rho}(t)]
+γ(t)(σρ~(t)σ+12{σ+σ,ρ~(t)}),\displaystyle+\gamma(t)\left(\sigma_{-}\tilde{\rho}(t)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\tilde{\rho}(t)\}\right)\ , (32)

provided we identify:

𝒮(t)\displaystyle\mathcal{S}(t) =2Im(c˙1(t)c1(t))\displaystyle=-2\imaginary\left(\frac{\dot{c}_{1}(t)}{c_{1}(t)}\right) (33a)
γ(t)\displaystyle\gamma(t) =2Re(c˙1(t)c1(t))\displaystyle=-2\real\left(\frac{\dot{c}_{1}(t)}{c_{1}(t)}\right) (33b)

The rate γ(t)\gamma(t) can be negative, corresponding to non-Markovian dynamics according to the CP non-divisibility criterion [22].

We focus on the excited state population and the coherence, which evolve in time according to:

ρ~11(t)\displaystyle\tilde{\rho}_{11}(t) =|c1(t)|2=|c1(0)|2exp(0tγ(t)𝑑t)\displaystyle=|c_{1}(t)|^{2}=|c_{1}(0)|^{2}\exp{-\int_{0}^{t}\gamma(t^{\prime})dt^{\prime}} (34a)
ρ~01(t)\displaystyle\tilde{\rho}_{01}(t) =c0c1(t)=c0c1(0)exp(120t(i𝒮(t)γ(t))𝑑t).\displaystyle=c_{0}{c}_{1}^{*}(t)=c_{0}{c}_{1}^{*}(0)\exp{\frac{1}{2}\int_{0}^{t}\left(i\mathcal{S}(t^{\prime})-\gamma(t^{\prime})\right)dt^{\prime}}\ . (34b)

Details of the derivation above can be found in Appendix A.

We next discuss the solutions for the three spectral densities of Eq. 18. In each case, we express the solution in terms of c1(t)c_{1}(t) or its Laplace transform.

II.4 Exact solution for three different spectral densities

II.4.1 J1=|g|2δ(ωωc)J_{1}=|g|^{2}\delta(\omega-\omega_{c})

As a toy example, we consider the impulse bath spectral density, which replaces the continuum of bath modes with a single mode. Consequently, we do not expect irreversibility and indeed, the solution is oscillatory. To demonstrate this, we circumvent the Laplace transform and instead use Eq. 22 to write:

f(t)\displaystyle f(t) =|g|2ei(Ω0ωc)t.\displaystyle=|g|^{2}e^{i(\Omega_{0}-\omega_{c})t}\ . (35)

As a result, Eq. 23 becomes:

c˙1(t)\displaystyle\dot{c}_{1}(t) =|g|20t𝑑tei(Ω0ωc)(tt)c1(t).\displaystyle=-|g|^{2}\int_{0}^{t}dt^{\prime}e^{i(\Omega_{0}-\omega_{c})(t-t^{\prime})}c_{1}(t^{\prime})\ . (36)

Differentiating both sides yields:

c¨1(t)i(Ω0ωc)c˙1(t)+|g|2c1(t)=0,\displaystyle\ddot{c}_{1}(t)-i(\Omega_{0}-\omega_{c})\dot{c}_{1}(t)+|g|^{2}c_{1}(t)=0\ , (37)

whose solution is:

c1(t)c1(0)\displaystyle\frac{c_{1}(t)}{c_{1}(0)} =ei(Ω0ωc)t/2(cos(t2δ)iΩ0ωcδsin(t2δ)),\displaystyle=e^{i(\Omega_{0}-\omega_{c})t/2}\left(\cos(\frac{t}{2}\delta)-i\frac{\Omega_{0}-\omega_{c}}{\delta}\sin(\frac{t}{2}\delta)\right), (38)

where δ\delta is a real number:

δ=(Ω0ωc)2+4g2.\displaystyle\delta=\sqrt{(\Omega_{0}-\omega_{c})^{2}+4g^{2}}\ . (39)

Since the solution is perfectly periodic, any master equation approximation with a non-trivial dissipator term will deviate from this exact solution for sufficiently long times. Note that the excited state population |c1(t)|2|c1(0)|2[(Ω0ωc)2(Ω0ωc)2+4g2,1]|c_{1}(t)|^{2}\in|c_{1}(0)|^{2}[\frac{(\Omega_{0}-\omega_{c})^{2}}{(\Omega_{0}-\omega_{c})^{2}+4g^{2}},1].

The Lamb shift and decay rate are now found from Eq. 33 to be:

𝒮(t)\displaystyle\mathcal{S}(t) =(1(Ω0ωc)2δ2)Ω0ωcδcot2(tδ2)+(Ω0ωc)2δ2\displaystyle=\left(1-\frac{(\Omega_{0}-\omega_{c})^{2}}{\delta^{2}}\right)\frac{\frac{\Omega_{0}-\omega_{c}}{\delta}}{\cot^{2}\left(\frac{t\delta}{2}\right)+\frac{(\Omega_{0}-\omega_{c})^{2}}{\delta^{2}}} (40a)
γ(t)\displaystyle\gamma(t) =(1(Ω0ωc)2δ2)cot(tδ2)cot2(tδ2)+(Ω0ωc)2δ2\displaystyle=\left(1-\frac{(\Omega_{0}-\omega_{c})^{2}}{\delta^{2}}\right)\frac{\cot\left(\frac{t\delta}{2}\right)}{\cot^{2}\left(\frac{t\delta}{2}\right)+\frac{(\Omega_{0}-\omega_{c})^{2}}{\delta^{2}}} (40b)

II.4.2 J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}

Now we turn our attention to the Ohmic spectral density. The memory kernel, as given in Eq. 22, takes the form:

f(t)=ηωc2eiΩ0t(1+iωct)2.\displaystyle f(t)=\frac{\eta\omega_{c}^{2}e^{i\Omega_{0}t}}{(1+i\omega_{c}t)^{2}}\ . (41)

To obtain the Laplace transform analytically, we integrate by parts and obtain:

f^(s)=\displaystyle\hat{f}(s)= η[(siΩ0)eΩ0+isωc(iπ2Ci(siΩ0ωc)\displaystyle\eta\Big{[}\left(s-i\Omega_{0}\right)e^{-\frac{\Omega_{0}+is}{\omega_{c}}}\left(i\frac{\pi}{2}-\operatorname{Ci}\left(\frac{s-i\Omega_{0}}{\omega_{c}}\right)\right.
iSi(siΩ0ωc))iωc],\displaystyle\left.\quad-i\operatorname{Si}\left(\frac{s-i\Omega_{0}}{\omega_{c}}\right)\right)-i\omega_{c}\Big{]}\ , (42)

where the sine and cosine integral functions are defined as:

Si(z)\displaystyle\operatorname{Si}(z) 0zsin(t)t𝑑t=π/2zsin(t)t𝑑t\displaystyle\equiv\int_{0}^{z}\frac{\sin(t)}{t}dt=\pi/2-\int_{z}^{\infty}\frac{\sin(t)}{t}dt (43a)
Ci(z)\displaystyle\operatorname{Ci}(z) zcos(t)t𝑑t\displaystyle\equiv-\int_{z}^{\infty}\frac{\cos(t)}{t}dt (43b)
Ei(z)\displaystyle\operatorname{Ei}(z) zett𝑑t,\displaystyle\equiv-\int_{-z}^{\infty}\frac{e^{-t}}{t}dt\ , (43c)

where for future reference we have also defined the exponential integral function.

Due to the absence of analytic expressions for the Laplace transforms of the trigonometric integral functions, we use the numerical inverse Laplace transform of Eq. 24 to obtain c1(t)c_{1}(t). The Lamb shift and decay rate are then calculated numerically using Eq. 33.

II.4.3 J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega)

Finally, we compute the memory kernel for the triangular spectral density, for which we obtain:

f(t)=0ωcηωeit(Ω0ω)=ηeit(Ω0ωc)1eiωct+iωctt2.\displaystyle f(t)=\int_{0}^{\omega_{c}}\eta\omega e^{it(\Omega_{0}-\omega)}=\eta e^{it(\Omega_{0}-\omega_{c})}\frac{1-e^{i\omega_{c}t}+i\omega_{c}t}{t^{2}}\ . (44)

The Laplace transform is:

f^(s)=iηωcη(siΩ0)ln(siΩ0siΩ0+iωc),\displaystyle\hat{f}(s)=-i\eta\omega_{c}-\eta(s-i\Omega_{0})\ln\left(\frac{s-i\Omega_{0}}{s-i\Omega_{0}+i\omega_{c}}\right)\ , (45)

and once more the (numerical) inverse Laplace transform of Eq. 24 yields c1(t)c_{1}(t). Finally, the Lamb shift and decay rate are again computed numerically using Eq. 33.

III Approximation methods

In this section, we compute the excited state population predictions for the damped Jaynes-Cummings model of the previous section using three different Markovian master equations and using the TCL to second and fourth orders. In each case, we first provide a brief summary of the underlying theory of the corresponding master equation, both to assist the reader who may be unfamiliar with this theory and to establish our notation.

III.1 Coarse-Grained Lindblad Equation (CG-LE)

We follow the derivation of Ref. [23]. Let us choose a dimensionless, fixed and orthogonal system operator basis for (S)\mathcal{B}(\mathcal{H}_{S}) as {Sα}α=0M\{S_{\alpha}\}_{\alpha=0}^{M} with S0=IS_{0}=I and M=d21M=d^{2}-1, where d=dim(S)d=\dim(\mathcal{H}_{S}) and

Tr(SαSβ)=1Nαδαβ,\displaystyle\mathrm{Tr}(S_{\alpha}^{\dagger}S_{\beta})=\frac{1}{N_{\alpha}}\delta_{\alpha\beta}\ , (46)

where NαN_{\alpha} is a normalization factor. We can then always write the system-bath interaction Hamiltonian in the following form:

HSB=αgαSαBα,\displaystyle H_{SB}=\sum_{\alpha}g_{\alpha}S_{\alpha}\otimes B_{\alpha}\ , (47)

where {Bα}\{B_{\alpha}\} are dimensionless bath operators and {gα}\{g_{\alpha}\} are the coupling coefficients. In the interaction picture, with US(t)=eiHStU_{S}(t)=e^{-iH_{S}t} and UB(t)=eiHBtU_{B}(t)=e^{-iH_{B}t}, we obtain:

H~SB\displaystyle\tilde{H}_{SB} =αgαSα(t)Bα(t)\displaystyle=\sum_{\alpha}g_{\alpha}S_{\alpha}(t)\otimes B_{\alpha}(t) (48a)
Sα(t)\displaystyle S_{\alpha}(t) =US(t)SαUS(t)=βpαβ(t)Sβ\displaystyle=U_{S}^{\dagger}(t)S_{\alpha}U_{S}(t)=\sum_{\beta}p_{\alpha\beta}(t)S_{\beta} (48b)
Bα(t)\displaystyle B_{\alpha}(t) =UB(t)BαUB(t)=βqαβ(t)Bβ,\displaystyle=U_{B}^{\dagger}(t)B_{\alpha}U_{B}(t)=\sum_{\beta}q_{\alpha\beta}(t)B_{\beta}\ , (48c)

with initial conditions pαβ(0)=qαβ(0)=δαβp_{\alpha\beta}(0)=q_{\alpha\beta}(0)=\delta_{\alpha\beta}. The dynamics of the total system+bath are described by Eqs. 4 and 5. The system state is obtained by tracing over the bath and, assuming the initial state is factorized [ρSB(0)=ρ(0)ρB(0)\rho_{SB}(0)=\rho(0)\otimes\rho_{B}(0)], can be represented as a completely positive quantum dynamical map:

ρ~(t)=TrB[ρ~SB(t)]=i=0MK~i(t)ρ(0)K~i(t)\displaystyle\tilde{\rho}(t)=\mathrm{Tr}_{B}[\tilde{\rho}_{SB}(t)]=\sum_{i=0}^{M}\tilde{K}_{i}(t)\rho(0)\tilde{K}_{i}^{\dagger}(t) (49)

where {K~i}\{\tilde{K}_{i}\} are Kraus operators in the interaction picture, which are defined as:

K~i={μν}(t)=λμν|U~(t)|μ,\displaystyle\tilde{K}_{i=\{\mu\nu\}}(t)=\sqrt{\lambda_{\mu}}\bra{\nu}\tilde{U}(t)\ket{\mu}\ , (50)

where the initial bath state is spectrally decomposed as ρB(0)=μλμ|μμ|\rho_{B}(0)=\sum_{\mu}\lambda_{\mu}|{\mu}\rangle\!\langle\mu|. Using a standard Dyson series expansion of U~(t)\tilde{U}(t) similar to Eq. 6, we can write:

K~i(t)\displaystyle\tilde{K}_{i}(t) =λμδμνI+n=1Ki(n)(t)\displaystyle=\sqrt{\lambda_{\mu}}\delta_{\mu\nu}I+\sum_{n=1}^{\infty}K_{i}^{(n)}(t) (51a)
=α=0Mbiα(t)Sα=bi0I+α=1Mn=1biα(n)(t)Sα,\displaystyle=\sum_{\alpha=0}^{M}b_{i\alpha}(t)S_{\alpha}=b_{i0}I+\sum_{\alpha=1}^{M}\sum_{n=1}^{\infty}b^{(n)}_{i\alpha}(t)S_{\alpha}\ , (51b)

where Ki(n)(t)K_{i}^{(n)}(t) is the nnth order term in the Dyson series, and the second line is an expansion in the system operator basis. In the weak-coupling limit (maxαgαt1\max_{\alpha}g_{\alpha}t\ll 1), the higher-order terms in the expansion of Ki(n)(t)K_{i}^{(n)}(t) become negligible, i.e., Ki(n+1)gαtKi(n)\|K_{i}^{(n+1)}\|\sim g_{\alpha}t\|K_{i}^{(n)}\|. Therefore, we can approximate the exact Kraus operators by truncating the expansion to first order. This yields:

K~i(1)(t)\displaystyle\tilde{K}_{i}^{(1)}(t) =iλμν|0t𝑑t1H~SB(t1)|μ\displaystyle=-i\sqrt{\lambda_{\mu}}\bra{\nu}\int_{0}^{t}dt_{1}\tilde{H}_{SB}(t_{1})\ket{\mu} (52a)
=iλμαgα0t𝑑t1Sα(t1)ν|Bα(t1)|μ\displaystyle=-i\sqrt{\lambda_{\mu}}\sum_{\alpha}g_{\alpha}\int_{0}^{t}dt_{1}S_{\alpha}(t_{1})\bra{\nu}B_{\alpha}(t_{1})\ket{\mu} (52b)
=itλμαβγgαSβν|Bγ|μΓαβγ(t),\displaystyle=-it\sqrt{\lambda_{\mu}}\sum_{\alpha\beta\gamma}g_{\alpha}S_{\beta}\bra{\nu}B_{\gamma}\ket{\mu}\Gamma_{\alpha}^{\beta\gamma}(t)\ , (52c)

where:

Γαβγ(t)1t0t𝑑t1pαβ(t1)qαγ(t1).\displaystyle\Gamma_{\alpha}^{\beta\gamma}(t)\equiv\frac{1}{t}\int_{0}^{t}dt_{1}p_{\alpha\beta}(t_{1})q_{\alpha\gamma}(t_{1})\ . (53)

Then, to match Eqs. 52c and 51b, we have:

biα(1)(t)=itλμαα′′gαν|Bα′′|μΓααα′′(t),\displaystyle b^{(1)}_{i\alpha}(t)=-it\sqrt{\lambda_{\mu}}\sum_{\alpha^{\prime}\alpha^{\prime\prime}}g_{\alpha^{\prime}}\bra{\nu}B_{\alpha^{\prime\prime}}\ket{\mu}\Gamma^{\alpha\alpha^{\prime\prime}}_{\alpha^{\prime}}(t)\ , (54)

and Eq. 51 implies that bi0=λμδμνb_{i0}=\sqrt{\lambda_{\mu}}\delta_{\mu\nu}. Next, we can construct the process matrix χ(t)\chi(t) (closely related to the Choi matrix), where

χαβ(t)=i=μνbiα(t)biβ(t),\chi_{\alpha\beta}(t)=\sum_{i=\mu\nu}b_{i\alpha}(t)b^{*}_{i\beta}(t)\ , (55)

and truncate it to lowest order (n1n\leq 1) in the Dyson expansion as follows:

χ00(t)\displaystyle\chi_{00}(t) =μλμ=1\displaystyle=\sum_{\mu}\lambda_{\mu}=1 (56)
χα0(1)(t)\displaystyle\chi^{(1)}_{\alpha 0}(t) =itαα′′gαBα′′BΓααα′′(t),α1\displaystyle=-it\sum_{\alpha^{\prime}\alpha^{\prime\prime}}g_{\alpha^{\prime}}\expectationvalue{B_{\alpha^{\prime\prime}}}_{B}\Gamma_{\alpha^{\prime}}^{\alpha\alpha^{\prime\prime}}(t),~{}~{}\alpha\geq 1 (57)
χαβ(1)(t)\displaystyle\chi^{(1)}_{\alpha\beta}(t) =t2αα′′ββ′′gαgβBβ′′Bα′′B×\displaystyle=t^{2}\sum_{\alpha^{\prime}\alpha^{\prime\prime}\beta^{\prime}\beta^{\prime\prime}}g_{\alpha^{\prime}}g_{\beta^{\prime}}^{*}\expectationvalue{B^{\dagger}_{\beta^{\prime\prime}}B_{\alpha^{\prime\prime}}}_{B}\times
Γααα′′(t)(Γβββ′′(t)),α,β1,\displaystyle\Gamma^{\alpha\alpha^{\prime\prime}}_{\alpha^{\prime}}(t)\left(\Gamma^{\beta\beta^{\prime\prime}}_{\beta^{\prime}}(t)\right)^{*},~{}~{}\alpha,\beta\geq 1\ , (58)

where XBTr(ρBX)\expectationvalue{X}_{B}\equiv\mathrm{Tr}(\rho_{B}X). Let us also define

Xj1τjτ(j+1)τX(t)𝑑t,\expectationvalue{X}_{j}\equiv\frac{1}{\tau}\int_{j\tau}^{(j+1)\tau}X(t)dt\ , (59)

where we call τ\tau the coarse-graining timescale. Note that χαβ(0)=δα0δβ0\chi_{\alpha\beta}(0)=\delta_{\alpha 0}\delta_{\beta 0}. Then:

aαβχ˙αβ(1)0=1τ(χαβ(1)(τ)χαβ(0))=χαβ(1)(τ)τ\displaystyle a_{\alpha\beta}\equiv\expectationvalue{\dot{\chi}^{(1)}_{\alpha\beta}}_{0}=\frac{1}{\tau}(\chi^{(1)}_{\alpha\beta}(\tau)-\chi_{\alpha\beta}(0))=\frac{\chi^{(1)}_{\alpha\beta}(\tau)}{\tau} (60)

unless α=β=0\alpha=\beta=0, in which case we have χ˙000=0\expectationvalue{\dot{\chi}_{00}}_{0}=0.

It can be shown [23] (see also [24]) that starting from Eq. 49, substituting the various expansions above, rearranging terms, and assuming that Eq. 60 can be extended to any interval [t,t+τ][t,t+\tau] (essentially an assumption of Markovianity), that one arrives at the Lindblad equation in the interaction picture:

ρ~˙(t)\displaystyle\dot{\tilde{\rho}}(t) =i[HLS,ρ~(t)]\displaystyle=-i[H_{LS},\tilde{\rho}(t)]
+α,β=1Maαβ(Sαρ~(t)Sβ12{SβSα,ρ~(t)}),\displaystyle+\sum_{\alpha,\beta=1}^{M}a_{\alpha\beta}(S_{\alpha}\tilde{\rho}(t)S_{\beta}^{\dagger}-\frac{1}{2}\{S_{\beta}^{\dagger}S_{\alpha},\tilde{\rho}(t)\})\ , (61)

where the Lamb shift is given by:

HLS\displaystyle H_{LS} =i2αχ˙α0Sαχ˙α0Sα\displaystyle=\frac{i}{2}\sum_{\alpha}\expectationvalue{\dot{\chi}_{\alpha 0}}S_{\alpha}-\expectationvalue{\dot{\chi}_{\alpha 0}}^{*}S_{\alpha}^{\dagger} (62a)
=12αϕαSα+ϕαSα\displaystyle=\frac{1}{2}\sum_{\alpha}\phi_{\alpha}S_{\alpha}+\phi_{\alpha}^{*}S_{\alpha}^{\dagger} (62b)

with

ϕααα′′gαBα′′BΓααα′′(τ),\displaystyle\phi_{\alpha}\equiv\sum_{\alpha^{\prime}\alpha^{\prime\prime}}g_{\alpha^{\prime}}\expectationvalue{B_{\alpha^{\prime\prime}}}_{B}\Gamma^{\alpha\alpha^{\prime\prime}}_{\alpha^{\prime}}(\tau)\ , (63)

and the decoherence rates are:

aαβ=ταα′′ββ′′gαgβBβ′′Bα′′BΓααα′′(τ)Γβββ′′(τ).\displaystyle a_{\alpha\beta}=\tau\sum_{\alpha^{\prime}\alpha^{\prime\prime}\beta^{\prime}\beta^{\prime\prime}}g_{\alpha^{\prime}}g_{\beta^{\prime}}^{*}\expectationvalue{B^{\dagger}_{\beta^{\prime\prime}}B_{\alpha^{\prime\prime}}}_{B}\Gamma_{\alpha^{\prime}}^{\alpha\alpha^{\prime\prime}}(\tau)\Gamma_{\beta^{\prime}}^{\beta\beta^{\prime\prime}}(\tau)^{*}\ . (64)

The choice of the coarse-graining timescale τ\tau is crucial. It can be understood as a free optimization parameter, constrained by the inequality

τSτ1/ωc,\tau_{S}\ll\tau\ll 1/\omega_{c}\ , (65)

where τS\tau_{S} is the timescale over which ρ~(t)\tilde{\rho}(t) changes, which arises from the replacement of

ρ~˙(t)0ρ~(τ)ρ(0)τ\langle\dot{\tilde{\rho}}(t)\rangle_{0}\equiv\frac{\tilde{\rho}(\tau)-{\rho}(0)}{\tau} (66)

by ρ~˙(t)\dot{\tilde{\rho}}(t) in arriving at Eq. 61.

For the spin-boson model Eq. 10, we have S+=σ+S_{+}=\sigma_{+}, S=σS_{-}=\sigma_{-}, Bk=bkB_{k}=b_{k} or bkb_{k}^{\dagger}. The system-bath interaction Hamiltonian in the interaction picture is given by Eq. 11. From Eqs. 48b and 48c, we obtain:

p±±(t)\displaystyle p^{\pm\pm}(t) =e±iΩ0t,p±(t)=0\displaystyle=e^{\pm i\Omega_{0}t}\ ,\quad p^{\pm\mp}(t)=0 (67a)
qkk′′±±(t)\displaystyle q^{\pm\pm}_{kk^{\prime\prime}}(t) =δkk′′eiωkt,qkk′′±(t)=0,\displaystyle=\delta_{kk^{\prime\prime}}e^{\mp i\omega_{k}t}\ ,\quad q^{\pm\mp}_{kk^{\prime\prime}}(t)=0\ , (67b)

where the ++ or - superscripts indicate that the corresponding bath operator is bkb_{k} or bkb_{k}^{\dagger}, respectively.

Assuming that the initial state of the bath (cavity) is the zero-temperature vacuum state ρB(0)=|vv|\rho_{B}(0)=|{v}\rangle\!\langle v|, we obtain the standard bosonic expectation values:

bkblB\displaystyle\langle{b_{k}^{\dagger}b_{l}}\rangle_{B} =bkB=bkB=bkblB=bkblB=0\displaystyle=\langle{b_{k}^{\dagger}}\rangle_{B}=\langle{b_{k}}\rangle_{B}=\langle{b_{k}^{\dagger}b_{l}^{\dagger}}\rangle_{B}=\langle{b_{k}b_{l}}\rangle_{B}=0
bkblB\displaystyle\langle b_{k}b_{l}^{\dagger}\rangle_{B} =δkl.\displaystyle=\delta_{kl}. (68)

Then, by utilizing Eq. 53, we obtain a slightly modified expression for Γ\Gamma up to first order:

Γα,(αk)β,(βk′′)(t)=1t0t𝑑t1pαβ(t1)qkk′′αβ(t1),\displaystyle\Gamma_{\alpha,(\alpha^{\prime}k)}^{\beta,(\beta^{\prime}k^{\prime\prime})}(t)=\frac{1}{t}\int_{0}^{t}dt_{1}p^{\alpha\beta}(t_{1})q^{\alpha^{\prime}\beta^{\prime}}_{kk^{\prime\prime}}(t_{1})\ , (69)

where α,β,α,β{+,}\alpha,\beta,\alpha^{\prime},\beta^{\prime}\in\{+,-\}. Using Eq. 67, we have the following non-zero Γ\Gamma’s:

Γ+,(+k)+,(+k)(t)\displaystyle\Gamma_{+,(+k)}^{+,(+k)}(t) =ei(Ω0ωk)t1i(Ω0ωk)t\displaystyle=\frac{e^{i(\Omega_{0}-\omega_{k})t}-1}{i(\Omega_{0}-\omega_{k})t} (70a)
Γ,(k),(k)(t)\displaystyle\Gamma_{-,(-k)}^{-,(-k)}(t) =ei(Ω0ωk)t1i(Ω0ωk)t.\displaystyle=\frac{e^{-i(\Omega_{0}-\omega_{k})t}-1}{-i(\Omega_{0}-\omega_{k})t}\ . (70b)

We also have a slightly modified expression for bb by using Eq. 54:

bμν,α\displaystyle b_{\mu\nu,\alpha} =itλμ(αk)gkαμ|Bkα|νΓα,(αk)α,(αk),\displaystyle=-it\sqrt{\lambda_{\mu}}\sum_{(\alpha^{\prime}k^{\prime})}g^{\alpha^{\prime}}_{k^{\prime}}\bra{\mu}B^{\alpha^{\prime}}_{k^{\prime}}\ket{\nu}\Gamma^{\alpha,(\alpha^{\prime}k^{\prime})}_{\alpha,(\alpha^{\prime}k^{\prime})}\ , (71)

where BkαB^{\alpha}_{k} represents bkb_{k} when α=+\alpha=+ and bkb_{k}^{\dagger} when α=\alpha=-. The Lamb shift rates are given by Eq. 63, and vanish:

ϕα=(αk)gkαBkαBΓα,(αk)α,(αk)(τ)=0,\displaystyle\phi_{\alpha}=\sum_{(\alpha^{\prime}k^{\prime})}g^{\alpha^{\prime}}_{k^{\prime}}\expectationvalue{B^{\alpha^{\prime}}_{k^{\prime}}}_{B}\Gamma^{\alpha,(\alpha^{\prime}k^{\prime})}_{\alpha,(\alpha^{\prime}k^{\prime})}(\tau)=0\ , (72)

since the expectation values of creation and annihilation operators between vacuum states vanish, as indicated by Section III.1. This result will be seen to undermine the quality of the CG-LE and C-LE when we perform a comparison with the exact results in Section IV.

The decoherence rates are:

aαβ(τ)\displaystyle a_{\alpha\beta}(\tau) =τ(αk)(βl)gkαglβBlβBkαB\displaystyle=\tau\sum_{(\alpha^{\prime}k^{\prime})(\beta^{\prime}l^{\prime})}g^{\alpha^{\prime}}_{k^{\prime}}g_{l^{\prime}}^{-\beta^{\prime}}\expectationvalue{B_{l^{\prime}}^{-\beta^{\prime}}B^{\alpha^{\prime}}_{k^{\prime}}}_{B}
×Γα,(αk)α,(αk)(τ)Γβ,(βl)β,(βl)(τ).\displaystyle\qquad\times\Gamma_{\alpha,(\alpha^{\prime}k^{\prime})}^{\alpha,(\alpha^{\prime}k^{\prime})}(\tau)\Gamma_{\beta,(\beta^{\prime}l^{\prime})}^{\beta,(\beta^{\prime}l^{\prime})}(\tau)^{*}\ . (73)

Using Sections III.1 and 70, we obtain:

a++(τ)\displaystyle a_{++}(\tau) =k|gk|2τsinc2((Ω0+ωk)τ2),\displaystyle=\sum_{k}|g_{k}|^{2}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}+\omega_{k})\tau}{2}\right)\ , (74)
a(τ)\displaystyle a_{--}(\tau) =k|gk|2τsinc2((Ω0ωk)τ2),\displaystyle=\sum_{k}|g_{k}|^{2}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega_{k})\tau}{2}\right)\ , (75)
γ(τ)\displaystyle\gamma(\tau) a(τ)=k|gk|2τsinc2((Ω0ωk)τ2),\displaystyle\equiv a_{--}(\tau)=\sum_{k}|g_{k}|^{2}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega_{k})\tau}{2}\right)\ , (76)

where sinc(x)sin(x)/x\operatorname{sinc}(x)\equiv\sin(x)/x. Introducing the spectral density

J(ω)=k|gk|2δ(ωωk),J(\omega)=\sum_{k}|g_{k}|^{2}\delta(\omega-\omega_{k})\ , (77)

we can write this as

γ(τ)=0𝑑ωJ(ω)τsinc2((Ω0ω)τ2).\gamma(\tau)=\int_{0}^{\infty}d\omega J(\omega)\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)\ . (78)

Let us now define

δ¯(x,y)12πysinc2(xy2),y0.\bar{\delta}(x,y)\equiv\frac{1}{2\pi}y\operatorname{sinc}^{2}\left(\frac{xy}{2}\right)\ ,\quad y\geq 0\ . (79)

This function behaves similarly to the Dirac-δ\delta function:

δ¯(x,y)𝑑x=1\displaystyle\int_{-\infty}^{\infty}\bar{\delta}(x,y)dx=1 (80a)
limyδ¯(x,y)=δ(x),\displaystyle\lim_{y\to\infty}\bar{\delta}(x,y)=\delta(x)\ , (80b)

i.e., it is sharply peaked at x=0x=0, and the peak becomes sharper as yy grows. The peak width is 1/y\sim 1/y. We can thus also write

γ(τ)=2π0𝑑ωJ(ω)δ¯(Ω0ω,τ).\gamma(\tau)=2\pi\int_{0}^{\infty}d\omega J(\omega)\bar{\delta}\left(\Omega_{0}-\omega,\tau\right)\ . (81)

We will show below that this representation allows us to express the RWA-LE as the τ\tau\to\infty limit of the CG-LE and C-LE results, as expected on general grounds [18].

Finally, we obtain the interaction picture Lindblad equation as:

ρ~˙(t)\displaystyle\dot{\tilde{\rho}}(t) =γ(τ)(σρ~(t)σ+12{σ+σ,ρ~(t)}).\displaystyle=\gamma(\tau)\left(\sigma_{-}\tilde{\rho}(t)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\tilde{\rho}(t)\}\right)\ . (82)

Taking matrix elements, we find that the populations and coherences are decoupled. Solving for the excited state population and coherence, we obtain, respectively:

ρ~11(t)\displaystyle\tilde{\rho}_{11}(t) =ρ11(0)eγ(τ)t\displaystyle=\rho_{11}(0)e^{-\gamma(\tau)t} (83a)
ρ~01(t)\displaystyle\tilde{\rho}_{01}(t) =ρ01(0)e12γ(τ)t,\displaystyle=\rho_{01}(0)e^{-\frac{1}{2}\gamma(\tau)t}\ , (83b)

which is to be contrasted with the exact solution given in Eq. 34. The coarse-graining time τ\tau can be chosen to optimize the agreement with the exact solution.

III.1.1 J1(ω)=|g|2δ(ωωc)J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c})

For the impulse spectral density we find, using Eq. 81:

γ(τ)\displaystyle\gamma(\tau) =|g|2τsinc2((Ω0ωc)τ2)\displaystyle=|g|^{2}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega_{c})\tau}{2}\right) (84a)
=2π|g|2δ¯(Ω0ω,τ).\displaystyle=2\pi|g|^{2}\bar{\delta}\left(\Omega_{0}-\omega,\tau\right)\ . (84b)

III.1.2 J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}

For the Ohmic spectral density we find, using Eq. 81:

γ(τ)\displaystyle\gamma(\tau) =η0ωeω/ωcτsinc2((Ω0ω)τ2)𝑑ω\displaystyle=\eta\int_{0}^{\infty}\omega e^{-\omega/\omega_{c}}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)d\omega (85a)
=2πη0ωeω/ωcδ¯(Ω0ω,τ)𝑑ω\displaystyle=2\pi\eta\int_{0}^{\infty}\omega e^{-\omega/\omega_{c}}\bar{\delta}\left(\Omega_{0}-\omega,\tau\right)d\omega (85b)
=ητeΩ0ωc[(1Ω0ωciΩ0τ)Ei(Ω0ωc+iΩ0τ)+\displaystyle=\frac{\eta}{\tau}e^{-\frac{\Omega_{0}}{\omega_{c}}}\left[\left(1-\frac{\Omega_{0}}{\omega_{c}}-i\Omega_{0}\tau\right)\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}+i\Omega_{0}\tau\right)+\right.
(1Ω0ωc+iΩ0τ)Ei(Ω0ωciΩ0τ)\displaystyle\left(1-\frac{\Omega_{0}}{\omega_{c}}+i\Omega_{0}\tau\right)\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}-i\Omega_{0}\tau\right) (85c)
+2(Ω0ωc1)Ei(Ω0ωc)]+2ητ(1cosΩ0τ),\displaystyle\left.+2\left(\frac{\Omega_{0}}{\omega_{c}}-1\right)\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}\right)\right]+\frac{2\eta}{\tau}(1-\cos\Omega_{0}\tau)\ ,

where the last equality is derived in Section B.1.

III.1.3 J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega)

For the triangular spectral density we find, using Eq. 81:

γ(τ)\displaystyle\gamma(\tau) =η0ωcωτsinc2((Ω0ω)τ2)𝑑ω\displaystyle=\eta\int_{0}^{\omega_{c}}\omega\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)d\omega (86a)
=2πη0ωcωδ¯(Ω0ω,τ)𝑑ω\displaystyle=2\pi\eta\int_{0}^{\omega_{c}}\omega\bar{\delta}\left(\Omega_{0}-\omega,\tau\right)d\omega (86b)
=2ητ(ln(ωcΩ01)+Ci(τΩ0)Ci(τ(ωcΩ0))\displaystyle=\frac{2\eta}{\tau}\Bigg{(}\ln\left(\frac{\omega_{c}}{\Omega_{0}}-1\right)+\operatorname{Ci}\left(\tau\Omega_{0}\right)-\operatorname{Ci}\left(\tau\left(\omega_{c}-\Omega_{0}\right)\right)
+ωc(cos(τ(ωcΩ0))1)ωcΩ0cos(τ(ωcΩ0))\displaystyle+\frac{\omega_{c}\left(\cos\left(\tau\left(\omega_{c}-\Omega_{0}\right)\right)-1\right)}{\omega_{c}-\Omega_{0}}-\cos\left(\tau\left(\omega_{c}-\Omega_{0}\right)\right)
+cos(τΩ0))+2ηΩ0(Si(τ(ωcΩ0))+Si(τΩ0)),\displaystyle+\cos\left(\tau\Omega_{0}\right)\Bigg{)}+2\eta\Omega_{0}\left(\operatorname{Si}\left(\tau\left(\omega_{c}-\Omega_{0}\right)\right)+\operatorname{Si}\left(\tau\Omega_{0}\right)\right)\ , (86c)

where the last equality is derived in Section B.2.

III.2 Cumulant Lindblad Equation C-LE

This section briefly reviews an alternative derivation of the Lindblad equation, based on a cumulant expansion [18]. Similarly to the CG-LE, the C-LE approach also uses a coarse-graining time scale that can be optimized to approximate the exact result. Despite using a rather different approach to deriving the Lindblad equation, we will show that for the problem we study in this work, the C-LE ultimately results in identical expressions for the master equation and its parameters (and hence also the solution, of course) as the CG-LE.

We start by writing the system-bath interaction Hamiltonian of Eq. 1 as:

HSB=λαSαBα\displaystyle H_{SB}=\lambda\sum_{\alpha}S_{\alpha}\otimes B_{\alpha} (87)

where SαS_{\alpha} and BαB_{\alpha} are the system and bath operators, respectively (not necessarily Hermitian), and λ\lambda is a dimensionless parameter to be used below for a series expansion, which we eventually set equal to 11. Note that unlike Eq. 47, the SαS_{\alpha} are now not a basis, and the BαB_{\alpha} have dimensions of energy since they include the coupling constants gαg_{\alpha}. Assuming a factorized initial condition ρSB(0)=ρ(0)ρB(0)\rho_{SB}(0)=\rho(0)\otimes\rho_{B}(0), we associate a CPTP map Λλ\Lambda_{\lambda} to the reduced density matrix of Eq. 7:

ρ~(t)=Λλ(t)ρ(0).\displaystyle\tilde{\rho}(t)=\Lambda_{\lambda}(t)\rho(0)\ . (88)

This CPTP map can be related to Eq. 6 by introducing superoperators K(n)K^{(n)}, which collect terms with matching power of λ\lambda:

Λλ(t)=exp(n=1λnK(n)(t))\displaystyle\Lambda_{\lambda}(t)=\exp\left(\sum_{n=1}^{\infty}\lambda^{n}K^{(n)}(t)\right) (89)

This is known as the cumulant expansion. The first-order term is then:

K(1)(t)ρ(0)=i0t𝑑sTrB([H~(s),ρSB(0)]),\displaystyle K^{(1)}(t)\rho(0)=-i\int_{0}^{t}ds\mathrm{Tr}_{B}\left([\tilde{H}(s),\rho_{SB}(0)]\right)\ , (90)

which can be eliminated by shifting the bath operators BαB_{\alpha}, assuming stationarity, i.e., [HB,ρB(0)]=0[H_{B},\rho_{B}(0)]=0 (see below and Appendix C). Moving on to the second order in λ\lambda, we have:

K(2)(t)\displaystyle K^{(2)}(t) ρ(0)=0t𝑑s0s𝑑sTrB[H~(s),[H~(s),ρSB(0)]],\displaystyle\rho(0)=-\int_{0}^{t}ds\int_{0}^{s}ds^{\prime}\mathrm{Tr}_{B}[\tilde{H}(s),[\tilde{H}(s^{\prime}),\rho_{SB}(0)]]\ , (91)

where the double commutator can be rearranged using:

TrB\displaystyle\mathrm{Tr}_{B} [Sα(s)Bα(s),[Sβ(s)Bβ(s),ρ(0)ρB(0)]]\displaystyle[S_{\alpha}(s)\otimes B_{\alpha}(s),[S_{\beta}(s^{\prime})\otimes B_{\beta}(s^{\prime}),\rho(0)\otimes\rho_{B}(0)]]
=\displaystyle= [Sα(s),Sβ(s)ρ(0)]Tr(Bα(s)Bβ(s)ρB)\displaystyle[S_{\alpha}(s),S_{\beta}(s^{\prime})\rho(0)]\mathrm{Tr}(B_{\alpha}(s)B_{\beta}(s^{\prime})\rho_{B})-
[Sα(s),ρ(0)Sβ(s)]Tr(Bβ(s)Bα(s)ρB)\displaystyle[S_{\alpha}(s),\rho(0)S_{\beta}(s^{\prime})]\mathrm{Tr}(B_{\beta}(s^{\prime})B_{\alpha}(s)\rho_{B}) (92)

Introducing the bath correlation function

αβ(s,s)Tr[Bα(s)Bβ(s)ρB],\displaystyle\mathcal{B}_{\alpha\beta}(s,s^{\prime})\equiv\mathrm{Tr}[B_{\alpha}(s)B_{\beta}(s^{\prime})\rho_{B}]\ , (93)

we have:

αβ(s,s)=αβ(s,s).\displaystyle\mathcal{B}_{\alpha\beta}(s,s^{\prime})=\mathcal{B}_{\alpha\beta}^{*}(s^{\prime},s)\ . (94)

To obtain a reduced form of the second-order cumulant in Eq. 91, it is useful to define a new variable that includes the double integration of the bath correlation function as follows:

αβω(t)0t𝑑s0s𝑑seiω(ss)αβ(s,s)\displaystyle\mathcal{B}_{\alpha\beta\omega}(t)\equiv\int_{0}^{t}ds\int_{0}^{s}ds^{\prime}e^{i\omega(s-s^{\prime})}\mathcal{B}_{\alpha\beta}(s,s^{\prime}) (95)

We introduce two more variables that will be associated with the Lamb shift and the decoherence rate:

Qαβω(t)12i(αβω(t)αβω(t))=Im(αβω(t)),\displaystyle Q_{\alpha\beta\omega}(t)\equiv\frac{1}{2i}\left(\mathcal{B}_{\alpha\beta\omega}(t)-\mathcal{B}^{*}_{\alpha\beta\omega}(t)\right)=\imaginary(\mathcal{B}_{\alpha\beta\omega}(t))\ , (96)

and

bαβω(t)\displaystyle b_{\alpha\beta\omega}(t) 0t𝑑s0t𝑑seiω(ss)αβ(s,s)\displaystyle\equiv\int_{0}^{t}ds\int_{0}^{t}ds^{\prime}e^{i\omega(s-s^{\prime})}\mathcal{B}_{\alpha\beta}(s,s^{\prime}) (97a)
=αβω(t)+αβω(t)\displaystyle=\mathcal{B}_{\alpha\beta\omega}(t)+\mathcal{B}^{*}_{\alpha\beta\omega}(t) (97b)
=2Re(αβω(t)),\displaystyle=2\real(\mathcal{B}_{\alpha\beta\omega}(t))\ , (97c)

where the equality in the second line is shown in Appendix D.

Explicitly, in this section, α{+,}\alpha\in\{+,-\}, and as in Eq. 11, in the interaction picture the system operators are S±(t)=σ±e±iΩ0tS_{\pm}(t)=\sigma_{\pm}e^{\pm i\Omega_{0}t} and the bath operators are B+(t)=kgkeiωktbk,B(t)=kgkeiωktbkB_{+}(t)=\sum_{k}g_{k}e^{-i\omega_{k}t}b_{k},~{}B_{-}(t)=\sum_{k}g_{k}^{*}e^{i\omega_{k}t}b_{k}^{\dagger}. The initial bath state ρB=|vv|\rho_{B}=|{v}\rangle\!\langle v| and the commutation rules Section III.1 give us just one non-zero bath correlation function:

+(s,s)\displaystyle\mathcal{B}_{+-}(s,s^{\prime}) =Tr[ρBB+(s)B(s)]\displaystyle=\mathrm{Tr}[\rho_{B}B_{+}(s)B_{-}(s^{\prime})] (98a)
=l,lglglei(ωlsωls)Tr[ρB(0)blbl]\displaystyle=\sum_{l,l^{\prime}}g_{l}g_{l^{\prime}}^{*}e^{-i(\omega_{l}s-\omega_{l^{\prime}}s^{\prime})}\mathrm{Tr}[\rho_{B}(0)b_{l}b_{l^{\prime}}^{\dagger}] (98b)
+(s,s)\displaystyle\mathcal{B}_{-+}(s,s^{\prime}) =++(s,s)=(s,s)=0.\displaystyle=\mathcal{B}_{++}(s,s^{\prime})=\mathcal{B}_{--}(s,s^{\prime})=0\ . (98c)

The double commutator Section III.2 can now be simplified as follows:

α,βTrB\displaystyle\sum_{\alpha,\beta}\mathrm{Tr}_{B} [Sα(s)Bα(s),[Sβ(s)Bβ(s),ρ(0)ρB(0)]]\displaystyle[S_{\alpha}(s)\otimes B_{\alpha}(s),[S_{\beta}(s^{\prime})\otimes B_{\beta}(s^{\prime}),\rho(0)\otimes\rho_{B}(0)]]
=eiΩ0(ss)[σ+,σρ(0)]+(s,s)\displaystyle=e^{i\Omega_{0}(s-s^{\prime})}[\sigma_{+},\sigma_{-}\rho(0)]\mathcal{B}_{+-}(s,s^{\prime})
eiΩ0(ss)[σ,ρ(0)σ+]+(s,s).\displaystyle\quad-e^{-i\Omega_{0}(s-s^{\prime})}[\sigma_{-},\rho(0)\sigma_{+}]\mathcal{B}_{+-}(s^{\prime},s)\ . (99)

Using Eq. 94, we have:

K(2)(t)ρ(0)\displaystyle K^{(2)}(t)\rho(0) (100a)
=+,Ω0(t)[σ+,σρ(0)]++,Ω0(t)[σ,ρ(0)σ+]\displaystyle\quad=-\mathcal{B}_{+-,\Omega_{0}}(t)[\sigma_{+},\sigma_{-}\rho(0)]+\mathcal{B}_{+-,\Omega_{0}}^{*}(t)[\sigma_{-},\rho(0)\sigma_{+}]
=iIm(+,Ω0(t))[σ+σ,ρ(0)]\displaystyle\quad=-i\imaginary(\mathcal{B}_{+-,\Omega_{0}}(t))[\sigma_{+}\sigma_{-},\rho(0)] (100b)
+2Re(+,Ω0(t))(σρ(0)σ+12{σ+σ,ρ(0)}).\displaystyle\qquad+2\real(\mathcal{B}_{+-,\Omega_{0}}(t))\left(\sigma_{-}\rho(0)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\rho(0)\}\right)\ .

Hence, the second-order cumulant takes the form:

K(2)(t)ρ(0)=iQ+,Ω0(t)[σ+σ,ρ(0)]\displaystyle K^{(2)}(t)\rho(0)=-iQ_{+-,\Omega_{0}}(t)[\sigma_{+}\sigma_{-},\rho(0)]
+b+,Ω0(t)(σρ(0)σ+12{σ+σ,ρ(0)}).\displaystyle\quad+b_{+-,\Omega_{0}}(t)\left(\sigma_{-}\rho(0)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\rho(0)\}\right)\ . (101)

Hence, the state in the interaction picture after the CP map in Eq. 88 using a truncation up to the second order in the cumulant expansion, is:

ρ~(t)\displaystyle\tilde{\rho}(t) =Λλ(t)ρ(0)exp(λ2K(2)(t))ρ(0)\displaystyle=\Lambda_{\lambda}(t)\rho(0)\approx\exp\left(\lambda^{2}K^{(2)}(t)\right)\rho(0) (102a)
=ρ(0)iλ2Q+,Ω0(t)[σ+σ,ρ(t)]\displaystyle={\rho}(0)-i\lambda^{2}Q_{+-,\Omega_{0}}(t)[\sigma_{+}\sigma_{-},\rho(t)] (102b)
+λ2b+,Ω0(t)(σρ(t)σ+12{σ+σ,ρ(t)}).\displaystyle\quad+\lambda^{2}b_{+-,\Omega_{0}}(t)\left(\sigma_{-}\rho(t)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\rho(t)\}\right)\ .

Now we use the coarse-graining method by averaging over the coarse-graining timescale τ\tau as in Eq. 60:

b˙+,Ω0(t)0\displaystyle\langle\dot{b}_{+-,\Omega_{0}}(t)\rangle_{0} =b+,Ω0(τ)τ\displaystyle=\frac{b_{+-,\Omega_{0}}(\tau)}{\tau} (103a)
Q˙+,Ω0(t)0\displaystyle\langle\dot{Q}_{+-,\Omega_{0}}(t)\rangle_{0} =Q+,Ω0(τ)τ,\displaystyle=\frac{Q_{+-,\Omega_{0}}(\tau)}{\tau}\ , (103b)

which, when applied to Eq. 102b, yields:

ρ~˙(t)0=iλ2Q˙+,Ω0(t)0[σ+σ,ρ(0)]\displaystyle\langle\dot{\tilde{\rho}}(t)\rangle_{0}=-i\lambda^{2}\langle\dot{Q}_{+-,\Omega_{0}}(t)\rangle_{0}[\sigma_{+}\sigma_{-},\rho(0)] (104)
+λ2b˙+,Ω0(t)0(σρ(0)σ+12{σ+σ,ρ(0)}),\displaystyle+\lambda^{2}\langle\dot{b}_{+-,\Omega_{0}}(t)\rangle_{0}\left(\sigma_{-}\rho(0)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\rho(0)\}\right)\ ,

where we also used Eq. 66. Let us now define

𝒮(τ)\displaystyle\mathcal{S}(\tau) 2Q˙+,Ω0(t)0=2Im(+,Ω0(τ))τ\displaystyle\equiv 2\langle\dot{Q}_{+-,\Omega_{0}}(t)\rangle_{0}=\frac{2\imaginary(\mathcal{B}_{+-,\Omega_{0}}(\tau))}{\tau} (105a)
γ(τ)\displaystyle\gamma(\tau) b˙+,Ω0(t)0=2Re(+,Ω0(τ))τ,\displaystyle\equiv\langle\dot{b}_{+-,\Omega_{0}}(t)\rangle_{0}=\frac{2\real(\mathcal{B}_{+-,\Omega_{0}}(\tau))}{\tau}\ , (105b)

where we used Eqs. 96 and 97c.

For a general ρB(0)\rho_{B}(0) obtained by tracing out the system from Eq. 14 we find that [ρB(0),HB]0[\rho_{B}(0),H_{B}]\neq 0 (as shown in Appendix E), which means that the bath correlation function αβ(s,s)\mathcal{B}_{\alpha\beta}(s,s^{\prime}) is not stationary. However, it is for the vacuum bath state ρB(0)=|vv|\rho_{B}(0)=|{v}\rangle\!\langle v| which we assume to be the case throughout, so we can write

+(s,s)=+(ss)=0𝑑ωJ(ω)eiω(ss).\mathcal{B}_{+-}(s,s^{\prime})=\mathcal{B}_{+-}(s-s^{\prime})=\int_{0}^{\infty}d\omega J(\omega)e^{-i\omega(s-s^{\prime})}\ . (106)

We then obtain, using Eq. 96:

S(τ)\displaystyle S(\tau) =1τIm0τ𝑑s0s𝑑seiω(ss)0𝑑ωJ(ω)eiω(ss)\displaystyle=\frac{1}{\tau}\imaginary\int_{0}^{\tau}ds\int_{0}^{s}ds^{\prime}e^{i\omega(s-s^{\prime})}\int_{0}^{\infty}d\omega J(\omega)e^{-i\omega(s-s^{\prime})} (107a)
=τ2Im0𝑑ωJ(ω)=0,\displaystyle=\frac{\tau}{2}\imaginary\int_{0}^{\infty}d\omega J(\omega)=0\ , (107b)

since the spectral density is real. I.e., just like in the CG-LE case [Eq. 72], the Lamb shift vanishes.

For the decay rate we now have, using Eq. 97b:

γ(τ)\displaystyle\gamma(\tau) =1τ0𝑑ωJ(ω)0τ𝑑s0τ𝑑sei(Ω0ω)(ss)\displaystyle=\frac{1}{\tau}\int_{0}^{\infty}d\omega J(\omega)\int_{0}^{\tau}ds\int_{0}^{\tau}ds^{\prime}e^{i(\Omega_{0}-\omega)(s-s^{\prime})} (108a)
=0𝑑ωJ(ω)τsinc2((Ω0ω)τ2),\displaystyle=\int_{0}^{\infty}d\omega J(\omega)\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)\ , (108b)

which is identical to the CG-LE result, Eq. 78.

Moreover, similar to how we arrived at Eq. 61, assuming Markovianity in the sense that Eq. 104 can be extended to any interval [t,t+τ][t,t+\tau] we again arrive at the Lindblad equation in the interaction picture, after replacing ρ~˙(t)0ρ~˙(t)\langle\dot{\tilde{\rho}}(t)\rangle_{0}\mapsto\dot{\tilde{\rho}}(t), and setting λ=1\lambda=1. The form of this Lindblad equation is identical to Eq. 82. In particular, both the excited state population and the coherence are the same as in Eq. 83. Thus, the end results of the C-LE and CG-LE are identical for the model considered in this work.

III.3 Rotating Wave Approximation Lindblad Equation (RWA-LE)

The rotating wave approximation (RWA) drops the non-secular (off-diagonal) frequency terms which appear in the C-LE [see, e.g., Eq. 98b]. This approximation is based on the idea that the terms with ωω\omega\neq\omega^{\prime} are rapidly oscillating if t|ωω|1t\gg|\omega-\omega^{\prime}|^{-1}, which thus (roughly) average to zero. Since we will assume that tτBt\gg\tau_{B}, where τB\tau_{B} is the bath correlation time (the time over which the bath correlation function decays), the former assumption is consistent provided we also assume that the Bohr frequency differences satisfy minωω|ωω|>1/τB\min_{\omega\neq\omega^{\prime}}|\omega-\omega^{\prime}|>1/\tau_{B}. Combining this with the weak coupling assumption, we obtain:

g1τB<minωω|ωω|.\displaystyle g\ll\frac{1}{\tau_{B}}<\min_{\omega\neq\omega^{\prime}}|\omega-\omega^{\prime}|\ . (109)

By considering the weak coupling limit, taking the bath correlation timescale as the inverse of the cutoff frequency, and considering the Bohr frequencies {0,±Ω0}\{0,\pm\Omega_{0}\}, Eq. 109 becomes:

η1<Ω0/ωc.\displaystyle\eta\ll 1<\Omega_{0}/\omega_{c}\ . (110)

Furthermore, the Born approximation states that for a sufficiently large bath, the composite state factorizes:

ρ~SB(t)ρ~(t)ρB.\displaystyle\tilde{\rho}_{SB}(t)\approx\tilde{\rho}(t)\otimes\rho_{B}. (111)

Thus, up to second order in the Dyson series, the system state evolves according to:

ρ~˙\displaystyle\dot{\tilde{\rho}} =TrB[H~(t),0t𝑑τ[H~(tτ),ρ~SB(tτ)]]\displaystyle=-\mathrm{Tr}_{B}[\tilde{H}(t),\int_{0}^{t}d\tau[\tilde{H}(t-\tau),\tilde{\rho}_{SB}(t-\tau)]] (112a)
=α,βTrB[Aα(t)Bα(t),0tdτ[Aβ(tτ)\displaystyle=-\sum_{\alpha,\beta}\mathrm{Tr}_{B}[A_{\alpha}(t)\otimes B_{\alpha}(t),\int_{0}^{t}d\tau[A_{\beta}(t-\tau)
Bβ(tτ),ρ~(tτ)ρB],\displaystyle\qquad\otimes B_{\beta}(t-\tau),\tilde{\rho}(t-\tau)\otimes\rho_{B}], (112b)

where in our case α,β{+,}\alpha,\beta\in\{+,-\}. It is useful to define the stationary (single-variable) bath correlation, a special case of Eq. 93:

+(t,tτ)=0𝑑ωJ(ω)eiωτ+(τ).\displaystyle\mathcal{B}_{+-}(t,t-\tau)=\int_{0}^{\infty}d\omega J(\omega)e^{-i\omega\tau}\equiv\mathcal{B}_{+-}(\tau)\ . (113)

Now if we assume that tτBt\gg\tau_{B}, then ρ~(tτ)ρ~(t)\tilde{\rho}(t-\tau)\approx\tilde{\rho}(t). We discuss the limitations of this approximation in Appendix F, where we show that it can lead to an unbounded error.

Expanding the double commutator in terms of the bath correlation function, we obtain:

α,βTrB[Sα(t)Bα,[Sβ(tτ)Bβ(tτ),ρ~(t)ρB]]\displaystyle\sum_{\alpha,\beta}\mathrm{Tr}_{B}[S_{\alpha}(t)\otimes B_{\alpha},[S_{\beta}(t-\tau)\otimes B_{\beta}(t-\tau),\tilde{\rho}(t)\otimes\rho_{B}]]
=[S+(t),S(tτ)ρ~(t)]+(τ)\displaystyle\quad=[S_{+}(t),S_{-}(t-\tau)\tilde{\rho}(t)]\mathcal{B}_{+-}(\tau)
[S(t),ρ~(t)S+(tτ)]+(τ).\displaystyle\quad\quad-[S_{-}(t),\tilde{\rho}(t)S_{+}(t-\tau)]\mathcal{B}_{+-}(-\tau)\ . (114)

Consequently, substituting Section III.3 back into Eq. 112, we can write:

ρ~˙(t)=0t𝑑τ+(τ)eiΩ0τ[σ+,σρ~(t)]\displaystyle\dot{\tilde{\rho}}(t)=-\int_{0}^{t}d\tau\mathcal{B}_{+-}(\tau)e^{i\Omega_{0}\tau}[\sigma_{+},\sigma_{-}\tilde{\rho}(t)]
+0t𝑑τ+(τ)eiΩ0τ[σ,ρ~(t)σ+].\displaystyle\quad+\int_{0}^{t}d\tau\mathcal{B}_{+-}(-\tau)e^{-i\Omega_{0}\tau}[\sigma_{-},\tilde{\rho}(t)\sigma_{+}]\ . (115)

To arrive at a Lindblad form, we complete the Markovian approximation by setting the upper limit of the integral to be \infty. This is justified since the bath correlation decays rapidly to zero for t1/ωct\gg 1/\omega_{c}. Now, let

Γαβ(ω)\displaystyle\Gamma_{\alpha\beta}(\omega) 0𝑑ταβ(τ)eiωτ\displaystyle\equiv\int_{0}^{\infty}d\tau\mathcal{B}_{\alpha\beta}(\tau)e^{i\omega\tau} (116a)
=0𝑑ωJ(ω)0𝑑τei(ωω)τ\displaystyle=\int_{0}^{\infty}d\omega^{\prime}J(\omega^{\prime})\int_{0}^{\infty}d\tau e^{i(\omega-\omega^{\prime})\tau} (116b)
=πJ(ω)+i0𝑑ωJ(ω)𝒫(1ωω),\displaystyle=\pi J(\omega)+i\int_{0}^{\infty}d\omega^{\prime}\ J(\omega^{\prime})\mathcal{P}\left(\frac{1}{\omega-\omega^{\prime}}\right)\ , (116c)

where we used the identity

0𝑑τeixτ=πδ(x)+i𝒫(1x),\int_{0}^{\infty}d\tau e^{ix\tau}=\pi\delta(x)+i\mathcal{P}\left(\frac{1}{x}\right)\ , (117)

and where the Cauchy principal value is defined as

𝒫(1x)[f]=limϵ0ϵϵf(x)x𝑑x,\mathcal{P}\left(\frac{1}{x}\right)[f]=\lim_{\epsilon\to 0}\int_{-\epsilon}^{\epsilon}\frac{f(x)}{x}dx\ , (118)

for smooth functions ff with compact support on the real line \mathbb{R}.

We show in Appendix G that taking the complex conjugate of Γαβ\Gamma_{\alpha\beta} yields:

Γ±(ω)=0𝑑τ±(τ)eiωτ.\displaystyle\Gamma^{*}_{{\pm\mp}}(\omega)=\int_{0}^{\infty}d\tau\mathcal{B}_{{\pm\mp}}(-\tau)e^{-i\omega\tau}\ . (119)

This simplifies Section III.3 into Lindblad form:

ρ~˙(t)\displaystyle\dot{\tilde{\rho}}(t) =Γ+(Ω0)[σ+,σρ~]+Γ+(Ω0)[σ,ρ~σ+]\displaystyle=-\Gamma_{+-}(\Omega_{0})[\sigma_{+},\sigma_{-}\tilde{\rho}]+\Gamma_{+-}^{*}(\Omega_{0})[\sigma_{-},\tilde{\rho}\sigma_{+}] (120a)
=iIm[Γ+(Ω0)][σ+σ,ρ(t)]\displaystyle=-i\imaginary[\Gamma_{+-}(\Omega_{0})][\sigma_{+}\sigma_{-},\rho(t)] (120b)
+2Re[Γ+(Ω0)](σρ(t)σ+12{σ+σ,ρ(t)}).\displaystyle\quad+2\real[\Gamma_{+-}(\Omega_{0})]\left(\sigma_{-}\rho(t)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\rho(t)\}\right)\ .

This result has the same form as the exact Section II.3, but with a time-independent Lamb shift and decay rate:

𝒮\displaystyle\mathcal{S} =2Im[Γ+(Ω0)]=20𝑑ωJ(ω)𝒫(1Ω0ω)\displaystyle=2\imaginary[\Gamma_{+-}(\Omega_{0})]=2\int_{0}^{\infty}d\omega^{\prime}\ J(\omega^{\prime})\mathcal{P}\left(\frac{1}{\Omega_{0}-\omega^{\prime}}\right) (121a)
γ\displaystyle\gamma =2Re[Γ+(Ω0)]=2πJ(Ω0).\displaystyle=2\real[\Gamma_{+-}(\Omega_{0})]=2\pi J(\Omega_{0})\ . (121b)

This last result is consistent with the finding that the RWA-LE is the τ\tau\to\infty limit of the C-LE [18], since it follows from Eqs. 80b and 81 that

limτγ(τ)=2π0𝑑ωJ(ω)δ(Ω0ω)=2πJ(Ω0).\lim_{\tau\to\infty}\gamma(\tau)=2\pi\int_{0}^{\infty}d\omega J(\omega)\delta(\Omega_{0}-\omega)=2\pi J(\Omega_{0})\ . (122)

The population and coherence are given by the τ\tau\to\infty limit of Eq. 83, i.e.,

ρ~11(t)\displaystyle\tilde{\rho}_{11}(t) =ρ11(0)eγt\displaystyle=\rho_{11}(0)e^{-\gamma t} (123a)
ρ~01(t)\displaystyle\tilde{\rho}_{01}(t) =ρ01(0)e12γt,\displaystyle=\rho_{01}(0)e^{-\frac{1}{2}\gamma t}\ , (123b)

We are now ready to present the results for the three spectral densities.

III.3.1 J1(ω)=|g|2δ(ωωc)J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c})

For the impulse spectral density we find, using Eq. 121:

𝒮\displaystyle\mathcal{S} =2|g|2𝒫(1Ω0ωc)\displaystyle=2|g|^{2}\mathcal{P}\left(\frac{1}{\Omega_{0}-\omega_{c}}\right) (124a)
γ\displaystyle\gamma =2π|g|2δ(Ω0ωc).\displaystyle=2\pi|g|^{2}\delta(\Omega_{0}-\omega_{c})\ . (124b)

This means that the decay rate either vanishes or is singular at Ω0=ωc\Omega_{0}=\omega_{c}. Hence, the RWA-LE is unsuitable for describing the model with this spectral density.

III.3.2 J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}

We can immediately write down the decay rate as γ=2πJ2(Ω0)\gamma=2\pi J_{2}(\Omega_{0}). However, the Cauchy principal value complicates the calculation of the Lamb shift, so we use a direct method instead.

For the Ohmic spectral density, the bath correlation in Eq. 113 takes the form:

+(τ)=η0ωeω/ωceiωτ𝑑ω=ηωc2(1+iωcτ)2.\displaystyle\mathcal{B}_{+-}(\tau)=\eta\int_{0}^{\infty}\omega e^{-\omega/\omega_{c}}e^{-i\omega\tau}d\omega=\frac{\eta\omega_{c}^{2}}{(1+i\omega_{c}\tau)^{2}}\ . (125)

The one-sided Fourier integral in Eq. 116a becomes:

Γ+(Ω0)=ηωc0d(ωcτ)eiΩ0τ(1+iωcτ)2\displaystyle\Gamma_{+-}(\Omega_{0})=\eta\omega_{c}\int_{0}^{\infty}d(\omega_{c}\tau)\frac{e^{i\Omega_{0}\tau}}{(1+i\omega_{c}\tau)^{2}} (126a)
=iηωc+ηΩ0eΩ0/ωc(π+iEi(Ω0/ωc)),\displaystyle\quad=-i\eta\omega_{c}+\eta\Omega_{0}e^{-\Omega_{0}/\omega_{c}}(\pi+i\operatorname{Ei}\left(\Omega_{0}/\omega_{c}\right))\ , (126b)

where we derive the second equality in Section H.1.

Thus, the Lamb shift and the decay rate are:

𝒮\displaystyle\mathcal{S} =2J2(Ω0)Ei(Ω0ωc)2ηωc\displaystyle=2J_{2}(\Omega_{0})\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}\right)-2\eta\omega_{c} (127a)
γ\displaystyle\gamma =2πJ2(Ω0).\displaystyle=2\pi J_{2}(\Omega_{0})\ . (127b)

III.3.3 J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega)

We can once more immediately write down the decay rate as γ=2πJ3(Ω0)\gamma=2\pi J_{3}(\Omega_{0}), but a direct calculation is again advantageous for arriving at the form of the Lamb shift.

For the triangular spectral density, the bath correlation in Eq. 113 takes the form:

+(τ)=η0ωcωeiωτ𝑑ω=ηeiωcτ(1+iωcτ)1τ2\displaystyle\mathcal{B}_{+-}(\tau)=\eta\int_{0}^{\omega_{c}}\omega e^{-i\omega\tau}d\omega=\eta\frac{e^{-i\omega_{c}\tau}(1+i\omega_{c}\tau)-1}{\tau^{2}} (128)

The one-sided Fourier integral in Eq. 116a becomes:

Γ+(Ω0)=ηωc0d(ωcτ)eiωcτ(1+iωcτ)1(ωcτ)2eiΩ0τ\displaystyle\Gamma_{+-}(\Omega_{0})=\eta\omega_{c}\int_{0}^{\infty}d(\omega_{c}\tau)\frac{e^{-i\omega_{c}\tau}(1+i\omega_{c}\tau)-1}{(\omega_{c}\tau)^{2}}e^{i\Omega_{0}\tau} (129)

Thus, as we derive in Section H.2, the Lamb shift and the decay rate are:

𝒮\displaystyle\mathcal{S} =2Ω0ln|ωcΩ01|2ηωc\displaystyle=-2\Omega_{0}\ln\left|\frac{\omega_{c}}{\Omega_{0}}-1\right|-2\eta\omega_{c} (130a)
γ\displaystyle\gamma =2πJ3(Ω0).\displaystyle=2\pi J_{3}(\Omega_{0})\ . (130b)

Note that Eq. 110 requires Ω0>ωc\Omega_{0}>\omega_{c}, but in this case it follows from Eq. 130b that γ\gamma vanishes. This breakdown of the validity conditions of the Markov approximation, along with the issue of the potentially unbounded approximation error discussed in Appendix F, highlights that the RWA-LE has limited validity for the model we study here. Our simulation results reinforce these conclusions, as shown in Section IV.

III.4 TCL

In this section we briefly review the time-convolutionless formalism, closely following the presentation of [2] while adding a few pertinent details.

We can rewrite the Liouville equation in Eq. 3 as:

dρ~SBdtλρ~SB,[H~SB,],\displaystyle\frac{d\tilde{\rho}_{SB}}{dt}\equiv\lambda\mathcal{L}\tilde{\rho}_{SB}\ \ ,\qquad\mathcal{L}\equiv[\tilde{H}_{SB},\cdot]\ , (131)

where we introduce the Feshbach projection superoperator 𝒫\mathcal{P} via

𝒫ρSBTrB(ρ)SBρB,\displaystyle\mathcal{P}\rho_{SB}\equiv\mathrm{Tr}_{B}(\rho)_{SB}\otimes\rho_{B}\ , (132)

and its orthogonal complement 𝒬=𝒫\mathcal{Q}=\mathcal{I}-\mathcal{P}. For an arbitrary operator AA, we define:

A^𝒫A,A¯𝒬A,\displaystyle\hat{A}\equiv\mathcal{P}A\ ,\qquad\bar{A}\equiv\mathcal{Q}A\ , (133)

which leads via Eq. 131 to:

tρ^SB\displaystyle\partial_{t}\hat{\rho}_{SB} =λ^ρ^SB+λ^ρ¯SB\displaystyle=\lambda\hat{\mathcal{L}}\hat{\rho}_{SB}+\lambda\hat{\mathcal{L}}\bar{\rho}_{SB} (134a)
tρ¯SB\displaystyle\partial_{t}\bar{\rho}_{SB} =λ¯ρ^SB+λ¯ρ¯SB.\displaystyle=\lambda\bar{\mathcal{L}}\hat{\rho}_{SB}+\lambda\bar{\mathcal{L}}\bar{\rho}_{SB}\ . (134b)

The second equation has a solution given by:

ρ¯SB(t)=𝒢(t,0)ρ¯SB(0)+λ0t𝒢(t,t)¯(t)ρ¯SB(t)𝑑t,\displaystyle\bar{\rho}_{SB}(t)=\mathcal{G}(t,0)\bar{\rho}_{SB}(0)+\lambda\int_{0}^{t}\mathcal{G}(t,t^{\prime})\bar{\mathcal{L}}(t^{\prime})\bar{\rho}_{SB}(t^{\prime})dt^{\prime}\ , (135)

where:

𝒢(t,0)T+eλ0t¯(t)𝑑t.\displaystyle\mathcal{G}(t,0)\equiv T_{+}e^{\lambda\int_{0}^{t}\bar{\mathcal{L}}(t^{\prime})dt^{\prime}}\ . (136)

From Eq. 131, we obtain:

ρ~SB(t)=𝒰+(t,t)ρ~SB(t),\displaystyle\tilde{\rho}_{SB}(t)=\mathcal{U}_{+}(t,t^{\prime})\tilde{\rho}_{SB}(t^{\prime})\ , (137)

where

𝒰+(t,t)=T+eλtt¯(s)𝑑s.\displaystyle\mathcal{U}_{+}(t,t^{\prime})=T_{+}e^{\lambda\int_{t^{\prime}}^{t}\bar{\mathcal{L}}(s)ds}\ . (138)

We can back-propagate the system state as:

ρ~SB(t)=𝒰(t,t)ρ~SB(t),\displaystyle\tilde{\rho}_{SB}(t^{\prime})=\mathcal{U}_{-}(t^{\prime},t)\tilde{\rho}_{SB}(t^{\prime})\ , (139)

where 𝒰(t,t)=[𝒰+(t,t)]1\mathcal{U}_{-}(t^{\prime},t)=[\mathcal{U}_{+}(t,t^{\prime})]^{-1}. When we substitute Eq. 139 back into Eq. 135, it yields:

ρ¯SB(t)\displaystyle\bar{\rho}_{SB}(t) =𝒢(t,0)ρ¯SB(0)+Σ(t)ρSB(t)\displaystyle=\mathcal{G}(t,0)\bar{\rho}_{SB}(0)+\Sigma(t)\rho_{SB}(t) (140a)
Σ(t)\displaystyle\Sigma(t) λ0t𝒢(t,t)¯(t)𝒰^(t,t)𝑑t.\displaystyle\equiv\lambda\int_{0}^{t}\mathcal{G}(t,t^{\prime})\bar{\mathcal{L}}(t^{\prime})\hat{\mathcal{U}}_{-}(t^{\prime},t)dt^{\prime}\ . (140b)

Assuming that Σ\mathcal{I}-\Sigma is invertible and using Eq. 134, we arrive at:

tρ^SB(t)\displaystyle\partial_{t}\hat{\rho}_{SB}(t) =𝒥(t)ρ¯SB(0)+𝒦(t)ρ^SB(t)\displaystyle=\mathcal{J}(t)\bar{\rho}_{SB}(0)+\mathcal{K}(t)\hat{\rho}_{SB}(t) (141a)
𝒥(t)\displaystyle\mathcal{J}(t) λ^(t)[Σ]1𝒢(t,0)𝒬\displaystyle\equiv\lambda\hat{\mathcal{L}}(t)[\mathcal{I}-\Sigma]^{-1}\mathcal{G}(t,0)\mathcal{Q} (141b)
𝒦(t)\displaystyle\mathcal{K}(t) λ^[Σ]1𝒫.\displaystyle\equiv\lambda\hat{\mathcal{L}}[\mathcal{I}-\Sigma]^{-1}\mathcal{P}\ . (141c)

If the inhomogeneity 𝒥(t)\mathcal{J}(t) vanishes (as is the case for a factorized initial system-bath state), then the resulting master equation is time-local and known as the time-convolutionless (TCL) master equation. By expressing Σ\mathcal{I}-\Sigma as a geometric series, we obtain:

[Σ]1=n=0Σn(t).\displaystyle[\mathcal{I}-\Sigma]^{-1}=\sum_{n=0}^{\infty}\Sigma^{n}(t)\ . (142)

This allows us to write the TCL generator as an expansion in powers of λ\lambda:

𝒦(t)=λ^(t)(n=0Σn(t))𝒫=n=1λn𝒦n(t).\displaystyle\mathcal{K}(t)=\lambda\hat{\mathcal{L}}(t)\left(\sum_{n=0}^{\infty}\Sigma^{n}(t)\right)\mathcal{P}=\sum_{n=1}^{\infty}\lambda^{n}\mathcal{K}_{n}(t)\ . (143)

For Gaussian baths, like the one considered here, the odd-order terms vanish, i.e., 𝒦2n+1=0\mathcal{K}_{2n+1}=0, which follows from the vanishing multi-time bath correlation function utilizing Wick’s theorem [25, 26]. In general, the nn’th order term is given by [27, 28]:

𝒦n(t)=0t𝑑t10t𝑑t20tn2𝑑tn1(t)(t1)(tn1)oc,\displaystyle\mathcal{K}_{n}(t)=\int_{0}^{t}\!\!\!dt_{1}\int_{0}^{t}\!\!\!dt_{2}\cdots\int_{0}^{t_{n-2}}dt_{n-1}\langle\mathcal{L}(t)\mathcal{L}(t_{1})...\mathcal{L}(t_{n-1})\rangle_{\text{oc}}\ , (144)

where the ordered cumulants are defined as:

(t)\displaystyle\langle\mathcal{L}(t) (t1)(tn1)oc=\displaystyle\mathcal{L}(t_{1})...\mathcal{L}(t_{n-1})\rangle_{\text{oc}}=
(1)q𝒫(t)(ti)𝒫(tj)(tm)𝒫,\displaystyle\sum(-1)^{q}\mathcal{P}\mathcal{L}(t)\cdots\mathcal{L}(t_{i})\mathcal{P}\mathcal{L}(t_{j})\cdots\mathcal{L}(t_{m})\mathcal{P}\ , (145)

where the sum is over all possible arrangements of qq 𝒫\mathcal{P}’s and nn \mathcal{L}’s such that there is at least one \mathcal{L} between 𝒫\mathcal{P}’s and there is a time ordering ttitjtmt\geq...\geq t_{i}\geq t_{j}...\geq t_{m}.

Since Eq. 31 is already time-local, we can relate it to the TCL generator 𝒦(t)\mathcal{K}(t) via [16]:

𝒦S(t)ρ(t)=TrB[𝒦(t)(ρ(t)ρB)].\displaystyle\mathcal{K}_{S}(t)\rho(t)=\mathrm{Tr}_{B}[\mathcal{K}(t)(\rho(t)\otimes\rho_{B})]\ . (146)

This leads exactly to the ansatz given in Section II.3, i.e., the TCL master equation is given by

ρ~˙\displaystyle\dot{\tilde{\rho}} =𝒦S(t)ρ~=i2𝒮(t)[σ+σ,ρ~(t)]\displaystyle=\mathcal{K}_{S}(t)\tilde{\rho}=-\frac{i}{2}\mathcal{S}(t)[\sigma_{+}\sigma_{-},\tilde{\rho}(t)]
+γ(t)(σρ~(t)σ+12{σ+σ,ρ~(t)}).\displaystyle\qquad+\gamma(t)\left(\sigma_{-}\tilde{\rho}(t)\sigma_{+}-\frac{1}{2}\{\sigma_{+}\sigma_{-},\tilde{\rho}(t)\}\right)\ . (147)

The solution is given by Eq. 34, to the desired order in perturbation theory, i.e.:

ρ~11(t)\displaystyle\tilde{\rho}_{11}(t) =ρ11(0)exp(0tγ(n)(t)𝑑t)\displaystyle=\rho_{11}(0)\exp{-\int_{0}^{t}\gamma^{(n)}(t^{\prime})dt^{\prime}} (148a)
ρ~01(t)\displaystyle\tilde{\rho}_{01}(t) =ρ01(0)exp(120t(i𝒮(n)(t)γ(n)(t))𝑑t),\displaystyle=\rho_{01}(0)\exp{\frac{1}{2}\int_{0}^{t}\left(i\mathcal{S}^{(n)}(t^{\prime})-\gamma^{(n)}(t^{\prime})\right)dt^{\prime}}\ , (148b)

where n=2n=2 for TCL2, n=4n=4 for TCL4, etc.

To find the perturbative expansion of the Lamb shift and the decay rate, we observe that σ+\sigma_{+} is an eigenoperator of the generator:

𝒦S(t)σ+=12(γ(t)+iS(t))σ+.\displaystyle\mathcal{K}_{S}(t)\sigma_{+}=-\frac{1}{2}(\gamma(t)+iS(t))\sigma_{+}\ . (149)

Using the superoperator (t)=i[HSB(t),]\mathcal{L}(t)=-i[H_{SB}(t),\cdot], we can verify [29] that σ+ρB\sigma_{+}\otimes\rho_{B} is an eigenoperator of (t)(t1)\mathcal{L}(t)\mathcal{L}(t_{1}) with eigenvalue f(tt1)-f(t-t_{1}), where f(t)f(t) is defined in Eq. 22. Substituting Eq. 143 into Eq. 149, we obtain:

𝒦S(t)σ+=n=1(λ2)nt0t𝑑t1t0t1𝑑t2t0t2n2𝑑t2n1\displaystyle\mathcal{K}_{S}(t)\sigma_{+}=\sum_{n=1}^{\infty}(-\lambda^{2})^{n}\int_{t_{0}}^{t}dt_{1}\int_{t_{0}}^{t_{1}}dt_{2}\cdots\int_{t_{0}}^{t_{2n-2}}dt_{2n-1}
f(tt1)f(t2t3)f(t2n2t2n1)ocσ+.\displaystyle\quad\langle f(t-t_{1})f(t_{2}-t_{3})\cdots f(t_{2n-2}-t_{2n-1})\rangle_{\text{oc}}\sigma_{+}\ . (150)

Here, the terms in the summation follow the rule of considering all possible arrangements of the memory kernel f(titj)f(t_{i}-t_{j}). It is important to ensure that the times are properly ordered as tmtitjtt_{m}\leq...\leq t_{i}\leq t_{j}\leq...\leq t. Consequently, we can expand γ(t)\gamma(t) and S(t)S(t) as:

γ(t)=n=1λ2nγ2n(t),𝒮(t)=n=1λ2n𝒮2n(t).\displaystyle\gamma(t)=\sum_{n=1}^{\infty}\lambda^{2n}\gamma_{2n}(t)\ ,\quad\mathcal{S}(t)=\sum_{n=1}^{\infty}\lambda^{2n}\mathcal{S}_{2n}(t)\ . (151)

In particular, if we define the function

Z(t,t)0t𝑑t1f(tt1)\displaystyle Z(t,t^{\prime})\equiv\int_{0}^{t}dt_{1}f(t^{\prime}-t_{1})

we obtain, as shown in Ref. [16], the following expressions

γ2(t)+iS2(t)\displaystyle\gamma_{2}(t)+iS_{2}(t) =20t𝑑t1f(tt1)=2limttZ(t,t)\displaystyle=2\int_{0}^{t}dt_{1}f(t-t_{1})=2\lim_{t^{\prime}\rightarrow t}Z(t,t^{\prime}) (152a)
γ4(t)+iS4(t)\displaystyle\gamma_{4}(t)+iS_{4}(t) =20t𝑑t10t1𝑑t20t2𝑑t3\displaystyle=2\int_{0}^{t}dt_{1}\int_{0}^{t_{1}}dt_{2}\int_{0}^{t_{2}}dt_{3} (152b)
[f(tt2)f(t1t3)+f(tt3)f(t1t2)].\displaystyle\!\!\!\!\!\left[f(t-t_{2})f(t_{1}-t_{3})+f(t-t_{3})f(t_{1}-t_{2})\right]\ .

Thus,

𝒮2(t)[γ2(t)]=2Im[Re]limttZ(t,t)\displaystyle\mathcal{S}_{2}(t)[\gamma_{2}(t)]=2\imaginary[\real]\lim_{t^{\prime}\rightarrow t}Z(t,t^{\prime}) (153a)
𝒮4(t)[γ4(t)]=2Im[Re]0tdt10t1dt2(f(tt2)Z(t1,t2)\displaystyle\mathcal{S}_{4}(t)[\gamma_{4}(t)]=2\imaginary[\real]\int_{0}^{t}dt_{1}\int_{0}^{t_{1}}dt_{2}\left(f(t-t_{2})Z(t_{1},t_{2})\right.
+f(t1t2)Z(t,t2)),\displaystyle\qquad\left.+f(t_{1}-t_{2})Z(t,t_{2})\right)\ , (153b)

where the imaginary and real parts are taken for 𝒮\mathcal{S} and γ\gamma, respectively. The TCL2 and TCL4 cases correspond, respectively, to the following substitutions into Eq. 148:

TCL2:{𝒮(2)(t)=𝒮2(t)γ(2)(t)=γ2(t)\displaystyle\text{TCL2:}\begin{cases}\mathcal{S}^{(2)}(t)=\mathcal{S}_{2}(t)\\ \gamma^{(2)}(t)=\gamma_{2}(t)\end{cases} (154a)
TCL4:{𝒮(4)(t)=𝒮2(t)+𝒮4(t)g(4)(t)=γ2(t)+γ4(t)\displaystyle\text{TCL4:}\begin{cases}\mathcal{S}^{(4)}(t)=\mathcal{S}_{2}(t)+\mathcal{S}_{4}(t)\\ g^{(4)}(t)=\gamma_{2}(t)+\gamma_{4}(t)\end{cases} (154b)

This follows from Eq. 151 with λ=1\lambda=1 (recall that λ\lambda is a formal expansion parameter).

Refer to caption
(a)
Refer to caption
(b)
Figure 1: (a) Integrated trace-norm distance D[0,100]D_{[0,100]} between the exact solution and the Markovian approximations: CG-LE and RWA-LE, as a function of the dimensionless coarse-graining time ωcτ\omega_{c}\tau for the Ohmic bath spectral density J2J_{2} with η=1\eta=1 and Ω0/ωc=1\Omega_{0}/\omega_{c}=1. The minimum is obtained for a coarse-graining time τ=0.501939/ωc\tau=0.501939/\omega_{c}. (b) The coarse-graining at which D[0,100]D_{[0,100]} is minimized as a function of the qubit frequency Ω0/ωc\Omega_{0}/\omega_{c} for η=1\eta=1 for the three different spectral densities. The labeled red dots indicate the values of Ω0/ωc\Omega_{0}/\omega_{c} shown in each of the corresponding figures.

Next, we consider our three spectral densities.

III.4.1 J1=|g|2δ(ωωc)J_{1}=|g|^{2}\delta(\omega-\omega_{c})

We use the integral of the shifted memory kernel from Eq. 35:

Z(t,t1)=2|g|2sin(Ω0ωc)t12Ω0ωcei(Ω0ωc)(2tt1)2.\displaystyle Z(t,t_{1})=\frac{2|g|^{2}\sin\frac{(\Omega_{0}-\omega_{c})t_{1}}{2}}{\Omega_{0}-\omega_{c}}e^{i\frac{(\Omega_{0}-\omega_{c})(2t-t_{1})}{2}}\ . (155)

The second-order Lamb shift and decay rate are, using Eq. 153a:

𝒮2(t)\displaystyle\mathcal{S}_{2}(t) =2|g|21cos((Ω0ωc)t)Ω0ωc\displaystyle=2|g|^{2}\frac{1-\cos((\Omega_{0}-\omega_{c})t)}{\Omega_{0}-\omega_{c}} (156a)
γ2(t)\displaystyle\gamma_{2}(t) =2|g|2tsinc[(Ω0ωc)t)].\displaystyle=2|g|^{2}t\operatorname{sinc}[(\Omega_{0}-\omega_{c})t)]\ . (156b)

The fourth-order Lamb shift and decay rate are, using Eq. 153:

𝒮4(t)\displaystyle\mathcal{S}_{4}(t) =4|g|4sin2(Ω0ωc)t+(Ω0ωc)tsin(Ω0ωc)t(Ω0ωc)3\displaystyle=4|g|^{4}\frac{-\sin^{2}(\Omega_{0}-\omega_{c})t+(\Omega_{0}-\omega_{c})t\sin(\Omega_{0}-\omega_{c})t}{(\Omega_{0}-\omega_{c})^{3}} (157a)
γ4(t)\displaystyle\gamma_{4}(t) =2|g|42(Ω0ωc)tcos(Ω0ωc)tsin2(Ω0ωc)t(Ω0ωc)3\displaystyle=2|g|^{4}\frac{2(\Omega_{0}-\omega_{c})t\cos(\Omega_{0}-\omega_{c})t-\sin 2(\Omega_{0}-\omega_{c})t}{(\Omega_{0}-\omega_{c})^{3}} (157b)

III.4.2 J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}

Again we use the integral of the shifted memory kernel from Eq. 41:

Z(t,t1)\displaystyle Z(t,t_{1}) =iηωceiΩ0t1+iωctiηωceiΩ0(tt1)1+iωc(tt1)\displaystyle=\frac{i\eta\omega_{c}e^{i\Omega_{0}t}}{1+i\omega_{c}t}-\frac{i\eta\omega_{c}e^{i\Omega_{0}(t-t_{1})}}{1+i\omega_{c}(t-t_{1})}
+iJ(Ω0)(Ei(Ω0ωc(1+iωc(tt1)))\displaystyle\quad+iJ(\Omega_{0})\Big{(}\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}(1+i\omega_{c}(t-t_{1}))\right)
Ei(Ω0ωc(1+iωct))).\displaystyle\qquad-\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}(1+i\omega_{c}t)\right)\Big{)}\ . (158)

The second-order Lamb shift and decay rate are, using Eq. 153a:

𝒮2(t)\displaystyle\mathcal{S}_{2}(t) =2ηωc+2ηωccos(Ω0t)+ωctsin(Ω0t)1+(ωct)2\displaystyle=-2\eta\omega_{c}+2\eta\omega_{c}\frac{\cos(\Omega_{0}t)+\omega_{c}t\sin(\Omega_{0}t)}{1+(\omega_{c}t)^{2}}
2J2(Ω0)(Re[Ei(Ω0ωc(1+iωct))]Ei(Ω0ωc))\displaystyle-2J_{2}(\Omega_{0})\left(\real\left[\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}(1+i\omega_{c}t)\right)\right]-\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}\right)\right) (159a)
γ2(t)\displaystyle\gamma_{2}(t) =2ηωcωctcos(Ω0t)sin(Ω0t)1+(ωct)2\displaystyle=2\eta\omega_{c}\frac{\omega_{c}t\cos(\Omega_{0}t)-\sin(\Omega_{0}t)}{1+(\omega_{c}t)^{2}}
+2J2(Ω0)Im[Ei(Ω0ωc(1+iωct))].\displaystyle\quad+2J_{2}(\Omega_{0})\imaginary\left[\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}(1+i\omega_{c}t)\right)\right]\ . (159b)

The Markovian decay rate is obtained in the long time limit. Using limxEi(ix+1)=iπ\lim_{x\rightarrow\infty}\operatorname{Ei}(ix+1)=i\pi, we obtain:

γM=limtγ2(t)=2πηΩ0eΩ0ωc=2πJ(Ω0),\displaystyle\gamma_{M}=\lim_{t\rightarrow\infty}\gamma_{2}(t)=2\pi\eta\Omega_{0}e^{-\frac{\Omega_{0}}{\omega_{c}}}=2\pi J(\Omega_{0})\ , (160)

in agreement with Eq. 127b.

For the fourth-order decay rate, Eq. 153 does not admit a closed-form solution and needs to be evaluated numerically.

III.4.3 J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega)

We use the shifted integral of the memory kernel in Eq. 44:

Z(t,t1)\displaystyle Z(t,t_{1}) =ηt(tt1)(ei(Ω0ωc)(tt1)teiΩ0(tt1)t\displaystyle=\frac{\eta}{t(t-t_{1})}\left(e^{i(\Omega_{0}-\omega_{c})(t-t_{1})}t-e^{i\Omega_{0}(t-t_{1})}t\right.
+(eiΩ0tei(Ω0ωc)t)(tt1))\displaystyle\left.+(e^{i\Omega_{0}t}-e^{i(\Omega_{0}-\omega_{c})t})(t-t_{1})\right) (161)
+iηΩ0[Ei(i(Ω0ωc)t)Ei(iΩ0t)\displaystyle+i\eta\Omega_{0}\left[\operatorname{Ei}(i(\Omega_{0}-\omega_{c})t)-\operatorname{Ei}(i\Omega_{0}t)\right.
Ei(i(Ω0ωc)(tt1))+Ei(iΩ0(tt1))].\displaystyle\left.-\operatorname{Ei}(i(\Omega_{0}-\omega_{c})(t-t_{1}))+\operatorname{Ei}(i\Omega_{0}(t-t_{1}))\right]\ .

Thus:

𝒮2(t)\displaystyle\mathcal{S}_{2}(t) =2iη(sin(Ω0t)sin((Ω0ωc)t)t\displaystyle=-2i\eta\Big{(}\frac{\sin(\Omega_{0}t)-\sin((\Omega_{0}-\omega_{c})t)}{t}
Ω0ln|1ωcΩ0|ωc\displaystyle\quad-\Omega_{0}\ln|1-\frac{\omega_{c}}{\Omega_{0}}|-\omega_{c}
+Re(Ω0(Ei(i(Ω0ωc)t)Ei(iΩ0t))))\displaystyle\quad+\real(\Omega_{0}\left(\operatorname{Ei}(i(\Omega_{0}-\omega_{c})t)-\operatorname{Ei}(i\Omega_{0}t))\right)\Big{)} (162a)
γ2(t)\displaystyle\gamma_{2}(t) =2η(cos(Ω0t)cos((Ω0ωc)t)t\displaystyle=2\eta\Big{(}\frac{\cos(\Omega_{0}t)-\cos((\Omega_{0}-\omega_{c})t)}{t}
πΩ0Θ(ωcΩ0)\displaystyle\quad-\pi\Omega_{0}\Theta(\omega_{c}-\Omega_{0})
+Ω0Im[Ei(iΩ0t)Ei(i(Ω0ωc)t)]).\displaystyle\quad+\Omega_{0}\imaginary\left[\operatorname{Ei}(i\Omega_{0}t)-\operatorname{Ei}(i(\Omega_{0}-\omega_{c})t)\right]\Big{)}\ . (162b)

We recover the Markov approximation in the large tt limit:

limtγ2(t)=2πηJ3(Ω0).\displaystyle\lim_{t\rightarrow\infty}\gamma_{2}(t)=2\pi\eta J_{3}(\Omega_{0})\ . (163)

While the TCL result Eq. 162 does not entirely match the exact result, it does describe an oscillatory behavior similar to the exact solution, which is entirely absent in the Markovian limit. A similar phenomenon has been described in the context of non-equilibrium dynamics in Ref. [30].

Moreover, recall that the Markovian rate vanishes when Ω0>ωc\Omega_{0}>\omega_{c} [Eq. 130b], resulting in the absence of decay. In contrast, when Ω0>ωc\Omega_{0}>\omega_{c} we find:

0γ2(t)𝑑t=2η(ωcΩ0ωc+ln(Ω0ωcΩ0)),\displaystyle\int_{0}^{\infty}\gamma_{2}(t^{\prime})dt^{\prime}=2\eta\left(\frac{\omega_{c}}{\Omega_{0}-\omega_{c}}+\ln\left(\frac{\Omega_{0}-\omega_{c}}{\Omega_{0}}\right)\right)\ , (164)

so that the asymptotic limit for the population in the TCL2 approximation (which coincides with the Redfield equation), e0γ2(t)𝑑te^{-\int_{0}^{\infty}\gamma_{2}(t^{\prime})dt^{\prime}}, is non-zero. This qualitatively recovers the non-zero decay property of the exact solution.

For the fourth-order decay rate, Eq. 153 again does not admit a closed-form solution and needs to be evaluated numerically.

J2(ω)=ηωeω/ωcη=1Ω0/ωc=1\boxed{J_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}\qquad\eta=1\qquad\Omega_{0}/\omega_{c}=1}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 2: The trace-norm distance ρexactρapprox1\|\rho_{\text{exact}}-\rho_{\text{approx}}\|_{1} (a), decay rate ratio γ/γRWA\gamma/\gamma_{\text{RWA}} (b), Lamb shift ratio S/SRWAS/S_{\text{RWA}} (c), population ρ11\rho_{11} (d) and coherence in its real part Re(ρ01)\real(\rho_{01}) (e) and imaginary part Im(ρ01)\imaginary(\rho_{01}) (f) with an initial state ρ(0)=|++|\rho(0)=|{+}\rangle\!\langle+| for Ohmic spectral density J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}} as a function of dimensionless time ωct\omega_{c}t for coupling η=1\eta=1 and qubit frequency ratio Ω0/ωc=1\Omega_{0}/\omega_{c}=1. Five different approaches are depicted in the plots: the exact solution (Exact), TCL2 (Redfield), TCL4, CG-LE and RWA-LE (Markov). The CG-LE coarse graining time is τ=0.501939/ωc\tau=0.501939/\omega_{c}.
J2(ω)=ηωeω/ωcη=1/2Ω0/ωc=1\boxed{J_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}\qquad\eta=1/2\qquad\Omega_{0}/\omega_{c}=1}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 3: Same as Fig. 2 with weak coupling η=1/2\eta=1/2 and a qubit frequency equal to the cutoff: Ω0/ωc=1\Omega_{0}/\omega_{c}=1. The CG-LE coarse graining time is τ=0.659732/ωc\tau=0.659732/\omega_{c}.
J2(ω)=ηωeω/ωcη=1Ω0/ωc=1/2\boxed{J_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}\qquad\eta=1\qquad\Omega_{0}/\omega_{c}=1/2}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 4: Same as Fig. 2 with an Ohmic spectral density J2(ω)J_{2}(\omega) and the same coupling η=1\eta=1, but with a lower qubit frequency: Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2. The CG-LE coarse graining time is τ=0.999876/ωc\tau=0.999876/\omega_{c}.
J2(ω)=ηωeω/ωcη=1Ω0/ωc=4\boxed{J_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}\qquad\eta=1\qquad\Omega_{0}/\omega_{c}=4}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 5: Same as Fig. 2 with an Ohmic spectral density J2(ω)J_{2}(\omega) and the same coupling η=1\eta=1, but with a larger qubit frequency: Ω0/ωc=4\Omega_{0}/\omega_{c}=4. The CG-LE coarse graining time is τ=0.025953/ωc\tau=0.025953/\omega_{c}.
exact C-LE/CG-LE RWA-LE TCL2 TCL4
population ρ11\rho_{11} Eq. 34a Eq. 83a Eq. 123a Eqs. 148a and 154a Eqs. 148a and 154b
coherence ρ01\rho_{01} Eq. 34b Eq. 83b Eq. 123b Eqs. 148b and 154a Eqs. 148b and 154b
Lamb shift 𝒮\mathcal{S} J1J_{1} Eq. 40a 0 Eq. 124a Eq. 156a Eq. 157a
J2J_{2} 0 Eq. 127a Eq. 159
J3J_{3} 0 Eq. 130a Eq. 162
decay rate γ\gamma J1J_{1} Eq. 40b Eq. 84 Eq. 124b Eq. 156b Eq. 157b
J2J_{2} Eq. 85 Eq. 127b Eq. 159
J3J_{3} Eq. 86 Eq. 130b Eq. 162
Table 1: Analytical results. The equations listed in the table are explicit analytical results. The results for empty cells are obtained numerically.

IV Analysis

In this section, we compare the exact solution for the excited state population ρ11(t)\rho_{11}(t) and the coherence to the results of the various approximation schemes we described above. We also compare the exact Lamb shift and decay rate γ(t)\gamma(t) to the corresponding quantities predicted by these approximation schemes. We do this for all three spectral densities where possible, for the CG-LE/C-LE, RWA-LE, TCL2 (Redfield), and TCL4. These results are summarized with corresponding equation references in Table 1.

IV.1 Computation of the optimal coarse graining time in CG-LE

Before presenting the comparison to the exact results, we first explain our methodology for optimizing the coarse-graining time τ\tau within the framework of the CG-LE, since in the ensuing comparison we use the optimal τ\tau values. Recall that the coarse-graining time needs to satisfy the condition ωcτ1\omega_{c}\tau\ll 1 [Eq. 65].

To determine the appropriate coarse-graining timescale τ\tau, we minimize a metric that quantifies the deviation from the exact solution. We employ the integrated trace-norm distance norm [18] as our chosen metric, defined as:

𝒟[0,T]12T0T𝑑tρexact(t)ρapprox(t)1,\displaystyle\mathcal{D}_{[0,T]}\equiv\frac{1}{2T}\int_{0}^{T}dt\|\rho_{\text{exact}}(t)-\rho_{\text{approx}}(t)\|_{1}\ , (165)

where M1TrMM\|M\|_{1}\equiv\mathrm{Tr}\sqrt{M^{\dagger}M}. Since M=ρexactρapproxM=\rho_{\text{exact}}-\rho_{\text{approx}} is Hermitian and traceless, it can be written as (abba)\left(\begin{array}[]{cc}a&b\\ b^{*}&-a\\ \end{array}\right) in the qubit case, so that TrM2=2a2+|b|2\mathrm{Tr}\sqrt{M^{2}}=2\sqrt{a^{2}+|b|^{2}}, and we can simplify the integrand as follows:

ρexactρapprox1\displaystyle\|\rho_{\text{exact}}-\rho_{\text{approx}}\|_{1} (166)
=2(ρ11,exactρ11,approx)2+|ρ01,exactρ01,approx|2.\displaystyle=2\sqrt{(\rho_{11,\text{exact}}-\rho_{11,\text{approx}})^{2}+|\rho_{01,\text{exact}}-\rho_{01,\text{approx}}|^{2}}\ .

In Fig. 1(a), we illustrate the distance 𝒟\mathcal{D} between the exact solution and both the CG-LE and the RWA-LE as a function of the dimensionless quantity ωcτ\omega_{c}\tau, where we vary the coarse-graining time τ\tau. In this example, we use the parameters η=1\eta=1 and Ω0/ωc=1\Omega_{0}/\omega_{c}=1, and the numerical integration is performed up to a total time of ωcT=100\omega_{c}T=100. This upper limit is justified by the observation that the distance is minimized at relatively short times. For example, in Fig. 1(a) the minimum occurs at ωcτ=0.50139\omega_{c}\tau=0.50139, thus satisfying Eq. 65. Notably, the coarse-graining solution outperforms the RWA-LE solution, consistent with the findings reported in [18].

In Fig. 1(b), we present the minimum coarse-graining time as a function of the qubit frequency Ω0/ωc\Omega_{0}/\omega_{c} for η=1\eta=1 across the three different spectral models. The coarse-graining times used in the later figures are indicated by the red dots in the plot. We selected Ω0/ωc[0.15,2]\Omega_{0}/\omega_{c}\in[0.15,2] to account for a range of small to large qubit frequencies.

The most notable conclusion from Fig. 1(b) is that condition ωcτ1\omega_{c}\tau\ll 1 cannot be satisfied for the impulse (J1J_{1}) and triangular (J3J_{3}) spectral densities within the range of Ω0/ωc\Omega_{0}/\omega_{c} values shown. However, it is satisfied for the Ohmic spectral density (J2J_{2}), the most physically relevant of the three. We thus expect the CG-LE to perform poorly for J1J_{1} and J3J_{3}, but to perform relatively well for J2J_{2}. These expectations are borne out in our results below.

Note further that for the Ohmic spectral density, the coarse-graining timescale increases as the qubit frequency decreases, tending to the RWA-LE solution in line with Eq. 121. Conversely, for the spectral density J1=|g|2δ(ωωc)J_{1}=|g|^{2}\delta(\omega-\omega_{c}), characterized by strong non-Markovian behavior and oscillations that prevent the fitting of a Markovian exponential decay, the coarse-graining time diverges at ωωc\omega-\omega_{c} to attempt to fit the exact solution.

We remark that to ensure the validity of RWA-LE, the weak coupling limit Eq. 110 needs to be satisfied, a point we comment on in more detail below.

IV.2 Exact solution vs TCL and Markov approximations for the Ohmic spectral density

We now present the results of a comparison between the exact solution and the different master equations for the Ohmic spectral density, the most physically interesting of the three densities. Our general expectations are that the TCL approximations will capture some of the aspects of the non-Markovian dynamics and will thus outperform the two Markovian master equations, and that the CG-LE will outperform the RWA-LE due to the ability to optimize the coarse-graining timescale τ\tau in the former. This expectation depends on the validity condition ωcτ1\omega_{c}\tau\ll 1 not being violated, as explained in the previous section. Indeed, we find that this holds for all the examples we consider in this section, at least in the sense that we find ωcτ<1\omega_{c}\tau<1 in all cases.

For the initial condition, we consider a simple scenario where c1(0)=c0(0)=12c_{1}(0)=c_{0}(0)=\frac{1}{2} and ck(0)=0c_{k}(0)=0. In this setup, the system is initially in the state |+=12(|0+|1)\ket{+}=\frac{1}{\sqrt{2}}(\ket{0}+\ket{1}), while the bath is initially in the vacuum state |v\ket{v} of the cavity. To determine the amplitude for the exact solution, from which we obtain the population ρ11(t)=|c11(t)|2\rho_{11}(t)=|c_{11}(t)|^{2} of the excited state |1\ket{1}, we need to apply a numerical inverse Laplace transform to Eq. 24, where the Laplace transform of the memory kernel f^(s)\hat{f}(s) is given by Section II.4.2. Irrespective of the specific positive values of η\eta, ωc\omega_{c}, and Ω0\Omega_{0}, the roots of Eq. 24 are found to be complex. This means that the amplitude c1(t)c_{1}(t) is oscillatory. This trend persists even in cases of weak coupling, resulting in a damped oscillation of the amplitude.

IV.2.1 General observations

Figs. 2, 3, 4 and 5 offer a comparison of the temporal dynamics against the exact solution across four distinct approximations: TCL2, TCL4, CG-LE (==C-LE), and RWA-LE. The comparison metrics comprise the trace-norm distance ρexactρapprox1\|\rho_{\text{exact}}-\rho_{\text{approx}}\|_{1}, decay rate γ\gamma, Lamb shift SS, population ρ11\rho_{11}, and coherence ρ01\rho_{01}, all with the Ohmic spectral density J2J_{2}, but with different values of the coupling η\eta and qubit frequency ratio Ω0/ωc\Omega_{0}/\omega_{c}. The equations plotted for each curve are listed in Table 1.

During the initial time intervals (ωct1\omega_{c}t\approx 1), the TCL2 and TCL4 solutions generally exhibit much closer agreement with the exact solution than the Markovian CG-LE and RWA-LE, as evidenced especially by the trace-norm distance curves. The plots representing population and the real part of the coherence reveal that the TCL approximations aptly capture the Zeno effect observed in the exact solution – characterized by a gradual concave decay at short times. This stands in contrast to the monotonic exponential decay exhibited by the Markovian approaches.

Note that, as mentioned above in the discussion of Eq. 121, the TCL2 decay rate γ\gamma gradually approaches the asymptotic behavior of the RWA-LE decay rate. Concerning the CG-LE, we note that it exhibits a smaller trace-norm distance from the exact solution compared to the RWA-LE, as anticipated in Ref. [18]. This is due to the aforementioned ability to optimize the coarse-graining time. Both Markovian approximations exhibit a constant decay rate and Lamb shift, but in the case of the CG-LE, the Lamb shift SS is zero, resulting in the coherence being purely real.

IV.2.2 Weak and strong coupling

Recall that the RWA-LE needs to satisfy Eq. 110, in particular η1\eta\ll 1. In practice, we will use both weak (η=0.5\eta=0.5) and strong (η=1\eta=1) coupling. We thus expect the RWA-LE to be a relatively poor approximation in the later case, at least for relatively short evolutions.

Figs. 2 and 3 compare the results for weak and strong coupling. Using a qubit frequency equal to the cutoff frequency Ω0/ωc=1\Omega_{0}/\omega_{c}=1, Fig. 2 shows the strong coupling results, while Fig. 3 shows the weak coupling results.

After rising first, the exact decay rate [Eq. 33b] for strong coupling (η=1\eta=1) in Fig. 2(b) exhibits a negative trend and continues to exhibit non-monotonic behavior for even longer evolution times. Intervals during which γ\gamma turns negative correspond to non-Markovian dynamics, leading to an increase in population, as evidenced in Fig. 2(d). Consequently, during intermediate times, the approximation methods struggle to accurately match the exact solution, which displays a remarkably slow population decay.

In contrast, for weak coupling (η=1/2\eta=1/2), the decay rate in Fig. 3(b) remains positive, causing the population and coherence to approach zero more rapidly. This aligns closely with the behavior exhibited by the approximation methods, as demonstrated in Figs. 3(d), 3(e) and 3(f). This trend is also affirmed by the trace-norm distance plots in Figs. 2(a) and 3(a), where weaker coupling results in a smaller trace-norm distance difference between the approximation methods and the exact solution.

IV.2.3 Weak and strong qubit frequency

Fixing the coupling at η=1\eta=1 [technically at the upper limit of Eq. 110], we compare the cases where the qubit frequency Ω0\Omega_{0} is either less than or greater than the cutoff frequency ωc\omega_{c}, as shown in Figs. 4 and 5, respectively. Recall that Eq. 110 also imposes the validity condition Ω0/ωc>1\Omega_{0}/\omega_{c}>1 on the RWA-LE, so we expect better agreement in Fig. 5, as in indeed the case.

In more detail, from the trace-norm distance plots in Figs. 4(a) and 5(a), we observe that all the approximations exhibit closer agreement with the exact result when the qubit frequency is relatively high, especially at longer evolution times. This trend is also noticeable in the population (Figs. 4(d) and 5(d)) and coherence decay (Figs. 4(e), 5(e), 4(f) and 5(f)). For low qubit frequencies, as in Fig. 4(b), the exact decay rate γ\gamma exhibits non-Markovian negative phases. Correspondingly, the population in Fig. 4(d) increases initially and then stabilizes at a non-zero value. The approximations fail to match the exact solution’s behavior for extended periods, with the TCL approximation being effective primarily during short times (ωct<1.5\omega_{c}t<1.5). This behavior is characteristic of strong non-Markovian behavior, which is expected in models with non-flat spectral densities such as the ones considered here.

Conversely, for higher qubit frequencies in Fig. 5(b), the behavior of the TCL decay rate resembles that of the exact solution. Even the exact Lamb shift in Fig. 5(c) closely aligns with the approximation methods, with the notable exception of CG-LE (where the Lamb shift is zero). Consequently, in this scenario, the approximations reasonably reproduce the exact behavior, as demonstrated in Figs. 5(d), 5(e) and 5(f). The TCL4 approximation is particularly good according to all six of our metrics.

J1(ω)=|g|2δ(ωωc)|g|=1Ω0/ωc=1/2\boxed{J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c})\qquad|g|=1\qquad\Omega_{0}/\omega_{c}=1/2}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 6: Same as Fig. 2 but with an impulse bath spectral density J1(ω)=|g|2δ(ωωc)J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c}) with coupling |g|=1|g|=1 and a low qubit frequency: Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2. The CG-LE coarse graining time is τ=7.197305/ωc\tau=7.197305/\omega_{c}.
J1(ω)=|g|2δ(ωωc)|g|=1Ω0/ωc=2\boxed{J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c})\qquad|g|=1\qquad\Omega_{0}/\omega_{c}=2}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 7: Same as Fig. 2 but with an impulse bath spectral density J1(ω)J_{1}(\omega) with coupling |g|=1|g|=1 and a high qubit frequency: Ω0/ωc=2\Omega_{0}/\omega_{c}=2. The CG-LE coarse graining time is τ=3.609881/ωc\tau=3.609881/\omega_{c}.
J3(ω)=ηΘ(ωcω)η=1Ω0/ωc=1/2\boxed{J_{3}(\omega)=\eta\Theta(\omega_{c}-\omega)\qquad\eta=1\qquad\Omega_{0}/\omega_{c}=1/2}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 8: Same as Fig. 2 but with a triangular bath spectral density J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega) with coupling η=1\eta=1 and a low qubit frequency: Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2. The CG-LE coarse graining time is τ=2.631777/ωc\tau=2.631777/\omega_{c}.
J3(ω)=ηΘ(ωcω)η=1Ω0/ωc=1.8\boxed{J_{3}(\omega)=\eta\Theta(\omega_{c}-\omega)\qquad\eta=1\qquad\Omega_{0}/\omega_{c}=1.8}
Refer to caption
Refer to caption
(a) Trace-norm distance
Refer to caption
(b) Decay rate
Refer to caption
(c) Lamb shift
Refer to caption
(d) Population
Refer to caption
(e) Coherence – real part
Refer to caption
(f) Coherence – imaginary part
Figure 9: Same as Fig. 2 but with a triangular bath spectral density J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega) with coupling η=1\eta=1 and a high qubit frequency: Ω0/ωc=1.8\Omega_{0}/\omega_{c}=1.8. The CG-LE coarse graining time is τ=1.8629886/ωc\tau=1.8629886/\omega_{c}. Note that in panel (b), the RWA-LE decay rate is zero due to Eq. 130b.

IV.3 Exact solution vs TCL and Markov approximations for the impulse spectral density

For the impulse spectral density J1(ω)=|g|2δ(ωωc)J_{1}(\omega)=|g|^{2}\delta(\omega-\omega_{c}), the exact solution for the population |c1(t)|2|c_{1}(t)|^{2} of the excited population [see Eq. 38] is purely periodic. Consequently, no Markovian approximation with exponential decay can accurately capture the exact solution. This renders the RWA-LE inadequate for the description, as remarked in the discussion of Eq. 124b. However, since the CG-LE retains the coarse-graining time-scale as an optimization parameter, we use it to understand how close a Markovian approximation can still be to the exact solution in this rather extreme non-Markovian case, even though we have already concluded that the validity condition of the CG-LE cannot be satisfied [recall the discussion in Section IV.1].

Figs. 6 and 7 depict the time evolution of the same metrics as in the previous section but for the impulse spectral density J1J_{1} with two different qubit frequencies: Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2 and Ω0/ωc=2\Omega_{0}/\omega_{c}=2. For this spectral density, we have closed analytical expressions for the TCL Lamb shift [see Eqs. 156a and 157a] and decay rate [see Eqs. 156b and 157b]. In general, we observe that as expected, the CG-LE is a poor approximation to both the short-time and longer-time oscillatory behavior of the decay rate, population, and coherence.

In contrast, the TCL2 and TCL4 approximations in Eqs. 156b and 157b are fairly accurate for short-time dynamics. However, the populations and coherences under TCL4 deviate rapidly from the exact solution for longer times, as demonstrated in Figs. 6(d), 7(d), 6(e), 6(f), 7(e) and 7(f). The deviation is accentuated by the recurrent intervals of negative decay rates in the exact solution and TCL approximations. This is especially noticeable in TCL4, where the oscillations of the decay rates increase with time, taking on higher negative values, as seen in Figs. 6(b) and 7(b). Consequently, TCL4 develops an instability reflected in the diverging oscillation frequency of the coherence seen in Figs. 6(e) and 6(f). An explanation of the breakdown of the TCL approximation is given in Appendix I.

When comparing the low qubit frequency case (Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2) in Figs. 6(d), 6(e) and 6(f) with the high qubit frequency case (Ω0/ωc=2\Omega_{0}/\omega_{c}=2) in Figs. 7(d), 7(e) and 7(f), a distinction emerges. In the former, TCL2 exhibits a decay to zero similar to CG-LE, while in the latter, TCL2 mirrors the oscillatory behavior of the exact solution. This observation underscores the higher accuracy of the TCL approximation as the qubit frequency increases. However, while TCL4 is a better approximation at short times ωct<1.5\omega_{c}t<1.5, it is worse than TCL2 at long times for the impulse spectral density.

IV.4 Exact solution vs TCL and Markov approximations for the triangular spectral density

The triangular spectral density J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega) is intermediate between the impulse density J1(ω)J_{1}(\omega) and the Ohmic density J2(ω)J_{2}(\omega) because it matches J2(ω)J_{2}(\omega) for low frequencies and drops to zero for frequencies exceeding the cutoff. In Figs. 8 and 9, we perform a comparison for Ω0/ωc=1/2\Omega_{0}/\omega_{c}=1/2 and Ω0/ωc=1.8\Omega_{0}/\omega_{c}=1.8. Recall that here too, the validity condition of the CG-LE cannot be satisfied, as discussed in Section IV.1.

Similar to the impulse case J1J_{1}, for J3J_{3}, we observe a strong non-Markovian oscillatory behavior in the exact solution’s population and coherence, in Figs. 8(d), 8(e), 8(f), 9(d), 9(e) and 9(f), respectively. The decay rate (as seen in Figs. 8(b) and 9(b)) oscillates between positive and negative values.

As previously discussed, the TCL approximation is capable of capturing the quantum Zeno effect, which leads to a better agreement with the exact solution than the Markovian approximation does for short evolution times. However, a significant difference emerges in the exact decay rate between low and high qubit frequencies, as shown in Figs. 8(b) and 9(b). In the latter case, the decay rate oscillates and converges to zero (equivalent to the rate of the Markov approximations). Conversely, in the former case, the decay rate exhibits discontinuous behavior, rendering the approximation methods unsuitable for fitting the exact solution. In general, when Ω0<ωc\Omega_{0}<\omega_{c}, the discontinuous behavior of the decay rate leads to oscillatory population dynamics, where the population first decays to zero, then revives, and subsequently decays again. However, for Ω0>ωc\Omega_{0}>\omega_{c}, the decay rate converges to zero, resulting in non-decaying population dynamics, similar to what was observed for J1J_{1} in Fig. 7(d).

Furthermore, as depicted in Fig. 8(c), the exact Lamb shift SS diverges when Ω0<ωc\Omega_{0}<\omega_{c} in contrast to the zero Lamb shift in the CG-LE, the constant non-zero SS in the RWA-LE, and TCL2 (which converges to RWA). However, TCL4 diverges in an attempt to match the exact SS. In contrast, for the Lamb shift SS when Ω0>ωc\Omega_{0}>\omega_{c} (as shown in Fig. 9(c)), the TCL approximations align with the oscillatory behavior of the exact solution. Even the constant RWA-LE approximation closely resembles the asymptotic behavior. However, the CG solution maintains a zero Lamb shift.

Moving on to the population and coherence in Ω0<ωc\Omega_{0}<\omega_{c}, as illustrated in Figs. 8(d), 8(e) and 8(f), we observe that the approximations rapidly decay to zero, matching the exact solution at short times. However, the oscillatory behavior at later times remains beyond the reach of the approximations, similar to what was observed for J1J_{1} in Fig. 6(d).

In contrast, for the population and coherence at Ω0>ωc\Omega_{0}>\omega_{c}, as depicted in Figs. 9(d), 9(e) and 9(f), the TCL approximation exhibits an oscillatory and non-decaying behavior, akin to the exact solution, matching the Zeno effect at short times. TCL4, in particular, matches the exact solution well up to ωct2\omega_{c}t\approx 2. The CG-LE approximation performs much better than RWA-LE for short times in describing the excited state population, which remains constant in the latter due to its zero decay rate [as can be seen from Eq. 130b]. The two are comparable in describing the coherence. For long times, the Markov approximations, which display monotonic decay, fail to capture the exact behavior. CG-LE, which aims to minimize the trace-norm distance with the exact solution, decays slowly to achieve its objective.

In summary, for the triangular spectral density, when Ω0>ωc\Omega_{0}>\omega_{c}, the non-Markovian oscillatory behavior is captured qualitatively by the TCL approximation, as opposed to CG or RWA. In contrast, when Ω0<ωc\Omega_{0}<\omega_{c}, all approximations exhibit rapid decay and fail to accurately capture the exact solution.

V Summary and conclusions

In this work, we studied the dynamics of a qubit in a cavity interacting with bosonic baths described by three different spectral densities: impulse, Ohmic, and triangular. The model is exactly solvable within the single-excitation subspace, and this allowed us to perform a comprehensive comparison to a number of different master equations, both Markovian (CG-LE, C-LE, and RWA-LE) and non-Markovian (TCL2 and TCL4). Of the three spectral densities, the Ohmic model is the most physically relevant, and the other two were introduced primarily to allow us to reach closed-form analytical results, as well as to study extreme non-Markovian dynamics.

For weak coupling and a large qubit frequency, the Ohmic case leads to a quasi-exponential decay, characteristic of the Markovian limit. In this regime, therefore, we found that the purely Markovian master equations are able to approximate the exact solution rather well. Outside of this regime, in particular also for the impulse and triangular spectral densities, the Markovian master equations perform poorly, but we found that TCL is still relatively accurate, in particular for short evolution times. The TCL is also able to capture distinctly non-Markovian features such as bath-induced population and coherence oscillations.

Within the regime of validity of the Markovian master equations, we found the CG-LE and C-LE to be better approximations than the standard RWA-LE. This was achieved by optimizing the coarse-graining time to minimize the difference between the CG-LE and the exact solution.

Overall, this work shows that low-order quantum master equations can be accurate when operated in their guaranteed regime of validity (short evolution times, in particular), but significant caution must be exercised in trusting their predictions outside of these regimes, as they can dramatically deviate from the exact dynamics. The TCL approach stands out as significantly more accurate than the Markovian master equations, even when the latter are fine-tuned via the optimization of a free parameter such as the coarse-graining time. The standard Lindblad equation based on the rotating wave approximation is particularly suspect.

Future research may wish to address—within the context of the same analytically solvable model as studied here, or similar analytically solvable models—the accuracy of TCL at higher orders, as well as a variety of other master equations, such as the phenomenological post-Markovian master equation [31, 32], Floquet-based master equations for periodic driving [33, 34], time-dependent Markovian quantum master equations [35, 36, 37, 6, 38], the universal Lindblad equation [39], the geometric-arithmetic master equation [40, 41], regularized cumulant-based master equations [42], as well as various adiabatic master equations [43, 44, 45]. Another interesting direction worth considering is to incorporate more general initial states, such as a hermal Gibbs state, a coherent state [20] or a squeezed initial bath state.

Acknowledgements.
This research was supported by the ARO MURI grant W911NF-22-S-0007, by the National Science Foundation the Quantum Leap Big Idea under Grant No. OMA-1936388, and by the Defense Advanced Research Projects Agency (DARPA) under Contract No. HR001122C0063. We thank Prof. Todd Brun for useful discussions and Dawei Zhong for providing the analysis given in Appendix E.

Appendix A Detailed exact solution

Considering the Hamiltonian in Eq. 1, the general solution for the 11-excitation subspace is given by Eq. 14. This solution can be expressed as a linear combination of the basis eigenvectors {|ψ0,|ψ1,|φk}\{\ket{\psi}_{0},\ket{\psi}_{1},\ket{\varphi_{k}}\}, which are tensor products of the qubit system basis {|0,|1}\{\ket{0},\ket{1}\} and the 11-excitation bath basis {|v,|k}k1,2,..\{\ket{v},\ket{k}\}_{k\in 1,2,..}, as shown in Eq. 13. Here, |v\ket{v} denotes the vacuum state with no photons, and |k=bk|v\ket{k}=b_{k}^{\dagger}\ket{v} represents the state with one photon in mode kk. The coefficients of the linear combination satisfy the normalization condition given in Eq. 15. For the evolution of the closed system within the 11-excitation subspace, we utilize the Schrödinger equation with the interaction Hamiltonian in Eq. 11. Using

σ(t)|0\displaystyle\sigma_{-}(t)\ket{0} =σ+(t)|1=B(t)|v=0\displaystyle=\sigma_{+}(t)\ket{1}=B(t)\ket{v}=0 (167a)
σ+(t)|0\displaystyle\sigma_{+}(t)\ket{0} =eiΩ0t|1,σ(t)|1=eiΩ0t|0\displaystyle=e^{i\Omega_{0}t}\ket{1}\ ,\quad\sigma_{-}(t)\ket{1}=e^{-i\Omega_{0}t}\ket{0} (167b)
B(t)|k\displaystyle B(t)\ket{k} =gkeiωkt|v,B(t)|v=kgkeiωkt|k,\displaystyle=g_{k}e^{-i\omega_{k}t}\ket{v}\ ,\quad B^{\dagger}(t)\ket{v}=\sum_{k}g_{k}^{*}e^{i\omega_{k}t}\ket{k}\ , (167c)

this leads to Section II.3:

i|ϕ˙\displaystyle i\dot{\ket{\phi}} =c˙0(t)|ψ0+c˙1(t)|ψ1+k𝔠˙k(t)|φk\displaystyle=\dot{c}_{0}(t)\ket{\psi_{0}}+\dot{c}_{1}(t)\ket{\psi_{1}}+\sum_{k}\dot{\mathfrak{c}}_{k}(t)\ket{\varphi_{k}} (168a)
=λH~SB|ϕ(t)\displaystyle=\lambda\tilde{H}_{SB}\ket{\phi(t)} (168b)
=λ[σ+(t)B(t)+σ(t)B(t)]×\displaystyle=\lambda[\sigma_{+}(t)\otimes B(t)+\sigma_{-}(t)\otimes B^{\dagger}(t)]\times
(c0(t)|ψ0+c1(t)|ψ1+k𝔠˙k(t)|φk)\displaystyle\Big{(}c_{0}(t)\ket{\psi_{0}}+c_{1}(t)\ket{\psi_{1}}+\sum_{k}\dot{\mathfrak{c}}_{k}(t)\ket{\varphi_{k}}\Big{)} (168c)
=λ(σ+(t)|0B(t)k𝔠k(t)|k+\displaystyle=\lambda\Big{(}\sigma_{+}(t)\ket{0}\otimes B(t)\sum_{k}\mathfrak{c}_{k}(t)\ket{k}+
c1(t)σ(t)|1B(t)|v)\displaystyle\qquad\qquad c_{1}(t)\sigma_{-}(t)\ket{1}\otimes B^{\dagger}(t)\ket{v}\Big{)} (168d)
=λ(kgk𝔠k(t)ei(Ω0ωk)t|ψ1+\displaystyle=\lambda\Big{(}\sum_{k}g_{k}\mathfrak{c}_{k}(t)e^{i(\Omega_{0}-\omega_{k})t}\ket{\psi_{1}}+
kgkc1(t)ei(Ω0ωk)t|φk)\displaystyle\qquad\qquad\sum_{k}g_{k}^{*}c_{1}(t)e^{-i(\Omega_{0}-\omega_{k})t}\ket{\varphi_{k}}\Big{)} (168e)

We can obtain the joint system-bath density matrix using Eq. 14:

ρSB(t)=(|c0|2|00|+c0c1|01|+c1c0|10|\displaystyle\rho_{SB}(t)=\left(|c_{0}|^{2}|{0}\rangle\!\langle 0|+c_{0}c_{1}^{*}|{0}\rangle\!\langle 1|+c_{1}c_{0}^{*}|{1}\rangle\!\langle 0|\right.
+|c1|2|11|)|vv|+|00|(j,k𝔠j𝔠k|jk|\displaystyle\left.+|c_{1}|^{2}|{1}\rangle\!\langle 1|\right)\otimes|{v}\rangle\!\langle v|+|{0}\rangle\!\langle 0|\otimes\left(\sum_{j,k}\mathfrak{c}_{j}\mathfrak{c}_{k}^{*}|{j}\rangle\!\langle k|\right.
+k(c0𝔠k|vk|+c0𝔠k|kv|))+kc1𝔠k|10|\displaystyle\left.+\sum_{k}(c_{0}\mathfrak{c}_{k}^{*}|{v}\rangle\!\langle k|+c_{0}^{*}\mathfrak{c}_{k}|{k}\rangle\!\langle v|)\right)+\sum_{k}c_{1}\mathfrak{c}_{k}^{*}|{1}\rangle\!\langle 0|
|vk|+k𝔠kc1|01||kv|.\displaystyle\otimes|{v}\rangle\!\langle k|+\sum_{k}\mathfrak{c}_{k}c_{1}^{*}|{0}\rangle\!\langle 1|\otimes|{k}\rangle\!\langle v|\ . (169)

Taking the partial trace over the bath, we obtain the system state in the matrix form shown in Eq. 27. Its time derivative is given in Eq. 30. If we write the system density matrix in terms of the populations ρ00,ρ11\rho_{00},\rho_{11} and its coherences ρ01,ρ10\rho_{01},\rho_{10}, we have:

ρ11(t)\displaystyle\rho_{11}(t) =|c1(t)|2=1ρ00(t)\displaystyle=|c_{1}(t)|^{2}=1-\rho_{00}(t) (170a)
ρ01(t)\displaystyle\rho_{01}(t) =c0c˙1(t)=ρ10(t).\displaystyle=c_{0}\dot{c}^{*}_{1}(t)=\rho_{10}^{*}(t)\ . (170b)

Explicitly substituting the system density matrix into the ansatz Section II.3 along with Eq. 31, we obtain:

𝒦S(t)ρ=(γ(t)ρ1112(i𝒮(t)γ(t))ρ0112(i𝒮(t)γ(t))ρ10γ(t)ρ11)\displaystyle\mathcal{K}_{S}(t)\rho=\left(\begin{array}[]{cc}\gamma(t)\rho_{11}&\frac{1}{2}(i\mathcal{S}(t)-\gamma(t))\rho_{01}\\ \frac{1}{2}(-i\mathcal{S}(t)-\gamma(t))\rho_{10}&-\gamma(t)\rho_{11}\end{array}\right) (173)

Comparing this with Eq. 30 gives us the following differential equations:

t|c1(t)|2\displaystyle\partial_{t}|c_{1}(t)|^{2} =γ(t)|c1(t)|2\displaystyle=-\gamma(t)|c_{1}(t)|^{2} (174a)
c0c˙1(t)\displaystyle c_{0}^{*}\dot{c}_{1}(t) =12(i𝒮(t)+γ(t))c0c1(t).\displaystyle=-\frac{1}{2}\left(i\mathcal{S}(t)+\gamma(t)\right)c_{0}^{*}c_{1}(t)\ . (174b)

From the first equation, we obtain the population evolution Eq. 34a, and from the second one, we can take its real and imaginary parts to obtain Eq. 33.

Appendix B Simplification of the decay rate expressions for the CG-LE

B.1 Ohmic spectral density

Starting from Eq. 85 we have:

γ(τ)\displaystyle\gamma(\tau) =0ηωeω/ωcτsinc2((Ω0ω)τ2)𝑑ω\displaystyle=\int_{0}^{\infty}\eta\omega e^{-\omega/\omega_{c}}\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)d\omega (175a)
=2τηΩ0𝑑ν(ν+Ω0)e(ν+Ω0)/ωcν2(1cosντ),\displaystyle=\frac{2}{\tau}\eta\int_{-\Omega_{0}}^{\infty}d\nu\frac{(\nu+\Omega_{0})e^{-(\nu+\Omega_{0})/\omega_{c}}}{\nu^{2}}(1-\cos\nu\tau)\ , (175b)

where we used the change of variables ν=ωΩ0\nu=\omega-\Omega_{0}. This integral can be split into two parts. By using the exponential integral function Eq. 43c we have:

Ω0𝑑νeν/ωcν(1cosντ)=Ei(Ω0ωc)\displaystyle\int_{-\Omega_{0}}^{\infty}d\nu\frac{e^{-\nu/\omega_{c}}}{\nu}(1-\cos\nu\tau)=-\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}\right)
+12[Ei(Ω0ωc+iΩ0τ)+Ei(Ω0ωciΩ0τ)].\displaystyle\quad+\frac{1}{2}\left[\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}+i\Omega_{0}\tau\right)+\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}-i\Omega_{0}\tau\right)\right]\ . (176)

For the second integral, we can utilize integration by parts. Setting u=eν/ωc(1cosντ)u=e^{-\nu/\omega_{c}}(1-\cos\nu\tau) and dv=ν2dνdv=\nu^{-2}d\nu, we have:

Ω0u𝑑v=eΩ0/ωcΩ0(1cosΩ0τ)\displaystyle\int_{-\Omega_{0}}^{\infty}udv=-\frac{e^{\Omega_{0}/\omega_{c}}}{\Omega_{0}}(1-\cos\Omega_{0}\tau)
+Ω0𝑑νeν/ωcν(τsinντ1ωc(1cosντ)).\displaystyle+\int_{-\Omega_{0}}^{\infty}d\nu\frac{e^{-\nu/\omega_{c}}}{\nu}\left(\tau\sin\nu\tau-\frac{1}{\omega_{c}}(1-\cos\nu\tau)\right)\ . (177)

Since

Ω0𝑑νeν/ωcντsinντ=\displaystyle\int_{-\Omega_{0}}^{\infty}d\nu\frac{e^{-\nu/\omega_{c}}}{\nu}\tau\sin\nu\tau= (178)
i2τ[Ei(Ω0ωc+iΩ0τ)Ei(Ω0ωciΩ0τ)],\displaystyle-\frac{i}{2}\tau\left[\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}+i\Omega_{0}\tau\right)-\operatorname{Ei}\left(\frac{\Omega_{0}}{\omega_{c}}-i\Omega_{0}\tau\right)\right]\ ,

we can combine all the terms, and arrive at the final expression for γ(τ)\gamma(\tau) given in Eq. 85.

B.2 Triangular spectral density

Starting from Eq. 86 we have:

γ(τ)\displaystyle\gamma(\tau) =0ωcηωτsinc2((Ω0ω)τ2)𝑑ω\displaystyle=\int_{0}^{\omega_{c}}\eta\omega\tau\operatorname{sinc}^{2}\left(\frac{(\Omega_{0}-\omega)\tau}{2}\right)d\omega (179a)
=2τΩ0ωcΩ0𝑑νη(ν+Ω0)ν2(1cosντ).\displaystyle=\frac{2}{\tau}\int_{-\Omega_{0}}^{\omega_{c}-\Omega_{0}}d\nu\frac{\eta(\nu+\Omega_{0})}{\nu^{2}}(1-\cos\nu\tau)\ . (179b)

where we again used the change of variables ν=ωΩ0\nu=\omega-\Omega_{0}. Now for Ω0<ωc\Omega_{0}<\omega_{c}, we have:

Ω0ωcΩ0𝑑ν1cosντν=\displaystyle\int_{-\Omega_{0}}^{\omega_{c}-\Omega_{0}}d\nu\frac{1-\cos\nu\tau}{\nu}=
ln(ωcΩ0Ω0)Ci((ωcΩ0)τ)+Ci(Ω0τ),\displaystyle\ \ \ \ln\left(\frac{\omega_{c}-\Omega_{0}}{\Omega_{0}}\right)-\operatorname{Ci}((\omega_{c}-\Omega_{0})\tau)+\operatorname{Ci}(\Omega_{0}\tau)\ , (180)

where the sine and cosine integral functions are given in Eqs. 43a and 43b, respectively. The second term can be obtained by using integration by parts:

Ω0ωcΩ0𝑑ν1cosντν2\displaystyle\int_{-\Omega_{0}}^{\omega_{c}-\Omega_{0}}d\nu\frac{1-\cos\nu\tau}{\nu^{2}} =1cosντν|Ω0ωcΩ0\displaystyle=-\frac{1-\cos\nu\tau}{\nu}\bigg{|}_{-\Omega_{0}}^{\omega_{c}-\Omega_{0}}
+τΩ0ωcΩ0sin(ντ)ν.\displaystyle+\tau\int_{-\Omega_{0}}^{\omega_{c}-\Omega_{0}}\frac{\sin(\nu\tau)}{\nu}\ . (181)

The last term is the sine integral function. Combining, we arrive at the total rate as given by Eq. 86.

Appendix C Why the first order cumulant K(1)K^{(1)} in the C-LE can be made to vanish

Let B=Tr(ρBB)\langle{B}\rangle=\mathrm{Tr}(\rho_{B}B) and define a new, shifted bath operator:

BBBIB.B^{\prime}\equiv B-\langle{B}\rangle I_{B}\ . (182)

Its expectation value vanishes:

B=BBIB=0.\langle{B^{\prime}}\rangle=\langle{B}\rangle-\langle{B}\rangle\langle{I_{B}}\rangle=0\ . (183)

The corresponding bath interaction picture operator is:

B(t)\displaystyle B^{\prime}(t) =UB(t)BUB(t)=UB(t)(BBIB)UB(t)\displaystyle=U_{B}^{\dagger}(t)B^{\prime}U_{B}(t)=U_{B}^{\dagger}(t)(B-\langle{B}\rangle I_{B})U_{B}(t)
=B(t)BIB,\displaystyle=B(t)-\langle{B}\rangle I_{B}\ , (184)

where, as before, UB(t)=eiHBtU_{B}(t)=e^{-iH_{B}t}. Assuming stationarity [HB,ρB(0)]=0[H_{B},\rho_{B}(0)]=0 immediately implies [UB(t),ρB]=0[U_{B}(t),\rho_{B}]=0. In this case also B(t)=0\langle{B^{\prime}(t)}\rangle=0, since then:

B(t)\displaystyle\langle{B^{\prime}(t)}\rangle =Tr[UB(t)BUB(t)ρB]\displaystyle=\mathrm{Tr}[U_{B}^{\dagger}(t)B^{\prime}U_{B}(t)\rho_{B}] (185a)
=Tr[BUB(t)ρBUB(t)]=B=0.\displaystyle=\mathrm{Tr}[B^{\prime}U_{B}(t)\rho_{B}U_{B}^{\dagger}(t)]=\langle{B^{\prime}}\rangle=0\ . (185b)

Let

HSBAB,ΔHSBA,HSHS+ΔHS.H^{\prime}_{SB}\equiv A\otimes B^{\prime}\ ,\quad\Delta H_{S}^{\prime}\equiv\langle{B}\rangle A\ ,\quad H^{\prime}_{S}\equiv H_{S}+\Delta H^{\prime}_{S}\ . (186)

Correspondingly, the system-bath interaction can be written as

HSB=HSB+A(BB)=HSB+ΔHSIB.H_{SB}=H_{SB}^{\prime}+A\otimes(B-B^{\prime})=H_{SB}^{\prime}+\Delta H_{S}^{\prime}\otimes I_{B}\ . (187)

Thus, we can write the full Hamiltonian as:

H\displaystyle H =HSIB+HSB+ISHB\displaystyle=H_{S}\otimes I_{B}+H_{SB}+I_{S}\otimes H_{B} (188a)
=H0+HSB,H0HSIB+ISHB,\displaystyle=H^{\prime}_{0}+H^{\prime}_{SB}\ ,\quad H^{\prime}_{0}\equiv H^{\prime}_{S}\otimes I_{B}+I_{S}\otimes H_{B}\ , (188b)

where H0H^{\prime}_{0} defines a new, shifted interaction picture Hamiltonian.

Correspondingly, in this new interaction picture H~(t)=U0(t)HSBU0(t)=A(t)B(t)\tilde{H}^{\prime}(t)=U_{0}^{\prime{\dagger}}(t)H^{\prime}_{SB}U^{\prime}_{0}(t)=A(t)\otimes B^{\prime}(t), and we find, using Eq. 185:

TrB[H~(t),ρSB(0)]\displaystyle\mathrm{Tr}_{B}[\tilde{H}^{\prime}(t),\rho_{SB}(0)] =TrB[A(t)B(t),ρS(0)ρB(0)]\displaystyle=\mathrm{Tr}_{B}[A(t)\otimes B^{\prime}(t),\rho_{S}(0)\otimes\rho_{B}(0)]
=B(t)[A(t),ρS(0)]=0.\displaystyle=\langle{B^{\prime}(t)}\rangle[A(t),\rho_{S}(0)]=0\ . (189)

Therefore, K(1)(t)ρ(0)=0K^{\prime(1)}(t)\rho(0)=0, with K(1)K^{\prime(1)} defined within the shifted interaction picture and with the modified system-bath interaction HSBH^{\prime}_{SB}.

The extension to the case when HSBH_{SB} has the general form HSB=αAαBαH_{SB}=\sum_{\alpha}A_{\alpha}\otimes B_{\alpha} is immediate; in this case Bα=BαBαIBB^{\prime}_{\alpha}=B_{\alpha}-\langle{B_{\alpha}}\rangle I_{B} and:

HS=αBαAα,HSB=αAαBα.H_{S}^{\prime}=\sum_{\alpha}\langle{B_{\alpha}}\rangle A_{\alpha}\ ,\quad H^{\prime}_{SB}=\sum_{\alpha}A_{\alpha}\otimes B^{\prime}_{\alpha}\ . (190)

Now, for our bath operators B+(t)=kgkeiωktbkB_{+}(t)=\sum_{k}g_{k}e^{i\omega_{k}t}b_{k} and B=B+B_{-}=B_{+}^{*} and the bath state ρB=|vv|\rho_{B}=|{v}\rangle\!\langle v| we have that

B+(t)=Tr[ρBB+(t)]=kgkeiωktbk=0,\langle{B_{+}(t)}\rangle=\mathrm{Tr}[\rho_{B}B_{+}(t)]=\sum_{k}g_{k}e^{i\omega_{k}t}\langle{b_{k}}\rangle=0\ , (191)

and analogously B(t)=0\langle{B_{-}(t)}\rangle=0, both arising from the fact that the annihilation and creation operators average to zero [Section III.1]. As a result, in our case, in fact B(t)=B(t)B^{\prime}(t)=B(t).

Appendix D Proof of Eq. 97b

It is useful to relate αβω(t)\mathcal{B}_{\alpha\beta\omega}(t) with bαβω(t)b_{\alpha\beta\omega}(t):

αβω(t)=0t𝑑s0s𝑑sei(ω(ss)αβ(s,s)\displaystyle\mathcal{B}_{\alpha\beta\omega}(t)=\int_{0}^{t}ds\int_{0}^{s}ds^{\prime}e^{i(\omega(s-s^{\prime})}\mathcal{B}_{\alpha\beta}(s,s^{\prime}) (192)
=[0t𝑑s0t𝑑s0t𝑑sst𝑑s]ei(ωsωs)αβ(s,s)\displaystyle=\left[\int_{0}^{t}ds\int_{0}^{t}ds^{\prime}-\int_{0}^{t}ds\int_{s}^{t}ds^{\prime}\right]e^{i(\omega^{\prime}s-\omega s^{\prime})}\mathcal{B}_{\alpha\beta}(s,s^{\prime})
=[0t𝑑s0t𝑑s0t𝑑s0s𝑑s]eiω(ss)αβ(s,s)\displaystyle=\left[\int_{0}^{t}ds\int_{0}^{t}ds^{\prime}-\int_{0}^{t}ds^{\prime}\int_{0}^{s^{\prime}}ds\right]e^{i\omega(s-s^{\prime})}\mathcal{B}_{\alpha\beta}(s,s^{\prime})
=bαβω(t)0t𝑑s0s𝑑seiω(ss)αβ(s,s)\displaystyle=b_{\alpha\beta\omega}(t)-\int_{0}^{t}ds\int_{0}^{s}ds^{\prime}e^{-i\omega(s^{\prime}-s)}\mathcal{B}_{\alpha\beta}(s^{\prime},s)
=bαβω(t)αβω(t),\displaystyle=b_{\alpha\beta\omega}(t)-\mathcal{B}^{*}_{\alpha\beta\omega}(t)\ , (193)

where in the last line we used Eq. 94.

Appendix E [HB,ρB]0[H_{B},\rho_{B}]\neq 0

The bath state can be obtained via a partial trace of the system from the state Eq. 14, whose explicit pure density matrix is given in Appendix A:

ρB(t)=TrS|ϕ(t)ϕ(t)|=(|c0|2+|c1|2)|vv|\displaystyle\rho_{B}(t)=\mathrm{Tr}_{S}|{\phi(t)}\rangle\!\langle\phi(t)|=(|c_{0}|^{2}+|c_{1}|^{2})|{v}\rangle\!\langle v| (194)
+k𝔠k(t)c0|kv|+k𝔠k(t)c0|vk|+k,k𝔠k𝔠k|kk|.\displaystyle+\sum_{k}\mathfrak{c}_{k}(t)c_{0}^{*}|{k}\rangle\!\langle v|+\sum_{k}\mathfrak{c}_{k}^{*}(t)c_{0}|{v}\rangle\!\langle k|+\sum_{k,k^{\prime}}\mathfrak{c}_{k}\mathfrak{c}^{*}_{k^{\prime}}|{k}\rangle\!\langle k^{\prime}|\ .

Now using the bath Hamiltonian in Eq. 9 and using the identities nj|v=0n_{j}\ket{v}=0 and nj|k=δjk|jn_{j}\ket{k}=\delta_{jk}\ket{j}, we have:

HBρB\displaystyle H_{B}\rho_{B} =j,kωj𝔠j(t)𝔠k(t)|jk|+jωjc0𝔠j(t)|jv|\displaystyle=\sum_{j,k}\omega_{j}\mathfrak{c}_{j}(t)\mathfrak{c}_{k}^{*}(t)|{j}\rangle\!\langle k|+\sum_{j}\omega_{j}c_{0}^{*}\mathfrak{c}_{j}(t)|{j}\rangle\!\langle v| (195a)
=(ρBHB)ρBHB\displaystyle=(\rho_{B}H_{B})^{\dagger}\neq\rho_{B}H_{B} (195b)
=j,kωj𝔠k(t)𝔠j(t)|kj|+jωjc0𝔠j(t)|vj|.\displaystyle=\sum_{j,k}\omega_{j}\mathfrak{c}_{k}(t)\mathfrak{c}_{j}^{*}(t)|{k}\rangle\!\langle j|+\sum_{j}\omega_{j}c_{0}\mathfrak{c}_{j}^{*}(t)|{v}\rangle\!\langle j|. (195c)

In the particular case where ρB(0)=|vv|\rho_{B}(0)=|{v}\rangle\!\langle v|, or c0=0c_{0}=0 and 𝔠k(0)=0\mathfrak{c}_{k}(0)=0, it trivially follows that [HB,ρB(0)]=0[H_{B},\rho_{B}(0)]=0.

Appendix F Inadequacy of the Markov approximation

In the main text, we showed that we can reduce Eq. 112 to Eq. 120. However, this may lead to an unbounded approximation error, as we now show in detail.

Before the Markov approximation, Eq. 112 contains terms of the following form:

0t𝑑ταβ(±τ)e±iΩ0τσασβρ~(tτ),\displaystyle\int_{0}^{t}d\tau\mathcal{B}_{\alpha\beta}(\pm\tau)e^{\pm i\Omega_{0}\tau}\sigma_{\alpha}\sigma_{\beta}\tilde{\rho}(t-\tau)\ , (196)

where α,β{+,}\alpha,\beta\in\{+,-\}. The Markov approximation replaces the latter with

0𝑑ταβ(±τ)e±iΩ0τσασβρ~(t).\displaystyle\int_{0}^{\infty}d\tau\mathcal{B}_{\alpha\beta}(\pm\tau)e^{\pm i\Omega_{0}\tau}\sigma_{\alpha}\sigma_{\beta}\tilde{\rho}(t)\ . (197)

Therefore the approximation error is the difference between these two quantities, which we write as

δ=Δ1+Δ2Δ1+Δ2,\delta=\|\Delta_{1}+\Delta_{2}\|\leq\|\Delta_{1}\|+\|\Delta_{2}\|\ , (198)

where \|\cdot\| represents the operator norm and

Δ1\displaystyle\Delta_{1} 0𝑑ταβ(±τ)e±iΩ0τσασβ(ρ~(t)ρ~(tτ))\displaystyle\equiv\int_{0}^{\infty}d\tau\mathcal{B}_{\alpha\beta}(\pm\tau)e^{\pm i\Omega_{0}\tau}\sigma_{\alpha}\sigma_{\beta}(\tilde{\rho}(t)-\tilde{\rho}(t-\tau)) (199a)
Δ2\displaystyle\Delta_{2} t𝑑ταβ(±τ)e±iΩ0τσασβρ~(tτ).\displaystyle\equiv\int_{t}^{\infty}d\tau\mathcal{B}_{\alpha\beta}(\pm\tau)e^{\pm i\Omega_{0}\tau}\sigma_{\alpha}\sigma_{\beta}\tilde{\rho}(t-\tau)\ . (199b)

Let 1\|\cdot\|_{1} denote the trace norm and observe that ABAB1\|AB\|\leq\|A\|\|B\|_{1} for any pair of operators AA and BB.

For the Ohmic spectral density J2(ω)J_{2}(\omega) we have, using Eq. 125:

Δ2\displaystyle\|\Delta_{2}\| t𝑑τ|αβ(±τ)|σασβρ~(tτ)1\displaystyle\leq\int_{t}^{\infty}d\tau|\mathcal{B}_{\alpha\beta}(\pm\tau)|\,\|\sigma_{\alpha}\|\,\|\sigma_{\beta}\|\,\|\tilde{\rho}(t-\tau)\|_{1} (200a)
t𝑑τ|+(±τ)|=ηωctd(ωcτ)1+(ωcτ)2\displaystyle\leq\int_{t}^{\infty}d\tau|\mathcal{B}_{+-}(\pm\tau)|=\eta\omega_{c}\int_{t}^{\infty}\frac{d(\omega_{c}\tau)}{1+(\omega_{c}\tau)^{2}} (200b)
=(π2arctan(ωcτ))ηωc.\displaystyle=\left(\frac{\pi}{2}-\arctan(\omega_{c}\tau)\right)\eta\omega_{c}\ . (200c)

This quantity goes to zero for ωcτ1\omega_{c}\tau\gg 1 as required. On the other hand, the error term Δ1\Delta_{1} is unbounded. First, by the mean value theorem, there is a point t[tτ,t]t^{\prime}\in[t-\tau,t] such that:

ρ~(t)ρ~(tτ)τsupt[tτ,t]ρ~˙(t),\displaystyle\|\tilde{\rho}(t)-\tilde{\rho}(t-\tau)\|\leq\tau\sup_{t^{\prime}\in[t-\tau,t]}\|\dot{\tilde{\rho}}(t^{\prime})\|\ , (201)

so that:

Δ10𝑑ττ|Bαβ|supt[tτ,t]ρ~˙(t).\displaystyle\|\Delta_{1}\|\leq\int_{0}^{\infty}d\tau\,\tau|B_{\alpha\beta}|\sup_{t^{\prime}\in[t-\tau,t]}\|\dot{\tilde{\rho}}(t^{\prime})\|\ . (202)

We can bound ρ~˙(t)\|\dot{\tilde{\rho}}(t^{\prime})\| using the state evolution in Section III.3, where we undo the Markovian approximation by replacing ρ~(t)\tilde{\rho}(t) with ρ~(tτ)\tilde{\rho}(t-\tau) [i.e., returning to Eq. 112]:

ρ~˙(t)\displaystyle\|\dot{\tilde{\rho}}(t) 0tdτ|B+(τ)|[σ+,σρ~(tτ)]\displaystyle\|\leq\int_{0}^{t}d\tau|B_{+-}(\tau)|\,\|[\sigma_{+},\sigma_{-}\tilde{\rho}(t-\tau)]\|
+0t𝑑τ|B+(τ)|[σ+,ρ~(tτ)σ]\displaystyle\quad+\int_{0}^{t}d\tau|B_{-+}(-\tau)|\,\|[\sigma_{+},\tilde{\rho}(t-\tau)\sigma_{-}]\| (203a)
40t𝑑τ|B+(τ)|σ+σρ~(tτ)1\displaystyle\leq 4\int_{0}^{t}d\tau|B_{+-}(\tau)|\,\|\sigma_{+}\|\,\|\sigma_{-}\|\|\tilde{\rho}(t-\tau)\|_{1} (203b)
40t𝑑τ|B+(τ)|40𝑑τ|B+(τ)|\displaystyle\leq 4\int_{0}^{t}d\tau|B_{+-}(\tau)|\leq 4\int_{0}^{\infty}d\tau|B_{+-}(\tau)| (203c)
=2πηωc,\displaystyle=2\pi\eta\omega_{c}\ , (203d)

where in the last line we used Eq. 200c evaluated at t=0t=0.

While the integral

0τn|+(τ)|𝑑τ=π2ηωc1nsec(nπ/2)\displaystyle\int_{0}^{\infty}\tau^{n}|\mathcal{B}_{+-}(\tau)|d\tau=\frac{\pi}{2}\eta\omega_{c}^{1-n}\sec(n\pi/2) (204)

converges for |n|<1|n|<1, it does not for |n|1|n|\geq 1. Indeed, for the Ohmic spectral density we have:

0τ|+(τ)|𝑑τ=limτη2ln[1+(ωcτ)2],\displaystyle\int_{0}^{\infty}\tau|\mathcal{B}_{+-}(\tau)|d\tau=\lim_{\tau\rightarrow\infty}\frac{\eta}{2}\ln[1+(\omega_{c}\tau)^{2}]\ , (205)

which diverges. While this diverging upper bound does not prove that the error δ\delta itself diverges, it does suggest that this is indeed the case and that hence the approximation of replacing ρ(tτ)\rho(t-\tau) by ρ(t)\rho(t) is inaccurate. Indeed, the exact solution exhibits excited state population oscillations instead of purely Markovian exponential decay for all values of the parameters of the Ohmic density.

A similar situation arises for the triangular spectral density J3(ω)J_{3}(\omega). We find, numerically, that the integral 0τn|+(τ)|𝑑τ\int_{0}^{\infty}\tau^{n}|\mathcal{B}_{+-}(\tau)|d\tau [recall Eq. 128] diverges for n>0n>0, and therefore the Markov approximation is inadequate.

For a rigorous error bound, see Ref. [6].

Appendix G Proof of Eq. 119

Recall that Γαβ(ω)0𝑑ταβ(τ)eiωτ\Gamma_{\alpha\beta}(\omega)\equiv\int_{0}^{\infty}d\tau\mathcal{B}_{\alpha\beta}(\tau)e^{i\omega\tau} and HSB=σ+B++σBH_{SB}=\sigma_{+}\otimes B_{+}+\sigma_{-}\otimes B_{-} and B+=BB_{+}^{\dagger}=B_{-}. Thus, if αβ\alpha\neq\beta, Bα=BβB_{\alpha}=B^{\dagger}_{\beta}. It follows that

βα(τ)\displaystyle\mathcal{B}_{\beta\alpha}^{*}(\tau) =Tr[(ρBBβ(τ)Bα)]\displaystyle=\mathrm{Tr}\left[\left(\rho_{B}B_{\beta}(\tau)B_{\alpha}\right)^{\dagger}\right] (206a)
=Tr[ρBUB(τ)BαUB(τ)Bβ]\displaystyle=\mathrm{Tr}\left[\rho_{B}U_{B}(\tau)B^{\dagger}_{\alpha}U_{B}^{\dagger}(\tau)B^{\dagger}_{\beta}\right] (206b)
=Tr[ρBUB(τ)BβUB(τ)Bα]=βα(τ).\displaystyle=\mathrm{Tr}\left[\rho_{B}U_{B}(\tau)B_{\beta}U_{B}^{\dagger}(\tau)B_{\alpha}\right]=\mathcal{B}_{\beta\alpha}(-\tau)\ . (206c)

Hence, ±(τ)=±(τ)\mathcal{B}_{\pm\mp}^{*}(\tau)=\mathcal{B}_{\pm\mp}(-\tau) [as well as ±±(τ)=(τ)\mathcal{B}^{*}_{\pm\pm}(\tau)=\mathcal{B}_{\mp\mp}(-\tau), though we don’t use this result], which yields Γ±(ω)=0𝑑τ±(τ)eiωτ\Gamma^{*}_{{\pm\mp}}(\omega)=\int_{0}^{\infty}d\tau\mathcal{B}_{{\pm\mp}}(-\tau)e^{-i\omega\tau}.

Appendix H Simplification of the Lamb shift expressions for the RWA-LE

H.1 J2(ω)=ηωeω/ωcJ_{2}(\omega)=\eta\omega e^{-\omega/\omega_{c}}

Here we derive Eq. 126b. Our starting point is Eq. 126a, which we write as

Γ+(Ω0)/(ηωc)=0d(ωcτ)eiΩ0τ(1+iωcτ)2\displaystyle\Gamma_{+-}(\Omega_{0})/(\eta\omega_{c})=\int_{0}^{\infty}d(\omega_{c}\tau)\frac{e^{i\Omega_{0}\tau}}{(1+i\omega_{c}\tau)^{2}} (207a)
=0eiαxdx(1+ix)2\displaystyle\quad=\int_{0}^{\infty}\frac{e^{i\alpha x}dx}{(1+ix)^{2}} (207b)
=eiαxdx(1+ix)20eiαxdx(1+ix)2,\displaystyle\quad=\int_{-\infty}^{\infty}\frac{e^{i\alpha x}dx}{(1+ix)^{2}}-\int_{-\infty}^{0}\frac{e^{i\alpha x}dx}{(1+ix)^{2}}\ , (207c)

where α=Ω0/ωc\alpha=\Omega_{0}/\omega_{c} and x=ωcτx=\omega_{c}\tau.

The complex exponential integral is defined as follows [46]:

E1(z)zeuu𝑑u=zeuu𝑑u,\displaystyle\text{E}_{1}(z)\equiv\int_{z}^{\infty}\frac{e^{-u}}{u}du=-\int_{-\infty}^{-z}\frac{e^{u}}{u}du\ , (208)

where z=x+iyz=x+iy and |arg(z)|π/2|\arg(z)|\leq\pi/2. The following property holds for y>0y>0:

E1(y)=Ei(y)+iπ,\displaystyle-\text{E}_{1}(-y)=\operatorname{Ei}(y)+i\pi\ , (209)

where the real exponential integral Ei\operatorname{Ei} was defined in Eq. 43c.

The first integral on the right hand side of Eq. 207c can be computed via the residue theorem:

eiαxdx(1+ix)2=eiαxdx(xi)2=2πi(iαeα),\displaystyle\int_{-\infty}^{\infty}\frac{e^{i\alpha x}dx}{(1+ix)^{2}}=-\int_{-\infty}^{\infty}\frac{e^{i\alpha x}dx}{(x-i)^{2}}=-2\pi i(i\alpha e^{-\alpha})\ , (210)

since the second order pole of the analytic function is at x=ix=i, so that the residue is:

Res=ddxeiαx|x=i=iαeα.\displaystyle\text{Res}=\frac{d}{dx}e^{i\alpha x}\bigg{|}_{x=i}=i\alpha e^{-\alpha}\ . (211)

For the second integral on the right side of Eq. 207c we use a change of variable z=xiz=x-i and integrate by parts:

0eiαxdx(1+ix)2=0eiαxdx(xi)2=eαieiαzdzz2\displaystyle\int_{-\infty}^{0}\frac{e^{i\alpha x}dx}{(1+ix)^{2}}=-\int_{-\infty}^{0}\frac{e^{i\alpha x}dx}{(x-i)^{2}}=-e^{-\alpha}\int_{-\infty}^{-i}\frac{e^{i\alpha z}dz}{z^{2}} (212a)
=eα[eiαzz|i+iαieiαzz𝑑z]\displaystyle=-e^{-\alpha}\left[-\frac{e^{i\alpha z}}{z}\bigg{|}_{-\infty}^{-i}+i\alpha\int_{-\infty}^{-i}\frac{e^{i\alpha z}}{z}dz\right] (212b)
=iiαeαieiαzz𝑑z=iiαeααeuu𝑑u.\displaystyle=i-i\alpha e^{-\alpha}\int_{-\infty}^{-i}\frac{e^{i\alpha z}}{z}dz=i-i\alpha e^{-\alpha}\int_{-\infty}^{\alpha}\frac{e^{u}}{u}du\ . (212c)

Using the complex exponential integral Eq. 208 alongside Eq. 209, we have:

0eiαxdx(1+ix)2\displaystyle\int_{-\infty}^{0}\frac{e^{i\alpha x}dx}{(1+ix)^{2}} =iiαeα(E1(α))\displaystyle=i-i\alpha e^{-\alpha}(-\text{E}_{1}(-\alpha)) (213a)
=παeα+iiαeαEi(α).\displaystyle=\pi\alpha e^{-\alpha}+i-i\alpha e^{-\alpha}\operatorname{Ei}(\alpha)\ . (213b)

Hence, the integral of interest, Eq. 207b, becomes

0eiαxdx(1+ix)2\displaystyle\int_{0}^{\infty}\frac{e^{i\alpha x}dx}{(1+ix)^{2}} =παeαi+iαeαEi(α)\displaystyle=\pi\alpha e^{-\alpha}-i+i\alpha e^{-\alpha}\operatorname{Ei}(\alpha)
=i+αeα(π+iEi(α)).\displaystyle=-i+\alpha e^{-\alpha}(\pi+i\operatorname{Ei}(\alpha))\ . (214)

Collecting these results we obtain Eq. 126b.

H.2 J3(ω)=ηωΘ(ωcω)J_{3}(\omega)=\eta\omega\Theta(\omega_{c}-\omega)

Here we derive Eq. 130. Our starting point is Eq. 129, which we write as

Γ+(Ω0)/(ηωc)=0𝑑xeix(1+ix)1x2eiαx,\displaystyle\Gamma_{+-}(\Omega_{0})/(\eta\omega_{c})=\int_{0}^{\infty}dx\frac{e^{-ix}(1+ix)-1}{x^{2}}e^{i\alpha x}\ , (215)

where, again, α=Ω0/ωc\alpha=\Omega_{0}/\omega_{c} and x=ωcτx=\omega_{c}\tau.

We start from the following indefinite integral, solved via integration by parts:

eiuxx2𝑑x=eiuxx+iueiuxx𝑑x.\displaystyle\int\frac{e^{iux}}{x^{2}}dx=-\frac{e^{iux}}{x}+iu\int\frac{e^{iux}}{x}dx\ . (216)

Hence the integral in Eq. 215 is:

=0ei(α1)xx2𝑑x0eiαxx2𝑑x+i0ei(α1)xx𝑑x\displaystyle=\int_{0}^{\infty}\frac{e^{i(\alpha-1)x}}{x^{2}}dx-\int_{0}^{\infty}\frac{e^{i\alpha x}}{x^{2}}dx+i\int_{0}^{\infty}\frac{e^{i(\alpha-1)x}}{x}dx (217a)
=eiαx(1eix)x|0iα0eiαx(1eix)x𝑑x\displaystyle=\left.\frac{e^{i\alpha x}(1-e^{-ix})}{x}\right|_{0}^{\infty}-i\alpha\int_{0}^{\infty}\frac{e^{i\alpha x}(1-e^{-ix})}{x}dx (217b)
=i+2α0ei(2α1)usinc(u)𝑑u\displaystyle=-i+2\alpha\int_{0}^{\infty}e^{i(2\alpha-1)u}\operatorname{sinc}(u)du (217c)

To compute the last integral, we consider the real and imaginary parts separately. The real part is:

0cos(2α1)usinuu𝑑u\displaystyle\int_{0}^{\infty}\cos(2\alpha-1)u\frac{\sin u}{u}du (218a)
={0sin2αu+sin(22α)u2u𝑑u=π20<α<10sin2αusin(2α2)u2u𝑑u=0α>1\displaystyle=\begin{cases}\int_{0}^{\infty}\frac{\sin 2\alpha u+\sin(2-2\alpha)u}{2u}du=\frac{\pi}{2}&0<\alpha<1\\ \int_{0}^{\infty}\frac{\sin 2\alpha u-\sin(2\alpha-2)u}{2u}du=0&\alpha>1\end{cases} (218b)

The imaginary part is:

0sin(2α1)usinuu𝑑u\displaystyle\int_{0}^{\infty}\sin(2\alpha-1)u\frac{\sin u}{u}du (219a)
=0cos(2α2)ucos2αu2u𝑑u\displaystyle\quad=\int_{0}^{\infty}\frac{\cos(2\alpha-2)u-\cos 2\alpha u}{2u}du (219b)
={12ln(1α1)0<α<112ln(11α)α>1.\displaystyle\quad=\begin{cases}-\frac{1}{2}\ln\left(\frac{1}{\alpha}-1\right)&0<\alpha<1\\ -\frac{1}{2}\ln\left(1-\frac{1}{\alpha}\right)&\alpha>1\end{cases}\ . (219c)

Therefore the integral in Eq. 215 is

Γ+(Ω0)ηωc={παi(1+αln(1α1))0<α<1i(1+αln(11α))α>1.\displaystyle\frac{\Gamma_{+-}(\Omega_{0})}{\eta\omega_{c}}=\begin{cases}\pi\alpha-i\left(1+\alpha\ln\left(\frac{1}{\alpha}-1\right)\right)&0<\alpha<1\\ -i\left(1+\alpha\ln\left(1-\frac{1}{\alpha}\right)\right)&\alpha>1\end{cases}\ . (220)

Collecting these results and using Eq. 121 we obtain Eq. 130.

Appendix I Breakdown of the TCL approximation

For strong coupling or a small gap Ω0<ωc\Omega_{0}<\omega_{c}, as in Figs. 6 and 8, we observe that the TCL approximation is accurate for short times but fails to capture the subsequent oscillatory behavior. For these cases the operator IΣI-\Sigma in Eq. 141 is not invertible. For example, in the case of the impulse spectral density J1J_{1}, there is a common time t0t_{0} where the excited state population ρ11(t0)=|c1(t0)|2=0\rho_{11}(t_{0})=|c_{1}(t_{0})|^{2}=0 independently from the initial condition. This can be seen directly from Eq. 38 or from Eq. 40b when γ\gamma diverges. The time t0t_{0} is the minimum time tnt_{n} such that

tn=2|δ|(arctan|δ||Ω0ωc|+nπ),t_{n}=\frac{2}{|\delta|}\left(\arctan\frac{|\delta|}{|\Omega_{0}-\omega_{c}|}+n\pi\right), (221)

where nn is and integer and δ\delta is given in Eq. 39. Since TCL is a time-local master equation, it is impossible to invert the evolution for tt0t\geq t_{0} back to its initial condition. This implies that the TCL gives inconsistent result when the exact solution for the population vanishes.

A similar argument can be given for the triangular spectral density J3J_{3} as shown in Fig. 8. In contrast, this argument is not valid for the infinite range Ohmic spectral density J2J_{2}, where the population decreases but only vanishes as tt\to\infty.

References

  • Alicki and Lendi [2007] R. Alicki and K. Lendi, Quantum Dynamical Semigroups and Applications (Springer Science & Business Media, 2007).
  • Breuer and Petruccione [2002a] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, Oxford, 2002).
  • Rivas and Huelga [2012] A. Rivas and S. F. Huelga, Open Quantum Systems: An Introduction, SpringerBriefs in Physics (Springer-Verlag, Berlin Heidelberg, 2012).
  • Lindblad [1976] G. Lindblad, On the generators of quantum dynamical semigroups, Comm. Math. Phys. 48, 119 (1976).
  • Gorini et al. [1976] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Completely positive dynamical semigroups of N{N}-level systems, J. Math. Phys. 17, 821 (1976).
  • Mozgunov and Lidar [2020] E. Mozgunov and D. Lidar, Completely positive master equation for arbitrary driving and small level spacing, Quantum 4, 227 (2020).
  • Krovi et al. [2007] H. Krovi, O. Oreshkov, M. Ryazanov, and L. D., Phys. Rev. A. 76, 052117 (2007).
  • Jing and Wu [2018] J. Jing and L.-A. Wu, Decoherence and control of a qubit in spin baths: an exact master equation study, Scientific Reports 8, 1471 (2018).
  • Jaynes and Cummings [1963] E. T. Jaynes and F. W. Cummings, Comparison of quantum and semiclassical radiation theories with application to the beam maser, Proceedings of the IEEE 51, 89 (1963).
  • Garraway [1997] B. M. Garraway, Nonperturbative decay of an atomic system in a cavity, Physical Review A 55, 2290 (1997).
  • Carmichael [1993] H. Carmichael, An open systems approach to quantum optics, Lecture notes in physics No. m18 (Springer-Verlag, Berlin, 1993).
  • Koch et al. [2007a] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Charge-insensitive qubit design derived from the cooper pair box, Phys. Rev. A 76, 042319 (2007a).
  • Leppäkangas et al. [2018] J. Leppäkangas, J. Braumüller, M. Hauck, J.-M. Reiner, I. Schwenk, S. Zanker, L. Fritz, A. Ustinov, M. Weides, and M. Marthaler, Phys. Rev. A 97, 052321 (2018).
  • Stehli et al. [2023] A. Stehli, J. D. Brehm, T. Wolz, A. Schneider, H. Rotzinger, M. Weides, and A. V. Ustinov, Quantum emulation of the transient dynamics in the multistate landau-zener model, npj Quantum Information 9, 61 (2023).
  • Breuer et al. [1999] H.-P. Breuer, B. Kappler, and F. Petruccione, Stochastic wave-function method for non-markovian quantum master equations, Physical Review A 59, 1633 (1999).
  • Vacchini and Breuer [2010] B. Vacchini and H.-P. Breuer, Exact master equations for the non-markovian decay of a qubit, Physical Review A 81, 042103 (2010).
  • Koch et al. [2007b] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Charge-insensitive qubit design derived from the Cooper pair box, Physical Review A 76, 042319 (2007b).
  • Majenz et al. [2013] C. Majenz, T. Albash, H.-P. Breuer, and D. A. Lidar, Coarse graining can beat the rotating-wave approximation in quantum markovian master equations, Phys. Rev. A 88, 012103 (2013).
  • Roccati and Cilluffo [2024] F. Roccati and D. Cilluffo, Controlling markovianity with chiral giant atoms (2024), arXiv:2402.15556 [quant-ph] .
  • Ahmadi et al. [2024] B. Ahmadi, R. R. Rodríguez, R. Alicki, and M. Horodecki, Approximation scheme and non-hermitian renormalization for the description of atom-field-system evolution, Phys. Rev. A 109, 012408 (2024).
  • Leggett et al. [1987] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Dynamics of the dissipative two-state system, Rev. Mod. Phys. 59, 1 (1987).
  • Rivas et al. [2010] Á. Rivas, S. F. Huelga, and M. B. Plenio, Entanglement and non-markovianity of quantum evolutions, Physical review letters 105, 050403 (2010).
  • Lidar et al. [2001] D. A. Lidar, Z. Bihary, and K. Whaley, From completely positive maps to the quantum Markovian semigroup master equation, Chem. Phys. 268, 35 (2001).
  • Bacon et al. [1999] D. Bacon, D. A. Lidar, and K. B. Whaley, Robustness of decoherence-free subspaces for quantum computation, Phys. Rev. A 60, 1944 (1999).
  • Gaudin [1960] M. Gaudin, Une démonstration simplifiée du théorème de wick en mécanique statistique, Nuclear Physics 15, 89 (1960).
  • Fogedby [2022] H. C. Fogedby, Field-theoretical approach to open quantum systems and the Lindblad equation, Physical Review A 106, 022205 (2022).
  • Van Kampen [1974a] N. G. Van Kampen, A cumulant expansion for stochastic linear differential equations. i, Physica 74, 215 (1974a).
  • Van Kampen [1974b] N. G. Van Kampen, A cumulant expansion for stochastic linear differential equations. ii, Physica 74, 239 (1974b).
  • Breuer and Petruccione [2002b] H. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, 2002).
  • Ban [2010] M. Ban, Nonequilibrium dynamics of the dispersive Jaynes–Cummings model by a non-Markovian quantum master equation, Journal of Physics A: Mathematical and Theoretical 43, 335305 (2010).
  • Shabani and Lidar [2005] A. Shabani and D. A. Lidar, Completely positive post-markovian master equation via a measurement approach, Physical Review A 71, 020101 (2005).
  • Zhang et al. [2022] H. Zhang, B. Pokharel, E. M. Levenson-Falk, and D. Lidar, Predicting non-markovian superconducting-qubit dynamics from tomographic reconstruction, Physical Review Applied 17, 054018 (2022).
  • Alicki et al. [2006] R. Alicki, D. A. Lidar, and P. Zanardi, Internal consistency of fault-tolerant quantum error correction in light of rigorous derivations of the quantum Markovian limit, Phys. Rev. A 73, 052311 (2006).
  • Hartmann et al. [2017] M. Hartmann, D. Poletti, M. Ivanchenko, S. Denisov, and P. Hänggi, Asymptotic floquet states of open quantum systems: the role of interaction, New Journal of Physics 19, 083011 (2017).
  • Haddadfarshi et al. [2015] F. Haddadfarshi, J. Cui, and F. Mintert, Completely positive approximate solutions of driven open quantum systems, Physical Review Letters 114, 130402 (2015).
  • Yamaguchi et al. [2017] M. Yamaguchi, T. Yuge, and T. Ogawa, Markovian quantum master equation beyond adiabatic regime, Physical Review E 95, 012136 (2017).
  • Dann et al. [2018] R. Dann, A. Levy, and R. Kosloff, Time-dependent markovian quantum master equation, Physical Review A 98, 052129 (2018).
  • Meglio et al. [2023] G. D. Meglio, M. B. Plenio, and S. F. Huelga, Time dependent markovian master equation beyond the adiabatic limit (2023), arXiv:2304.06166 [quant-ph] .
  • Nathan and Rudner [2020] F. Nathan and M. S. Rudner, Universal lindblad equation for open quantum systems, Physical Review B 102, 115109 (2020).
  • Davidović [2020] D. Davidović, Completely Positive, Simple, and Possibly Highly Accurate Approximation of the Redfield Equation, Quantum 4, 326 (2020).
  • Davidović [2022] D. Davidović, Geometric-arithmetic master equation in large and fast open quantum systems, Journal of Physics A: Mathematical and Theoretical 55, 455301 (2022).
  • Winczewski et al. [2021] M. Winczewski, A. Mandarino, M. Horodecki, and R. Alicki, Bypassing the intermediate times dilemma for open quantum system (2021), arXiv:2106.05776 [quant-ph] .
  • Albash et al. [2012] T. Albash, S. Boixo, D. A. Lidar, and P. Zanardi, Quantum adiabatic markovian master equations, New J. of Phys. 14, 123016 (2012).
  • Yip et al. [2018] K. W. Yip, T. Albash, and D. A. Lidar, Quantum trajectories for time-dependent adiabatic master equations, Physical Review A 97, 022116 (2018).
  • Campos Venuti and Lidar [2018] L. Campos Venuti and D. A. Lidar, Error reduction in quantum annealing using boundary cancellation: Only the end matters, Phys. Rev. A 98, 022315 (2018).
  • Corrington [1961] M. S. Corrington, Applications of the Complex Exponential Integral, Mathematics of Computation 15, 1 (1961).