General Framework for Quantifying Dissipation Pathways in Open Quantum Systems. II. Numerical Validation and the Role of Non-Markovianity

Chang Woo Kim Department of Chemistry, Chonnam National University, Gwangju 61186, South Korea [email protected]    Ignacio Franco Department of Chemistry, University of Rochester, Rochester, New York 14627, USA Department of Physics, University of Rochester, Rochester, New York 14627, USA [email protected]
(June 8, 2024)
Abstract

In the previous paper [C. W. Kim and I. Franco, J. Chem. Phys. 160, 214111 (2024)], we developed a theory called MQME-D, which allows us to decompose the overall energy dissipation process in open quantum system dynamics into contributions by individual components of the bath when the subsystem dynamics is governed by a Markovian quantum master equation (MQME). Here, we contrast the predictions of MQME-D against the numerically exact results obtained by combining hierarchical equations of motion (HEOM) with a recently reported protocol for monitoring the statistics of the bath. Overall, MQME-D accurately captures the contributions of specific bath components to the overall dissipation while greatly reducing the computational cost as compared to exact computations using HEOM. The computations show that MQME-D exhibits errors originating from its inherent Markov approximation. We demonstrate that its accuracy can be significantly increased by incorporating non-Markovianity by exploiting time scale separations (TSS) in different components of the bath. Our work demonstrates that MQME-D combined with TSS can be reliably used to understanding how energy is dissipated in realistic open quantum system dynamics.

preprint: AIP/123-QED

I Introduction

In Paper I[1] of our two-paper series, we introduced a new theoretical framework that can be used to decompose the energy dissipated in the open quantum system dynamics into contributions from individual components of the bath. The new framework, referred to as MQME-D, targets the Markovian quantum master equations derived from the Nakajima-Zwanzig projection operator technique.[2, 3] By using it, we recovered the formulae for quantifying dissipation pathways reported in our earlier work,[4] which were developed in the context of Förster resonance energy transfer (FRET). We also derived and presented the expressions for other types of subsystem-bath interactions such as linearly coupled spin environment. In addition, the framework also allowed us to construct rigorous analytical proofs on thermodynamic principles such as energy conservation and detailed balance.

Here, we investigate the accuracy and computational efficiency of MQME-D as a method to decompose the overall dissipation during the open quantum system dynamics. An initial assessment of the strategy was done in Ref. 4 based on a comparison with a mixed-quantum classical (MQC) method.[5] However, the accuracy of the MQC method was limited by the zero-point energy leak and partial neglect of the interplay between the subsystem and bath,[6, 7] which both originate from the classical treatment of the trajectories.

Such an experience led us to develop alternative methods that do not rely on classical description of the bath. As a result, we successfully constructed a computational method[8] to quantify dissipation by individual bath modes based on numerically exact simulation methods such as hierarchical equations of motion (HEOM)[9, 10] or quasi-adiabatic propagator path integral (QUAPI).[11, 12] Using this method, we now can systematically assess the reliability of MQME-D by employing a numerically exact benchmark.

In this paper, we thoroughly examine the performance of MQME-D in various Hamiltonian models and simulation conditions with different subsystem-bath coupling strengths and temperatures. We find that, even when the subsystem dynamics is quite accurate, the dissipation predicted by MQME-D exhibits some disagreements with the benchmark calculation. Careful analysis reveals that a significant portion of the error is caused by the Markov approximation behind MQME, which does not properly reflect the response of the bath during the dynamics. We then incorporate non-Markovianity in the simulation by combining MQME-D with the time scale separation (TSS) method,[13] and demonstrate that the combination can substantially enhance the accuracy of the decomposition of the dissipation in some cases. In TSS, one separates the bath modes into slow and fast components depending on their characteristic frequencies. Only the fast component directly participates in the dynamics, while the effect of the slow component manifests as the static disorder which introduces the non-Markovianity. In the end, MQME-D with TSS (MQME-D+TSS) achieves nearly quantitative resolution of the dissipated energy in the frequency domain, demonstrating its capability to elucidate the key DOFs that govern the dynamics of realistic systems such as photosynthetic complexes or extended molecular aggregates.

The structure of the paper is as follows. Section II provides a brief overview of MQME-D and the Hamiltonian models used in this work. Section III applies the bare MQME-D to the Hamiltonian of a molecular dimer and contrasts the results against a numerically exact method. Section IV introduces the TSS strategy and benchmarks the accuracy of MQME-D+TSS by using different types of Hamiltonian models such as molecular dimer and spin-boson model. Section V presents a brief comparison between the computational efficiencies of MQME-D+TSS and numerically exact methods. Finally, Sec. VI concludes the paper by summarizing the main findings and discussing conceivable research directions for the future.

II Theoretical Background

The MQME-D theory was developed and discussed in Paper I[1]. For convenience, below we summarize the main ideas as needed to introduce the simulation strategy and subsequent discussion. First, Sec. II.1 introduces the Hamiltonian models used in this work. Then, Sec. II.2 presents the key equations of MQME-D for simulating the dissipation in the models presented in Sec. II.1.

II.1 Model Hamiltonian

To test the numerical accuracy of the theory, we study the dissipation in prototypical models of open quantum systems involving harmonic bath modes. However, as discussed in Ref. 1, the MQME-D framework is applicable to any (harmonic or anharmonic) environment as long as it is composed of independent bath degrees of freedom. The Hamiltonian of an open quantum system is written as

H^=H^sub+H^bath+H^int,^𝐻subscript^𝐻subsubscript^𝐻bathsubscript^𝐻int\hat{H}=\hat{H}_{\text{sub}}+\hat{H}_{\text{bath}}+\hat{H}_{\text{int}},over^ start_ARG italic_H end_ARG = over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT , (1)

where H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT, H^bathsubscript^𝐻bath\hat{H}_{\text{bath}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT, and H^intsubscript^𝐻int\hat{H}_{\text{int}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT are the Hamiltonian components describing the subsystem, the bath, and the subsystem-bath interaction, respectively. By denoting the subsystem states as {|A}ket𝐴\{\ket{A}\}{ | start_ARG italic_A end_ARG ⟩ }, H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT is written as

H^sub=AEA|AA|+AB<AVAB(|AB|+|BA|),subscript^𝐻subsubscript𝐴subscript𝐸𝐴ket𝐴bra𝐴subscript𝐴subscript𝐵𝐴subscript𝑉𝐴𝐵ket𝐴bra𝐵ket𝐵bra𝐴\hat{H}_{\text{sub}}=\sum_{A}E_{A}\ket{A}\bra{A}+\sum_{A}\sum_{B<A}V_{AB}(\ket% {A}\bra{B}+\ket{B}\bra{A}),over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT | start_ARG italic_A end_ARG ⟩ ⟨ start_ARG italic_A end_ARG | + ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_B < italic_A end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( | start_ARG italic_A end_ARG ⟩ ⟨ start_ARG italic_B end_ARG | + | start_ARG italic_B end_ARG ⟩ ⟨ start_ARG italic_A end_ARG | ) , (2)

where EAsubscript𝐸𝐴E_{A}italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT is the energy of the subsystem state |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩ and VABsubscript𝑉𝐴𝐵V_{AB}italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT is the coupling between the subsystem states |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩ and |Bket𝐵\ket{B}| start_ARG italic_B end_ARG ⟩. The harmonic bath modes and their interaction with the subsystem are described by

H^bath=j(p^j22+ωj2x^j22),subscript^𝐻bathsubscript𝑗superscriptsubscript^𝑝𝑗22superscriptsubscript𝜔𝑗2superscriptsubscript^𝑥𝑗22\hat{H}_{\text{bath}}=\sum_{j}\bigg{(}\frac{\hat{p}_{j}^{2}}{2}+\frac{\omega_{% j}^{2}\hat{x}_{j}^{2}}{2}\bigg{)},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( divide start_ARG over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) , (3)
H^int=A[|AA|j(ωj2dAjx^j+ωj2dAj22)].subscript^𝐻intsubscript𝐴delimited-[]tensor-productket𝐴bra𝐴subscript𝑗superscriptsubscript𝜔𝑗2subscript𝑑𝐴𝑗subscript^𝑥𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝑑𝐴𝑗22\hat{H}_{\text{int}}=\sum_{A}\bigg{[}\ket{A}\bra{A}\otimes\sum_{j}\bigg{(}-% \omega_{j}^{2}d_{Aj}\hat{x}_{j}+\frac{\omega_{j}^{2}d_{Aj}^{2}}{2}\bigg{)}% \bigg{]}.over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT [ | start_ARG italic_A end_ARG ⟩ ⟨ start_ARG italic_A end_ARG | ⊗ ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) ] . (4)

In the above, ωjsubscript𝜔𝑗\omega_{j}italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the characteristic frequency of the j𝑗jitalic_j-th bath mode, whose momentum and position are described by operators p^jsubscript^𝑝𝑗\hat{p}_{j}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and x^jsubscript^𝑥𝑗\hat{x}_{j}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, respectively. The state-dependent parameter dAjsubscript𝑑𝐴𝑗d_{Aj}italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT quantifies the strength of the subsystem-bath coupling and corresponds to the location of the minimum in the potential energy surface (PES) h^A=A|H^|Asubscript^𝐴bra𝐴^𝐻ket𝐴\hat{h}_{A}=\bra{A}\hat{H}\ket{A}over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = ⟨ start_ARG italic_A end_ARG | over^ start_ARG italic_H end_ARG | start_ARG italic_A end_ARG ⟩ along the coordinate xjsubscript𝑥𝑗x_{j}italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. Overall, Eq. (4) assumes that the bath only couples to diagonal part of H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT (Condon approximation) with linear dependence in {x^j}subscript^𝑥𝑗\{\hat{x}_{j}\}{ over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT }. The strength of the subsystem-bath interaction is collectively described by the bath spectral density (BSD), which is defined as

JAB(ω)=jωj3dAjdBj2δ(ωωj),subscript𝐽𝐴𝐵𝜔subscript𝑗superscriptsubscript𝜔𝑗3subscript𝑑𝐴𝑗subscript𝑑𝐵𝑗2𝛿𝜔subscript𝜔𝑗J_{AB}(\omega)=\sum_{j}\frac{\omega_{j}^{3}d_{Aj}d_{Bj}}{2}\delta(\omega-% \omega_{j}),italic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_ω ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_B italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_δ ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (5)

which is related to the generalized reorganization energy ΛABsubscriptΛ𝐴𝐵\Lambda_{AB}roman_Λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT via

ΛAB=jωj2dAjdBj2=0JAB(ω)ω𝑑ω.subscriptΛ𝐴𝐵subscript𝑗superscriptsubscript𝜔𝑗2subscript𝑑𝐴𝑗subscript𝑑𝐵𝑗2superscriptsubscript0subscript𝐽𝐴𝐵𝜔𝜔differential-d𝜔\Lambda_{AB}=\sum_{j}\frac{\omega_{j}^{2}d_{Aj}d_{Bj}}{2}=\int_{0}^{\infty}% \frac{J_{AB}(\omega)}{\omega}\;d\omega.roman_Λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_B italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_ω ) end_ARG start_ARG italic_ω end_ARG italic_d italic_ω . (6)

Analytical expressions of the BSD are frequently used as approximate models for describing realistic systems. One widely used form of the BSD is the Drude-Lorentz distribution expressed as

JDL(ω)=2Λπωcωω2+ωc2,subscript𝐽DL𝜔2Λ𝜋subscript𝜔c𝜔superscript𝜔2superscriptsubscript𝜔c2J_{\text{DL}}(\omega)=\frac{2\Lambda}{\pi}\frac{\omega_{\text{c}}\omega}{% \omega^{2}+\omega_{\text{c}}^{2}},italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 2 roman_Λ end_ARG start_ARG italic_π end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_ω end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (7)

which is often used to model the slow relaxation of the solvent due to the exponential form of the corresponding bath time correlation function in the high-temperature limit.[14] In Eq. (7), ΛΛ\Lambdaroman_Λ is the total reorganization energy of the bath and ωcsubscript𝜔c\omega_{\text{c}}italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT is the cutoff frequency which determines the relaxation time of the bath. Another model that will be used in this work is the Brownian oscillator whose BSD is expressed as

JBO(ω)=2Λγπ2ω02ω(ω2ω02)2+4γ2ω2.subscript𝐽BO𝜔2Λ𝛾𝜋2superscriptsubscript𝜔02𝜔superscriptsuperscript𝜔2superscriptsubscript𝜔0224superscript𝛾2superscript𝜔2J_{\text{BO}}(\omega)=\frac{2\Lambda\gamma}{\pi}\frac{2\omega_{0}^{2}\omega}{(% \omega^{2}-\omega_{0}^{2})^{2}+4\gamma^{2}\omega^{2}}.italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 2 roman_Λ italic_γ end_ARG start_ARG italic_π end_ARG divide start_ARG 2 italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω end_ARG start_ARG ( italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (8)

Here, ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the characteristic frequency of the oscillator and γ𝛾\gammaitalic_γ is the strength of the dam**. The time correlation function associated with Eq. (8) resembles the behavior of a damped harmonic oscillator,[14] which makes JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) a realistic model for describing molecular vibrations and photonic cavities.

II.2 MQME-D for a Bath of Harmonic Oscillators

According to Eq. (15) of Paper I,[1] the populations of the subsystem states in MQME are governed by a set of coupled first-order rate equations

P˙A(t)=BA[KBAPA(t)+KABPB(t)],subscript˙𝑃𝐴𝑡subscript𝐵𝐴delimited-[]subscript𝐾𝐵𝐴subscript𝑃𝐴𝑡subscript𝐾𝐴𝐵subscript𝑃𝐵𝑡\dot{P}_{A}(t)=\sum_{B\neq A}[-K_{BA}P_{A}(t)+K_{AB}P_{B}(t)],over˙ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_B ≠ italic_A end_POSTSUBSCRIPT [ - italic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) + italic_K start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_t ) ] , (9)

where PA(t)subscript𝑃𝐴𝑡P_{A}(t)italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) is the population of the state |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩, and KBAsubscript𝐾𝐵𝐴K_{BA}italic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT is the rate constant for the population transfer from |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩ to |Bket𝐵\ket{B}| start_ARG italic_B end_ARG ⟩ whose explicit expression is

KBA=2|VAB|22Re0exp(it(EBEA+ΛAA2ΛAB+ΛBB))×exp[gAA(t)+2gAB(t)gBB(t)]dt.subscript𝐾𝐵𝐴2superscriptsubscript𝑉𝐴𝐵2superscriptPlanck-constant-over-2-pi2Resuperscriptsubscript0𝑖superscript𝑡subscript𝐸𝐵subscript𝐸𝐴subscriptΛ𝐴𝐴2subscriptΛ𝐴𝐵subscriptΛ𝐵𝐵Planck-constant-over-2-pisubscript𝑔𝐴𝐴superscript𝑡2subscript𝑔𝐴𝐵superscript𝑡subscript𝑔𝐵𝐵superscript𝑡𝑑superscript𝑡\begin{split}K_{BA}&=\frac{2|V_{AB}|^{2}}{\hbar^{2}}\text{Re}\int_{0}^{\infty}% \exp\bigg{(}-\frac{it^{\prime}(E_{B}-E_{A}+\Lambda_{AA}-2\Lambda_{AB}+\Lambda_% {BB})}{\hbar}\bigg{)}\\ &\times\exp[-g_{AA}(t^{\prime})+2g_{AB}(t^{\prime})-g_{BB}(t^{\prime})]\>dt^{% \prime}.\end{split}start_ROW start_CELL italic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT end_CELL start_CELL = divide start_ARG 2 | italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG Re ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_i italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_E start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT - 2 roman_Λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ) end_ARG start_ARG roman_ℏ end_ARG ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × roman_exp [ - italic_g start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + 2 italic_g start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_g start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . end_CELL end_ROW (10)

Here, gAB(t)subscript𝑔𝐴𝐵superscript𝑡g_{AB}(t^{\prime})italic_g start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) is the line broadening function defined as

gAB(t)=10JAB(ω)[coth(βω2)1cos(ωt)ω2+isin(ωt)ωtω2]𝑑ω,subscript𝑔𝐴𝐵superscript𝑡1Planck-constant-over-2-pisuperscriptsubscript0subscript𝐽𝐴𝐵𝜔delimited-[]hyperbolic-cotangent𝛽Planck-constant-over-2-pi𝜔21𝜔superscript𝑡superscript𝜔2𝑖𝜔superscript𝑡𝜔superscript𝑡superscript𝜔2differential-d𝜔g_{AB}(t^{\prime})=\frac{1}{\hbar}\int_{0}^{\infty}J_{AB}(\omega)\bigg{[}\coth% \bigg{(}\frac{\beta\hbar\omega}{2}\bigg{)}\frac{1-\cos(\omega t^{\prime})}{% \omega^{2}}+i\>\frac{\sin(\omega t^{\prime})-\omega t^{\prime}}{\omega^{2}}% \bigg{]}\>d\omega,italic_g start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = divide start_ARG 1 end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_ω ) [ roman_coth ( divide start_ARG italic_β roman_ℏ italic_ω end_ARG start_ARG 2 end_ARG ) divide start_ARG 1 - roman_cos ( italic_ω italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_i divide start_ARG roman_sin ( italic_ω italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_ω italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] italic_d italic_ω , (11)

where β=1/kBT𝛽1subscript𝑘B𝑇\beta=1/k_{\text{B}}Titalic_β = 1 / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T is the inverse temperature. With the state populations propagated according to Eq. (10), the rate of dissipation E˙j(t)subscript˙𝐸𝑗𝑡\dot{E}_{j}(t)over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) into the j𝑗jitalic_j-th bath mode can be evaluated by

E˙j(t)=AB<A[𝒦BAjPA(t)+𝒦ABjPB(t)],subscript˙𝐸𝑗𝑡subscript𝐴subscript𝐵𝐴delimited-[]superscriptsubscript𝒦𝐵𝐴𝑗subscript𝑃𝐴𝑡superscriptsubscript𝒦𝐴𝐵𝑗subscript𝑃𝐵𝑡\dot{E}_{j}(t)=\sum_{A}\sum_{B<A}[\mathcal{K}_{BA}^{j}P_{A}(t)+\mathcal{K}_{AB% }^{j}P_{B}(t)],over˙ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_B < italic_A end_POSTSUBSCRIPT [ caligraphic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) + caligraphic_K start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_t ) ] , (12)

where the dissipation rate constants {𝒦BAj}superscriptsubscript𝒦𝐵𝐴𝑗\{\mathcal{K}_{BA}^{j}\}{ caligraphic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT } are calculated as

𝒦BAj=2|VAB|22(λAAj2λABj+λBBj)×Re0exp(it(EBEA+ΛAA2ΛAB+ΛBB))×exp[gAA(t)+2gAB(t)gBB(t)]×[cos(ωjt)icoth(βωj2)sin(ωjt)]dt.superscriptsubscript𝒦𝐵𝐴𝑗2superscriptsubscript𝑉𝐴𝐵2superscriptPlanck-constant-over-2-pi2superscriptsubscript𝜆𝐴𝐴𝑗2superscriptsubscript𝜆𝐴𝐵𝑗superscriptsubscript𝜆𝐵𝐵𝑗Resuperscriptsubscript0𝑖superscript𝑡subscript𝐸𝐵subscript𝐸𝐴subscriptΛ𝐴𝐴2subscriptΛ𝐴𝐵subscriptΛ𝐵𝐵Planck-constant-over-2-pisubscript𝑔𝐴𝐴superscript𝑡2subscript𝑔𝐴𝐵superscript𝑡subscript𝑔𝐵𝐵superscript𝑡delimited-[]subscript𝜔𝑗superscript𝑡𝑖hyperbolic-cotangent𝛽Planck-constant-over-2-pisubscript𝜔𝑗2subscript𝜔𝑗superscript𝑡𝑑superscript𝑡\begin{split}\mathcal{K}_{BA}^{j}&=\frac{2|V_{AB}|^{2}}{\hbar^{2}}(\lambda_{AA% }^{j}-2\lambda_{AB}^{j}+\lambda_{BB}^{j})\\ &\times\text{Re}\int_{0}^{\infty}\exp\bigg{(}-\frac{it^{\prime}(E_{B}-E_{A}+% \Lambda_{AA}-2\Lambda_{AB}+\Lambda_{BB})}{\hbar}\bigg{)}\\ &\times\exp[-g_{AA}(t^{\prime})+2g_{AB}(t^{\prime})-g_{BB}(t^{\prime})]\\ &\times\bigg{[}\cos(\omega_{j}t^{\prime})-i\coth\bigg{(}\frac{\beta\hbar\omega% _{j}}{2}\bigg{)}\sin(\omega_{j}t^{\prime})\bigg{]}\>dt^{\prime}.\end{split}start_ROW start_CELL caligraphic_K start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_CELL start_CELL = divide start_ARG 2 | italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_λ start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT - 2 italic_λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + italic_λ start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × Re ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_i italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_E start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT - 2 roman_Λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ) end_ARG start_ARG roman_ℏ end_ARG ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × roman_exp [ - italic_g start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + 2 italic_g start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_g start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × [ roman_cos ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_i roman_coth ( divide start_ARG italic_β roman_ℏ italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) roman_sin ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . end_CELL end_ROW (13)

In the above, λABj=ωj2dAjdBj/2superscriptsubscript𝜆𝐴𝐵𝑗superscriptsubscript𝜔𝑗2subscript𝑑𝐴𝑗subscript𝑑𝐵𝑗2\lambda_{AB}^{j}=\omega_{j}^{2}d_{Aj}d_{Bj}/2italic_λ start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_B italic_j end_POSTSUBSCRIPT / 2 is the reorganization energy associated with the j𝑗jitalic_j-th bath mode. At this point, we introduce specific Hamiltonian models which will be employed in the simulations in Sec. III and Sec. IV.

II.2.1 Local Bath Model

In this model, each bath mode only couples to a single subsystem state, and the Hamiltonian can be derived from Eqs. (1)–(4) by setting dAjdBj=0subscript𝑑𝐴𝑗subscript𝑑𝐵𝑗0d_{Aj}d_{Bj}=0italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_B italic_j end_POSTSUBSCRIPT = 0 when AB𝐴𝐵A\neq Bitalic_A ≠ italic_B. As a result, H^bathsubscript^𝐻bath\hat{H}_{\text{bath}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT and H^intsubscript^𝐻int\hat{H}_{\text{int}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT are reduced to

H^bath=AjA(p^j22+ωj2x^j22),subscript^𝐻bathsubscript𝐴subscript𝑗𝐴superscriptsubscript^𝑝𝑗22superscriptsubscript𝜔𝑗2superscriptsubscript^𝑥𝑗22\hat{H}_{\text{bath}}=\sum_{A}\sum_{j\in A}\bigg{(}\frac{\hat{p}_{j}^{2}}{2}+% \frac{\omega_{j}^{2}\hat{x}_{j}^{2}}{2}\bigg{)},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j ∈ italic_A end_POSTSUBSCRIPT ( divide start_ARG over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) , (14)
H^int=A[|AA|jA(ωj2dAjx^j+ωj2dAj22)],subscript^𝐻intsubscript𝐴delimited-[]tensor-productket𝐴bra𝐴subscript𝑗𝐴superscriptsubscript𝜔𝑗2subscript𝑑𝐴𝑗subscript^𝑥𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝑑𝐴𝑗22\hat{H}_{\text{int}}=\sum_{A}\bigg{[}\ket{A}\bra{A}\otimes\sum_{j\in A}\bigg{(% }-\omega_{j}^{2}d_{Aj}\hat{x}_{j}+\frac{\omega_{j}^{2}d_{Aj}^{2}}{2}\bigg{)}% \bigg{]},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT [ | start_ARG italic_A end_ARG ⟩ ⟨ start_ARG italic_A end_ARG | ⊗ ∑ start_POSTSUBSCRIPT italic_j ∈ italic_A end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_A italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) ] , (15)

where we have introduced the notation jA𝑗𝐴j\in Aitalic_j ∈ italic_A to express that the j𝑗jitalic_j-th bath mode exclusively belongs to |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩. For this model, the frequency-resolved rate of dissipation can be calculated by Eq. (75) of Paper I[1],

𝒟A(ω,t)=BA[𝒥BAA(ω)PA(t)+𝒥ABA(ω)PB(t)].subscript𝒟𝐴𝜔𝑡subscript𝐵𝐴delimited-[]superscriptsubscript𝒥𝐵𝐴𝐴𝜔subscript𝑃𝐴𝑡superscriptsubscript𝒥𝐴𝐵𝐴𝜔subscript𝑃𝐵𝑡\mathcal{D}_{A}(\omega,t)=\sum_{B\neq A}[\mathcal{J}_{BA}^{A}(\omega)P_{A}(t)+% \mathcal{J}_{AB}^{A}(\omega)P_{B}(t)].caligraphic_D start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) = ∑ start_POSTSUBSCRIPT italic_B ≠ italic_A end_POSTSUBSCRIPT [ caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) + caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_t ) ] . (16)

With this, 𝒟A(ω,t)dωsubscript𝒟𝐴𝜔𝑡𝑑𝜔\mathcal{D}_{A}(\omega,t)\>d\omegacaligraphic_D start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) italic_d italic_ω becomes the rate of dissipation through the frequency window [ω,ω+dω]𝜔𝜔𝑑𝜔[\omega,\omega+d\omega][ italic_ω , italic_ω + italic_d italic_ω ] at time t𝑡titalic_t for the vibrational modes coupled to molecule A𝐴Aitalic_A. For future reference, we also define the cumulative dissipation density as

A(ω,t)=0t𝒟A(ω,t)𝑑t,subscript𝐴𝜔𝑡superscriptsubscript0𝑡subscript𝒟𝐴𝜔superscript𝑡differential-dsuperscript𝑡\mathcal{E}_{A}(\omega,t)=\int_{0}^{t}\mathcal{D}_{A}(\omega,t^{\prime})\>dt^{% \prime},caligraphic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT caligraphic_D start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , (17)

which yields the total dissipated energy at t𝑡titalic_t when integrated over the frequency axis. We note that the subscript A𝐴Aitalic_A in 𝒟A(ω,t)subscript𝒟𝐴𝜔𝑡\mathcal{D}_{A}(\omega,t)caligraphic_D start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) and A(ω,t)subscript𝐴𝜔𝑡\mathcal{E}_{A}(\omega,t)caligraphic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) will be omitted when it is not needed for clarity.

The frequency-dependent profiles {𝒥BAC(ω)}superscriptsubscript𝒥𝐵𝐴𝐶𝜔\{\mathcal{J}_{BA}^{C}(\omega)\}{ caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT ( italic_ω ) } in Eq. (16) are given by

𝒥BAC(ω)=2|VAB|22JCC(ω)ωBA(ω)superscriptsubscript𝒥𝐵𝐴𝐶𝜔2superscriptsubscript𝑉𝐴𝐵2superscriptPlanck-constant-over-2-pi2subscript𝐽𝐶𝐶𝜔𝜔subscript𝐵𝐴𝜔\mathcal{J}_{BA}^{C}(\omega)=\frac{2|V_{AB}|^{2}}{\hbar^{2}}\frac{J_{CC}(% \omega)}{\omega}\mathcal{I}_{BA}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT ( italic_ω ) = divide start_ARG 2 | italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_J start_POSTSUBSCRIPT italic_C italic_C end_POSTSUBSCRIPT ( italic_ω ) end_ARG start_ARG italic_ω end_ARG caligraphic_I start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ( italic_ω ) (18)

where C𝐶Citalic_C is either A𝐴Aitalic_A or B𝐵Bitalic_B. This quantity, which was called “dissipative spectral density” in our earlier work,[4] captures how the energy dissipated by the transfer of state population from |Aket𝐴\ket{A}| start_ARG italic_A end_ARG ⟩ to |Bket𝐵\ket{B}| start_ARG italic_B end_ARG ⟩ is distributed among the bath modes coupled to |Cket𝐶\ket{C}| start_ARG italic_C end_ARG ⟩. Moreover, as JCC(ω)/ωsubscript𝐽𝐶𝐶𝜔𝜔J_{CC}(\omega)/\omegaitalic_J start_POSTSUBSCRIPT italic_C italic_C end_POSTSUBSCRIPT ( italic_ω ) / italic_ω is the density of reorganization energy along the frequency domain [Eq. (6)], the quantity BA(ω)subscript𝐵𝐴𝜔\mathcal{I}_{BA}(\omega)caligraphic_I start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ( italic_ω ) in Eq. (18) reflects the ability of the bath to induce dissipation per unit reorganization energy. For this reason, BA(ω)subscript𝐵𝐴𝜔\mathcal{I}_{BA}(\omega)caligraphic_I start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ( italic_ω ) was named as “dissipative potential” and is expressed as

BA(ω)=Re0exp(it(EBEA+ΛAA+ΛBB)gAA(t)gBB(t))×[cos(ωt)icoth(βω2)sin(ωt)]dt.subscript𝐵𝐴𝜔Resuperscriptsubscript0𝑖superscript𝑡subscript𝐸𝐵subscript𝐸𝐴subscriptΛ𝐴𝐴subscriptΛ𝐵𝐵Planck-constant-over-2-pisubscript𝑔𝐴𝐴superscript𝑡subscript𝑔𝐵𝐵superscript𝑡delimited-[]𝜔superscript𝑡𝑖hyperbolic-cotangent𝛽Planck-constant-over-2-pi𝜔2𝜔superscript𝑡𝑑superscript𝑡\begin{split}\mathcal{I}_{BA}(\omega)&=\text{Re}\int_{0}^{\infty}\exp\bigg{(}-% \frac{it^{\prime}(E_{B}-E_{A}+\Lambda_{AA}+\Lambda_{BB})}{\hbar}-g_{AA}(t^{% \prime})-g_{BB}(t^{\prime})\bigg{)}\\ &\times\bigg{[}\cos(\omega t^{\prime})-i\coth\bigg{(}\frac{\beta\hbar\omega}{2% }\bigg{)}\sin(\omega t^{\prime})\bigg{]}\>dt^{\prime}.\end{split}start_ROW start_CELL caligraphic_I start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ( italic_ω ) end_CELL start_CELL = Re ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_i italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_E start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT + roman_Λ start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ) end_ARG start_ARG roman_ℏ end_ARG - italic_g start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_g start_POSTSUBSCRIPT italic_B italic_B end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL × [ roman_cos ( italic_ω italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) - italic_i roman_coth ( divide start_ARG italic_β roman_ℏ italic_ω end_ARG start_ARG 2 end_ARG ) roman_sin ( italic_ω italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . end_CELL end_ROW (19)

An interesting consequence of Eqs. (16)–(19) is that, for two-level subsystems, the dissipation by both molecules become completely identical when JA(ω)=JB(ω)subscript𝐽𝐴𝜔subscript𝐽𝐵𝜔J_{A}(\omega)=J_{B}(\omega)italic_J start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω ) = italic_J start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ω ). This is because 𝒥BAA(ω)=𝒥BAB(ω)superscriptsubscript𝒥𝐵𝐴𝐴𝜔superscriptsubscript𝒥𝐵𝐴𝐵𝜔\mathcal{J}_{BA}^{A}(\omega)=\mathcal{J}_{BA}^{B}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) = caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ω ) and 𝒥ABA(ω)=𝒥ABB(ω)superscriptsubscript𝒥𝐴𝐵𝐴𝜔superscriptsubscript𝒥𝐴𝐵𝐵𝜔\mathcal{J}_{AB}^{A}(\omega)=\mathcal{J}_{AB}^{B}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) = caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_ω ) under such a condition [Eq. (18)], and inserting these relations in Eq. (16) gives 𝒟A(ω,t)=𝒟B(ω,t)subscript𝒟𝐴𝜔𝑡subscript𝒟𝐵𝜔𝑡\mathcal{D}_{A}(\omega,t)=\mathcal{D}_{B}(\omega,t)caligraphic_D start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω , italic_t ) = caligraphic_D start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ω , italic_t ). The validity of this corollary will be scrutinized in Sec. III.4.

II.2.2 Spin-Boson Model

In this model, a two-level subsystem is coupled to a single group of bath modes in an anti-correlated fashion. If we denote the two subsystem states as |+ket\ket{+}| start_ARG + end_ARG ⟩ and |ket\ket{-}| start_ARG - end_ARG ⟩, the Hamiltonian components H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT and H^intsubscript^𝐻int\hat{H}_{\text{int}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT are written as

H^sub=E2σ^z+Vσ^x,subscript^𝐻sub𝐸2subscript^𝜎𝑧𝑉subscript^𝜎𝑥\hat{H}_{\text{sub}}=\frac{E}{2}\hat{\sigma}_{z}+V\hat{\sigma}_{x},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT = divide start_ARG italic_E end_ARG start_ARG 2 end_ARG over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_V over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , (20)
H^int=σ^z[j(ωj2djx^j+ωj2dj22)],subscript^𝐻inttensor-productsubscript^𝜎𝑧delimited-[]subscript𝑗superscriptsubscript𝜔𝑗2subscript𝑑𝑗subscript^𝑥𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝑑𝑗22\hat{H}_{\text{int}}=\hat{\sigma}_{z}\otimes\bigg{[}\sum_{j}\bigg{(}-\omega_{j% }^{2}d_{j}\hat{x}_{j}+\frac{\omega_{j}^{2}d_{j}^{2}}{2}\bigg{)}\bigg{]},over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT = over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⊗ [ ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( - italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) ] , (21)

where we have used the Pauli spin operators σ^x=|+|+|+|subscript^𝜎𝑥ketbraketbra\hat{\sigma}_{x}=\ket{+}\bra{-}+\ket{-}\bra{+}over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = | start_ARG + end_ARG ⟩ ⟨ start_ARG - end_ARG | + | start_ARG - end_ARG ⟩ ⟨ start_ARG + end_ARG | and σ^z=|++|||subscript^𝜎𝑧ketbraketbra\hat{\sigma}_{z}=\ket{+}\bra{+}-\ket{-}\bra{-}over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = | start_ARG + end_ARG ⟩ ⟨ start_ARG + end_ARG | - | start_ARG - end_ARG ⟩ ⟨ start_ARG - end_ARG |.

III Markovian Dissipation

We now assess the accuracy of MQME-D by comparing its predictions with those obtained by using the numerically exact HEOM method.[9, 10] Our aim is to conduct a quantitative and systematic study on the extent to which the second-order perturbation theory and Markov approximation underlying MQME-D affect its reliability. In analogy with MQME-D, we call the decomposition of the dissipation into individual bath components based on HEOM as “HEOM-D.” In HEOM-D, the dissipation induced by a particular bath mode is indirectly elucidated by introduction of an additional “probe mode,” whose Hamiltonian closely resembles that of the mode we are interested in.[8] The introduction of the probe mode enables us to extract the dissipation by using the conventional protocol based on the extended subsystem,[15, 16, 17] without disturbing the analytical BSD required to construct HEOM. Details of this exact method are included in Appendix A.

For definitiveness, we focus on the dynamics of the local bath model (Sec. II.2.1) whose subsystem just consists of two states which we will denote as |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ and |2ket2\ket{2}| start_ARG 2 end_ARG ⟩. This model can be used to describe the dynamics within the single-excitation manifold of a molecular dimer, by map** |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ (|2ket2\ket{2}| start_ARG 2 end_ARG ⟩) onto the instance in which molecule 1 (2) is in its excited state while molecule 2 (1) remains in the ground state. When the electronic coupling arises from weak dipole-dipole interaction between the chromophores, the migration of excitation between them is referred to as Förster resonance energy transfer (FRET).[18]

III.1 Simulation Procedure for Molecular Dimer Model

We use Planck atomic unit system for which =kB=1Planck-constant-over-2-pisubscript𝑘B1\hbar=k_{\text{B}}=1roman_ℏ = italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT = 1. Table 1 summarizes the simulation conditions employed for the molecular dimer model. We examine the accuracy of MQME-D by either varying ΛΛ\Lambdaroman_Λ while kee** the temperature constant as T=1.0𝑇1.0T=1.0italic_T = 1.0 (conditions (i)–(iv) in Table 1) or varying T𝑇Titalic_T for a fixed reorganization energy Λ=0.2Λ0.2\Lambda=0.2roman_Λ = 0.2 (conditions (ii), (v), and (vi) in Table 1). From now on, we will refer these two sets of simulation conditions as “const-T𝑇Titalic_T” and “const-ΛΛ\Lambdaroman_Λ” series, respectively. For each simulation condition in Table 1, we also studied the effect of the energy gap ΔE=E1E2Δ𝐸subscript𝐸1subscript𝐸2\Delta E=E_{1}-E_{2}roman_Δ italic_E = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT on the accuracy by varying ΔEΔ𝐸\Delta Eroman_Δ italic_E as 0, 1 and 2, generating a total of 18 different simulation conditions. We fix V12=0.25subscript𝑉120.25V_{12}=0.25italic_V start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = 0.25 for the electronic coupling and ωc=0.5subscript𝜔c0.5\omega_{\text{c}}=0.5italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 0.5 for the cutoff frequency of the Drude-Lorentz BSDs [Eq. (7)] that couple to the chromophores. We also assume that the electronic excitation is initially localized at molecule 1 unless noted otherwise.

For MQME and MQME-D simulations, each BSD was discretized into 2000 harmonic oscillator modes by using the scheme described in Appendix B, which was originally reported in Ref.19. We used ωmax=15subscript𝜔max15\omega_{\text{max}}=15italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = 15 as the upper limit of frequency, which recovered 97.9% of the pristine reorganization energy of the analytical BSD. The time integrals required to calculate electronic [Eq. (10)] and dissipation [Eq. (18)] rate constants were evaluated by using the trapezoidal method with an integration grid size of 0.01 and the upper limit of integration of 5×1035superscript1035\times 10^{3}5 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. The rate equations for electronic populations [Eq. (9)] and dissipation [Eq. (16)] were propagated by using the fourth-order Runge-Kutta method with the time step of 0.01.

Numerically accurate benchmarks for dissipation were extracted by combining HEOM with the approach described in Ref. 8, along with the efficient low-temperature correction scheme recently reported in Ref. 20. Table 1 lists the number of hierarchy tiers Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT, the number of Matsubara low-temperature correction terms NMatsusubscript𝑁MatsuN_{\text{Matsu}}italic_N start_POSTSUBSCRIPT Matsu end_POSTSUBSCRIPT, and Huang-Rhys (H-R) factor of the probe mode sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT [Eq. (34)] used for the individual simulation conditions. For HEOM calculations, larger Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT and NMatsusubscript𝑁MatsuN_{\text{Matsu}}italic_N start_POSTSUBSCRIPT Matsu end_POSTSUBSCRIPT are required for stronger subsystem-bath interaction and lower temperature, respectively, while sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT is exclusively used for HEOM-D and must be small enough to satisfy the weak-coupling limit [Eq. (33)]. We scanned the frequency of the probe mode in the range of [0.1,3.0]0.13.0[0.1,3.0][ 0.1 , 3.0 ] with a constant spacing of 0.05. For all data points ω0.2𝜔0.2\omega\geq 0.2italic_ω ≥ 0.2, the number of vibrational quantum states used for describing the probe mode[8] was determined to make the initial bath density captures 99.9% of the total Boltzmann population. In the case of ω<0.2𝜔0.2\omega<0.2italic_ω < 0.2, the above criterion was slightly relaxed to 99.0% to cope with the steeply increasing computational burden of HEOM-D as ω𝜔\omegaitalic_ω decreases (Sec. V).

The steady state limits (t𝑡t\rightarrow\inftyitalic_t → ∞) were practically chosen as some finite time tsimsubscript𝑡simt_{\text{sim}}italic_t start_POSTSUBSCRIPT sim end_POSTSUBSCRIPT, which is differently defined for each simulation condition by visually inspecting the evolution of electronic population. The reduced density matrix (RDM) of the subsystem and the auxiliary density matrices (ADMs) in HEOM are propagated by using the adaptive RKF45 integrator.[21] The error function used for tuning the time step was determined as how much the trace of the RDM deviates from unity, which must be maintained as nought in the exact dynamics. To further improve the stability of the calculation near the steady state, we also prevented the time step from increasing beyond a pre-determined maximum value ΔtmaxΔsubscript𝑡max\Delta t_{\text{max}}roman_Δ italic_t start_POSTSUBSCRIPT max end_POSTSUBSCRIPT, which is also listed in Table 1 for all simulation conditions.

Table 1: Summary of the simulation conditions used for the molecular dimer model. The quantities ΛΛ\Lambdaroman_Λ and T𝑇Titalic_T define the subsystem-bath interaction. The other parameters specify the HEOM procedure. Each of the 6 simulation conditions in the table was combined with three different values of the energy gap ΔE=E1E2Δ𝐸subscript𝐸1subscript𝐸2\Delta E=E_{1}-E_{2}roman_Δ italic_E = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to yield 18 different conditions in total.
Simulation condition (i) (ii) (iii) (iv) (v) (vi)
Reorganization energy (ΛΛ\Lambdaroman_Λ) 0.05 0.2 1.0 2.0 0.2 0.2
Temperature (T𝑇Titalic_T) 1.0 1.0 1.0 1.0 0.5 0.25
Maximum time step (ΔtmaxΔsubscript𝑡max\Delta t_{\text{max}}roman_Δ italic_t start_POSTSUBSCRIPT max end_POSTSUBSCRIPT) 0.02 0.1 0.05 0.05 0.1 0.1
Number of hierarchy tiers (Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT) 4 7 10 13 7 7
Number of Matsubara terms (NMatsusubscript𝑁MatsuN_{\text{Matsu}}italic_N start_POSTSUBSCRIPT Matsu end_POSTSUBSCRIPT) 30 30 30 30 100 100
H-R factor of the probe mode (sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT) 2×1062superscript1062\times 10^{-6}2 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT

When the dynamics approaches the steady state, we observe that the total amount of dissipated energy computed with HEOM-D artificially increases at a constant rate for HEOM-D simulations even after the electronic dynamics has become stationary, violating energy conservation. Such a spurious behavior tends to become more severe when ΛΛ\Lambdaroman_Λ increases. A detailed numerical analysis of the drift is presented in the Supplementary Material for the case of Λ=1.0Λ1.0\Lambda=1.0roman_Λ = 1.0 [condition (iii)]. As shown, the drift cannot be eliminated by simply increasing Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT or NMatsusubscript𝑁MatsuN_{\text{Matsu}}italic_N start_POSTSUBSCRIPT Matsu end_POSTSUBSCRIPT, or decreasing ΔtmaxΔsubscript𝑡max\Delta t_{\text{max}}roman_Δ italic_t start_POSTSUBSCRIPT max end_POSTSUBSCRIPT or sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT. We therefore conclude that the dissipation does not arise from insufficient numerical convergence. In our results, we removed this artificial drift by applying a procedure similar to the one described in Ref. 4 [Eq. (S1) in the Supplementary Information]. This correction scheme assumes the linearity of the drift throughout the entire simulation, whose validity is also discussed in the Supplementary Material. Figure S4 of the Supplementary Material demonstrates that the corrected dissipation is robust to the choice of the simulation parameters, as long as the state population achieves convergence.

III.2 Electronic Dynamics

Refer to caption
Figure 1: Comparison between the dynamics of the state population represented as the population inversion σ^zdelimited-⟨⟩subscript^𝜎𝑧\langle\hat{\sigma}_{z}\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩, calculated by MQME (solid black) and HEOM (dashed purple) for the const-T𝑇Titalic_T series (T=1.0𝑇1.0T=1.0italic_T = 1.0). The results calculated by combining MQME with the time scale separation (gray) is also presented, which will be discussed in Sec. IV.2.1.
Refer to caption
Figure 2: Same as in Fig. 1, but for the const-ΛΛ\Lambdaroman_Λ series (Λ=0.2Λ0.2\Lambda=0.2roman_Λ = 0.2).

We first assess the reliability of MQME on describing the evolution of the state populations, which is the prerequisite for accurately calculating the dissipation. Before we present the results from the simulation, it is worthwhile to address how the applicability of MQME depends on the parameters related to the dynamics. According to Sec. II A of Paper I,[1] MQME is based on the assumption that the inter-state coupling is sufficiently weak so that it can be treated to second-order in perturbation. This implies that MQME will become accurate either when |ΔE||V12|much-greater-thanΔ𝐸subscript𝑉12|\Delta E|\gg|V_{12}|| roman_Δ italic_E | ≫ | italic_V start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT | in H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT, or when H^intsubscript^𝐻int\hat{H}_{\text{int}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT int end_POSTSUBSCRIPT induces strong thermal fluctuation arising from high T𝑇Titalic_T or large ΛΛ\Lambdaroman_Λ.[13]

We now analyze the accuracy of MQME based on the results obtained from the simulations. Figures 1 and 2 visualize the time-dependent population inversion σ^z(t)=P1(t)P2(t)delimited-⟨⟩subscript^𝜎𝑧𝑡subscript𝑃1𝑡subscript𝑃2𝑡\langle\hat{\sigma}_{z}(t)\rangle=P_{1}(t)-P_{2}(t)⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_t ) ⟩ = italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) - italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) calculated for the const-T𝑇Titalic_T and const-ΛΛ\Lambdaroman_Λ series, respectively, by using both MQME and HEOM. For Fig. 1, most of the cases [Fig. 1(c)–(l)] show good agreement between MQME and HEOM results, even though MQME slightly overestimates the rate of population transfer when ΔE1Δ𝐸1\Delta E\geq 1roman_Δ italic_E ≥ 1 [Fig. 1(e)–(l)]. The most noticeable disagreement between MQME and HEOM is observed for Λ=0.05Λ0.05\Lambda=0.05roman_Λ = 0.05 and ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Fig. 1(a)], for which HEOM shows oscillations in σ^z(t)delimited-⟨⟩subscript^𝜎𝑧𝑡\langle\hat{\sigma}_{z}(t)\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_t ) ⟩ while MQME does not. Such oscillations occur when the eigenstates of H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT undergo significant delocalization due to low |ΔE|Δ𝐸|\Delta E|| roman_Δ italic_E |, and the coherence between these eigenstates can persist for relatively long time due to weak subsystem-bath interaction [Fig. 1(a) and (b)]. It is easily expected that the accuracy of MQME will deteriorate under such conditions if we recall the discussions made above. The observed lack of oscillation actually shows a fundamental limitation of the MQME, as it can only predict monotonous, exponential dynamics. This can be explicitly shown by solving the coupled differential equations for the state populations [Eq. (9)] for our dimer model. As a result, we can derive the analytical expression for the population inversion as

σ^z(t)=σ^z()+[σ^z(0)σ^z()]e(K12+K21)t.delimited-⟨⟩subscript^𝜎𝑧𝑡delimited-⟨⟩subscript^𝜎𝑧delimited-[]delimited-⟨⟩subscript^𝜎𝑧0delimited-⟨⟩subscript^𝜎𝑧superscript𝑒subscript𝐾12subscript𝐾21𝑡\langle\hat{\sigma}_{z}(t)\rangle=\langle\hat{\sigma}_{z}(\infty)\rangle+\big{% [}\langle\hat{\sigma}_{z}(0)\rangle-\langle\hat{\sigma}_{z}(\infty)\rangle\big% {]}e^{-(K_{12}+K_{21})t}.⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_t ) ⟩ = ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( ∞ ) ⟩ + [ ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( 0 ) ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( ∞ ) ⟩ ] italic_e start_POSTSUPERSCRIPT - ( italic_K start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT + italic_K start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT . (22)

As the rate constants KABsubscript𝐾𝐴𝐵K_{AB}italic_K start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT calculated by Eq. (10) are always real, the state populations in MQME can only undergo a simple exponential decay without any oscillations.

The results for const-ΛΛ\Lambdaroman_Λ series in Fig. 2 exhibit a similar trend as observed in Fig. 1. Namely, MQME accurately describes the population transfer when ΔE1Δ𝐸1\Delta E\geq 1roman_Δ italic_E ≥ 1 [Fig. 2(d)–(k)] with slight overestimations in the rate, although the performance becomes poor for the cases with ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Fig. 2(a)–(c)]. The discrepancy between MQME and HEOM becomes more provoked as the temperature decreases, due to the reduced strength of the thermal fluctuation induced by the subsystem-bath interaction.

III.3 Dissipation Dynamics

Having confirmed the reliability of MQME for the electronic dynamics, we now examine the dissipation. Figure 3 shows how the dissipated energy is distributed along the frequency axis for const-T𝑇Titalic_T series, calculated by using both MQME-D and HEOM-D. As we have seen in Sec. II.2, MQME-D predicts the dissipation into both molecules to be exactly identical when J1(ω)=J2(ω)subscript𝐽1𝜔subscript𝐽2𝜔J_{1}(\omega)=J_{2}(\omega)italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ω ) = italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ω ).[4] With HEOM-D we can now put this statement under a close inspection, which was not attempted in Ref. 4 due to the limited accuracy of the benchmark simulation method employed therein. We therefore avoid the redundancy by plotting the dissipation for only one molecule for MQME-D simulations, while separately plotting for either molecules for HEOM-D simulations.

Refer to caption
Figure 3: Steady-state cumulative dissipation density (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) calculated for const-T𝑇Titalic_T series by using both MQME-D and HEOM-D. For MQME-D, the results for both molecules are identical and therefore plotted as a single profile. For HEOM-D, the results for the two molecules are plotted separately.

The dissipation can be conveniently visualized in the frequency domain by using the cumulative dissipation density defined in Eq. (17). Figure 3 shows (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) for const-T𝑇Titalic_T series obtained by both MQME-D and HEOM-D. When ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Fig. 3(a)–(d)], MQME-D predicts vanishing dissipation for all ω𝜔\omegaitalic_ω, although HEOM-D results clearly demonstrate net energy transfer from the subsystem to bath for all four values of ΛΛ\Lambdaroman_Λ. Such an incorrect behavior of MQME-D arises because EA=EBsubscript𝐸𝐴subscript𝐸𝐵E_{A}=E_{B}italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT makes 𝒥BAA(ω)=𝒥ABA(ω)superscriptsubscript𝒥𝐵𝐴𝐴𝜔superscriptsubscript𝒥𝐴𝐵𝐴𝜔\mathcal{J}_{BA}^{A}(\omega)=\mathcal{J}_{AB}^{A}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) = caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) [Eqs. (18) and (19)], while the detailed balance condition also imposes 𝒥BAA(ω)=𝒥ABA(ω)superscriptsubscript𝒥𝐵𝐴𝐴𝜔superscriptsubscript𝒥𝐴𝐵𝐴𝜔\mathcal{J}_{BA}^{A}(\omega)=-\mathcal{J}_{AB}^{A}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) = - caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) (Sec. II C of Paper I[1]). These two conditions can be simultaneously met only when both 𝒥ABA(ω)superscriptsubscript𝒥𝐴𝐵𝐴𝜔\mathcal{J}_{AB}^{A}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) and 𝒥BAA(ω)superscriptsubscript𝒥𝐵𝐴𝐴𝜔\mathcal{J}_{BA}^{A}(\omega)caligraphic_J start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_ω ) vanishes for all ω𝜔\omegaitalic_ω, leading to the absence of dissipation as shown in Fig. 3(a)–(d). In reality, however, energy is dissipated from the subsystem to the bath due to the population relaxation between the eigenstates of H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT. While such an aspect is naturally incorporated in HEOM-D, MQME-D cannot handle the effects arising from delocalized eigenstates due to its focus on the projected system density under second-order perturbation (Sec. II A of Paper I[1]).

When ΔEΔ𝐸\Delta Eroman_Δ italic_E becomes larger, we observe much better qualitative agreement between the predictions of MQME-D and HEOM-D. Of particular interest are the cases with ΔE=2Δ𝐸2\Delta E=2roman_Δ italic_E = 2, for which MQME-D exhibits semi-quantitative accuracy [Fig. 3(i)–(l)] with slight overestimation (underestimation) of the dissipation when ω𝜔\omegaitalic_ω is small (large). With a relatively small reorganization energy of Λ=0.05Λ0.05\Lambda=0.05roman_Λ = 0.05 [Fig. 3(i)], we observe that a substantial portion of the dissipation occurs through the region around ω=ΔEPlanck-constant-over-2-pi𝜔Δ𝐸\hbar\omega=\Delta Eroman_ℏ italic_ω = roman_Δ italic_E, which can be related to the vibronic resonance. Increasing ΛΛ\Lambdaroman_Λ makes the contribution of this channel gradually disappear while the dissipation becomes more concentrated near ω=0𝜔0\omega=0italic_ω = 0. However, the MQME-D calculations predict that (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) monotonously increases as ω𝜔\omegaitalic_ω approaches zero, which contradicts the steep drops observed in the HEOM-D counterparts [Fig. 3(k) and (l)]. Our previous study[4] clarified that the discrepancy at the low frequency is caused by the Markov approximation behind MQME-D, which neglects the quasi-static nature of the bath modes that delays their participation in the dynamics.

Refer to caption
Figure 4: Same as in Fig. 3, but for the const-ΛΛ\Lambdaroman_Λ series.

Finally, we inspect the effect of temperature on dissipation by looking into Fig. 4, which summarizes (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) for const-ΛΛ\Lambdaroman_Λ series calculated by both MQME-D and HEOM-D. As in Fig. 3, MQME-D erroneously predicts zero dissipation for ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Fig. 4(a)–(c)]. Nevertheless, the accuracy of MQME-D increases with ΔEΔ𝐸\Delta Eroman_Δ italic_E and reaches semi-quantitative level for ΔE=2Δ𝐸2\Delta E=2roman_Δ italic_E = 2 [Fig. 4(g)–(i)]. Apparently, lowering T𝑇Titalic_T from 1.0 to 0.25 enhances the influence of the vibronic resonance in dissipation, which is due to the reduction of the thermal fluctuation induced by the subsystem-bath interaction. Similar to what we have seen from Fig. 3, the dissipation calculated by MQME-D tends to be concentrated toward slightly lower frequency compared to the HEOM-D results. Again, this is because MQME-D is based on Markov approximation and therefore overestimates the contribution of the low-frequency bath modes on the dissipation.

III.4 Asymmetry of the Dissipation between Molecules

As we have stated earlier, MQME-D predicts the dissipation into the two molecules in a dimer to be exactly identical.[4] According to Figs. 3 and 4, this prediction is satisfied for large ω𝜔\omegaitalic_ω and small ΛΛ\Lambdaroman_Λ. However, HEOM-D calculations also show that some asymmetry do exist between 1(ω,)subscript1𝜔\mathcal{E}_{1}(\omega,\infty)caligraphic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ω , ∞ ) and 2(ω,)subscript2𝜔\mathcal{E}_{2}(\omega,\infty)caligraphic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ω , ∞ ), especially when ΔEΔ𝐸\Delta Eroman_Δ italic_E is small [Figs. 3(a)–(d)] or ΛΛ\Lambdaroman_Λ is relatively large [Figs. 3(k) and (l)]. The most dramatic cases are the conditions with ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Figs. 3(b)-3(d) and Figs. 4(a)-4(c)], where the directions of dissipation in the two molecules become completely opposite for low-frequency bath modes. This shows that the low-frequency modes of molecule 1 only lose energy from the reorganization, while those of molecule 2 actively absorb energy from the subsystem. It would be meaningful to conduct a further inspection about the origin of this asymmetry to clarify how it is connected to the approximations underlying MQME-D.

As the asymmetry becomes more pronounced when ω𝜔\omegaitalic_ω is small, we can speculate that the phenomenon is related to non-Markovianity. A possible source of the non-Markovianity in our dimer system is the non-equilibrium motion of the bath triggered by Franck-Condon transition to the excited state PES. To elaborate further on this phenomenon, we separately examine the behavior of the bath density in MQME and HEOM dynamics. In Sec. II B of Paper I,[1] we explained that MQME focuses only on the projected component of the density matrix for the system ρ^(t)^𝜌𝑡\hat{\rho}(t)over^ start_ARG italic_ρ end_ARG ( italic_t )

𝒫^ρ^(t)=A(PA(t)|AA|exp(βh^A)Trb[exp(βh^A)]),^𝒫^𝜌𝑡subscript𝐴tensor-productsubscript𝑃𝐴𝑡ket𝐴bra𝐴𝛽subscript^𝐴subscriptTrbdelimited-[]𝛽subscript^𝐴\hat{\mathcal{P}}\hat{\rho}(t)=\sum_{A}\bigg{(}P_{A}(t)\ket{A}\bra{A}\otimes% \frac{\exp(-\beta\hat{h}_{A})}{\text{Tr}_{\text{b}}[\exp(-\beta\hat{h}_{A})]}% \bigg{)},over^ start_ARG caligraphic_P end_ARG over^ start_ARG italic_ρ end_ARG ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) | start_ARG italic_A end_ARG ⟩ ⟨ start_ARG italic_A end_ARG | ⊗ divide start_ARG roman_exp ( - italic_β over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) end_ARG start_ARG Tr start_POSTSUBSCRIPT b end_POSTSUBSCRIPT [ roman_exp ( - italic_β over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) ] end_ARG ) , (23)

where h^A=A|H^|Asubscript^𝐴bra𝐴^𝐻ket𝐴\hat{h}_{A}=\bra{A}\hat{H}\ket{A}over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = ⟨ start_ARG italic_A end_ARG | over^ start_ARG italic_H end_ARG | start_ARG italic_A end_ARG ⟩ and TrbsubscriptTrb\text{Tr}_{\text{b}}Tr start_POSTSUBSCRIPT b end_POSTSUBSCRIPT denotes the trace over the subspace spanned by the bath DOFs. If we combine this condition with the explicit expressions for H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG and its components [Eqs. (1)–(4)], we can observe that the molecular vibrational modes in MQME dynamics are always in the thermal equilibrium associated with the PES of the relevant electronic state. As the shape of the PES for the ground and excited states are identical in our Hamiltonian model [Eqs. (14) and (15)], the excited state bath density is identical to the ground state density except its center is shifted to the minimum of the excited state PES. Under such a condition, the relative dissipation rates for the vibrational modes are solely determined by the distance between the minima of ground and excited state PESs. Because these distances are identical for both molecules when their spectral densities [Eq. (5)] are the same, the dissipation by the two molecules must be symmetric as already proven in Sec. II.2.

In contrast to MQME, HEOM assumes that the initial system density to be in the direct product form ρ^(0)=σ^(0)R^g^𝜌0tensor-product^𝜎0subscript^𝑅g\hat{\rho}(0)=\hat{\sigma}(0)\otimes\hat{R}_{\text{g}}over^ start_ARG italic_ρ end_ARG ( 0 ) = over^ start_ARG italic_σ end_ARG ( 0 ) ⊗ over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT g end_POSTSUBSCRIPT, where σ^(0)^𝜎0\hat{\sigma}(0)over^ start_ARG italic_σ end_ARG ( 0 ) is the initial subsystem density and

R^g=exp(βH^bath)Trb[exp(βH^bath)]subscript^𝑅g𝛽subscript^𝐻bathsubscriptTrbdelimited-[]𝛽subscript^𝐻bath\hat{R}_{\text{g}}=\frac{\exp(-\beta\hat{H}_{\text{bath}})}{\text{Tr}_{\text{b% }}[\exp(-\beta\hat{H}_{\text{bath}})]}over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT g end_POSTSUBSCRIPT = divide start_ARG roman_exp ( - italic_β over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT ) end_ARG start_ARG Tr start_POSTSUBSCRIPT b end_POSTSUBSCRIPT [ roman_exp ( - italic_β over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bath end_POSTSUBSCRIPT ) ] end_ARG (24)

is the equilibrium bath density on the PES of the electronic ground state.[22] Because we have set the initial electronic populations as P1(0)=1subscript𝑃101P_{1}(0)=1italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) = 1 and P2(0)=0subscript𝑃200P_{2}(0)=0italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) = 0, the bath modes coupled to molecule 1 undergo Franck-Condon excitation and start oscillating in the excited state PES while those coupled to molecule 2 do not. We can easily expect that such a difference in the dynamics of the bath density would cause asymmetry in the dissipation by affecting microscopic energy flows between vibronic quantum states.

To corroborate our explanation further, we directly calculate and visualize how the asymmetry in the dissipation is affected by the initial electronic populations. For this purpose, we have specifically chosen the case of ω=0.2𝜔0.2\omega=0.2italic_ω = 0.2 and Λ=1.0Λ1.0\Lambda=1.0roman_Λ = 1.0 which exhibits a noticeable difference between 1(ω,)subscript1𝜔\mathcal{E}_{1}(\omega,\infty)caligraphic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ω , ∞ ) and 2(ω,)subscript2𝜔\mathcal{E}_{2}(\omega,\infty)caligraphic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ω , ∞ ) [Fig. 3(k)]. We calculated the time-dependent dissipation (0.2,t)0.2𝑡\mathcal{E}(0.2,t)caligraphic_E ( 0.2 , italic_t ) for two different initial conditions: (i) P1(0)=1subscript𝑃101P_{1}(0)=1italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) = 1, P2(0)=0subscript𝑃200P_{2}(0)=0italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) = 0 and (ii) P1(0)=P2(0)=0.5subscript𝑃10subscript𝑃200.5P_{1}(0)=P_{2}(0)=0.5italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) = italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) = 0.5. The results plotted in Fig. 4 illustrate that the asymmetry under the original initial condition [Fig. 5(a)] disappears as expected when we induce the same amount of Franck-Condon excitation for both molecules by setting their initial electronic populations as equal [Fig. 5(b)]. Hence, it supports our claim that the asymmetry in the dissipation is indeed linked to the difference between the non-equilibrium motion of the bath. Intriguingly, although 1(0.2,t)subscript10.2𝑡\mathcal{E}_{1}(0.2,t)caligraphic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0.2 , italic_t ) in the early stage of the dynamics clearly exhibits oscillations arising from nuclear motions, such a feature is not visible in 2(0.2,t)subscript20.2𝑡\mathcal{E}_{2}(0.2,t)caligraphic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0.2 , italic_t ) even with some amount of excitation initially residing in molecule 2 [Fig. 5(b)].

Refer to caption
Figure 5: Densities of accumulated dissipation at ω=0.2𝜔0.2\omega=0.2italic_ω = 0.2, separately obtained for both molecules by using HEOM-D under simulation condition (iii) (Table 1) for ΔEΔ𝐸\Delta Eroman_Δ italic_E = 2. Two different initial electronic populations were employed to demonstrate that the asymmetry between the two molecules are related to non-equilibrium motion of the bath modes.

IV Non-Markovian Dissipation via Time Scale Separation (TSS)

In this section, we conduct extensive tests on MQME-D+TSS. Section IV.1 introduces the principles of TSS. Section IV.2 uses MQME-D+TSS to compute the dissipation in different types of Hamiltonian models with a wide range of simulation parameters. The results are then compared with HEOM-D to appraise the accuracy of our theory.

IV.1 Introduction to Time Scale Separation

In Sec. III.3, we have seen that one source of inaccuracy in MQME-D is the Markov approximation. This suggests that the reliability of MQME-D may be increased if we can somehow include non-Markovianity in the simulation. This task can be accomplished by employing TSS,[23, 13] which divides the BSD J(ω)𝐽𝜔J(\omega)italic_J ( italic_ω ) into “slow” Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) and “fast” Jfast(ω)subscript𝐽fast𝜔J_{\text{fast}}(\omega)italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ) components, respectively. The desired non-Markovianity is introduced in the simulation by prohibiting the bath modes in Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) from directly participating in the dynamics. This is practically achieved by defining the BSD components as

Jslow(ω)=S(ω,ω)J(ω),subscript𝐽slow𝜔𝑆𝜔superscript𝜔𝐽𝜔J_{\text{slow}}(\omega)=S(\omega,\omega^{*})J(\omega),italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) = italic_S ( italic_ω , italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_J ( italic_ω ) , (25a)
Jfast(ω)=[1S(ω,ω)]J(ω).subscript𝐽fast𝜔delimited-[]1𝑆𝜔superscript𝜔𝐽𝜔J_{\text{fast}}(\omega)=[1-S(\omega,\omega^{*})]J(\omega).italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ) = [ 1 - italic_S ( italic_ω , italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ] italic_J ( italic_ω ) . (25b)

Here, S(ω,ω)𝑆𝜔superscript𝜔S(\omega,\omega^{*})italic_S ( italic_ω , italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) is the splitting function which monotonically decays as ω𝜔\omegaitalic_ω increases, with the speed of decay controlled by the cutoff frequency ωsuperscript𝜔\omega^{*}italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. In our simulations, by following the similar approach as in Ref. 13, we use the splitting function of the form

S(ω,ω)={η[1(ω/ω)2]2,ω<ω,0,ωω,𝑆𝜔superscript𝜔cases𝜂superscriptdelimited-[]1superscript𝜔superscript𝜔22𝜔superscript𝜔0𝜔superscript𝜔S(\omega,\omega^{*})=\left\{\begin{array}[]{cc}\eta[1-(\omega/\omega^{*})^{2}]% ^{2},&\omega<\omega^{*},\\ 0,&\omega\geq\omega^{*},\end{array}\right.italic_S ( italic_ω , italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = { start_ARRAY start_ROW start_CELL italic_η [ 1 - ( italic_ω / italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL start_CELL italic_ω < italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL 0 , end_CELL start_CELL italic_ω ≥ italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , end_CELL end_ROW end_ARRAY (26)

where we have added an extra scaling factor η<1𝜂1\eta<1italic_η < 1 to the original expression to ensure the numerical convergence of the improper integrals [Eqs. (10) and (19)].

We now treat Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) as a source of static disorder, and modulate the state energies of the subsystem {EA}subscript𝐸𝐴\{E_{A}\}{ italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT } by adding Gaussian random noise whose standard deviation is

σslow=10Jslow(ω)coth(βω2)𝑑ω,subscript𝜎slow1Planck-constant-over-2-pisuperscriptsubscript0subscript𝐽slow𝜔hyperbolic-cotangent𝛽Planck-constant-over-2-pi𝜔2differential-d𝜔\sigma_{\text{slow}}=\frac{1}{\hbar}\int_{0}^{\infty}J_{\text{slow}}(\omega)% \coth\bigg{(}\frac{\beta\hbar\omega}{2}\bigg{)}\>d\omega,italic_σ start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) roman_coth ( divide start_ARG italic_β roman_ℏ italic_ω end_ARG start_ARG 2 end_ARG ) italic_d italic_ω , (27)

while Jfast(ω)subscript𝐽fast𝜔J_{\text{fast}}(\omega)italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ) governs the dissipation in individual realizations of the disorder. The final result is calculated by averaging over sufficiently large number of realizations. Each noise trajectory follows MQME whose BSD is Jfast(ω)subscript𝐽fast𝜔J_{\text{fast}}(\omega)italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ), and therefore exhibits the characteristics of the bare MQME such as exponential time-dependence and detailed balance PA(ω)/PB(ω)=exp[β(EBEA)]subscript𝑃𝐴𝜔subscript𝑃𝐵𝜔𝛽subscript𝐸𝐵subscript𝐸𝐴P_{A}(\omega)/P_{B}(\omega)=\exp[\beta(E_{B}-E_{A})]italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_ω ) / italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( italic_ω ) = roman_exp [ italic_β ( italic_E start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ) ]. However, because the static disorder induces variation in the state energies, such properties are not satisfied after averaging over the entire set of trajectories.

IV.2 Dissipation Dynamics

We now benchmark the accuracy of MQME-D+TSS against different types of model Hamiltonians for open quantum systems. In Sec. IV.2.1, we first explore the performance of MQME-D + TSS for the molecular dimer model which was already employed to benchmark the original MQME-D in Sec. III. Section IV.2.2 applies MQME-D to the spin-boson Hamiltonian with Brownian oscillator BSD.

IV.2.1 Molecular Dimer

Refer to caption
Figure 6: Split of the Drude-Lorentz BSD [Eq. (7)] into slow and fast components by employing Eqs. (25) and (26) with η=0.99𝜂0.99\eta=0.99italic_η = 0.99 and ω=0.2superscript𝜔0.2\omega^{*}=0.2italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2.

We examine the same molecular dimer model used in Sec. III.3, defined by the local bath Hamiltonian [Eqs. (2), (14), and (15)]. We apply TSS by constructing the splitting function [Eq. (26)] with η=0.99𝜂0.99\eta=0.99italic_η = 0.99 and ω=0.2superscript𝜔0.2\omega^{*}=0.2italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2, and averaging over 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT noise trajectories for conditions (ii)–(vi) in Table 1. For the condition (i) with Λ=0.05Λ0.05\Lambda=0.05roman_Λ = 0.05, the number of trajectories was increased to 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT to ensure numerical convergence. Figure 6 illustrates how TSS splits the Drude-Lorentz BSD used in our simulations. All other simulation procedures remain the same as those explained in Sec. III.1.

We first examine the effect of TSS in population dynamics. Figure 1 plots σ^z(t)delimited-⟨⟩subscript^𝜎𝑧𝑡\langle\hat{\sigma}_{z}(t)\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_t ) ⟩ obtained from both bare MQME and MQME+TSS obtained for const-T𝑇Titalic_T series. It is observed that TSS modifies the rate of population relaxation, although the direction and magnitude of the influence show complicated dependence on both ΛΛ\Lambdaroman_Λ and ΔEΔ𝐸\Delta Eroman_Δ italic_E. Nevertheless, there are cases for which TSS substantially increases the accuracy of MQME. In particular, nearly perfect matches between MQME-TSS and HEOM are observed for Figs. 1(i) and (j). In Figs. 1(e) and (f), TSS also leads to better agreements with HEOM-D results by prolonging the relaxation. Nevertheless, the accuracy of MQME-D+TSS may become further improved by using different values of ωsuperscript𝜔\omega^{*}italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. On the other hand, Figs. 1 (c) and (d) also demonstrate that TSS does not always positively effect the accuracy. Similar trends are observed for const-ΛΛ\Lambdaroman_Λ series (Fig. 2).

Refer to caption
Figure 7: Cumulative dissipation density at the steady state (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ), calculated for const-T𝑇Titalic_T series by combining MQME-D+TSS with the cutoff frequency ω=0.2superscript𝜔0.2\omega^{*}=0.2italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2. The averages of the HEOM-D results (dashed purple) for both BSDs in Fig. 3 are plotted together to benchmark the accuracy of the results.
Refer to caption
Figure 8: Similar to Fig. 7, but for const-ΛΛ\Lambdaroman_Λ series. In this case, the benchmark is the averages of the HEOM-D results in Fig. 4.

We now examine the dissipation predicted by MQME-D+TSS to see how the added non-Markovianity affects the accuracy of the method. We note that TSS does not save the lack of asymmetry in MQME, as we are applying the same S(ω,ω)𝑆𝜔superscript𝜔S(\omega,\omega^{*})italic_S ( italic_ω , italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) to the BSDs of the two molecules. We therefore compare the resolved dissipation against the HEOM-D results averaged over the two molecules.

Figures 7 and 8 display (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ) calculated by MQME-D+TSS, along with the averaged (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ) obtained from HEOM-D simulations. Compared to the results obtained without TSS (Figs. 2 and 3), it is apparent that TSS improves the accuracy of MQME-D for both const-T𝑇Titalic_T (Fig. 7) and const-ΛΛ\Lambdaroman_Λ (Fig. 8) series, especially for the cases with ΔE=2Δ𝐸2\Delta E=2roman_Δ italic_E = 2 [Figs. 7(i)–(l) and 8(g)–(i)]. In particular, the agreement near ω=0𝜔0\omega=0italic_ω = 0 became remarkably better as Jfast(ω)subscript𝐽fast𝜔J_{\text{fast}}(\omega)italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ) does not exhibit strong subsystem-bath interaction anymore in that region. The dissipation originally in this region is redirected toward higher frequencies, which alleviates the underestimation of (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ) near ω=2𝜔2\omega=2italic_ω = 2 by MQME-D. Such an effect is also pronounced for ΔE=1Δ𝐸1\Delta E=1roman_Δ italic_E = 1 [Figs. 7(e)–(h) and 8(d)–(f)], although some discrepancy still remains.

Finally, for the homodimer case of ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0 [Figs. 1(a)–(d) and 2(a)–(c)], we observe some amount of dissipation in contrast to the vanishing dissipation in Figs. 1 and 2. This is because the individual noise trajectories exhibit nonzero ΔEΔ𝐸\Delta Eroman_Δ italic_E due to the static disorder arising from Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ). However, because MQME loses its reliability when ΔE=0Δ𝐸0\Delta E=0roman_Δ italic_E = 0, MQME-D is also not accurate enough to make meaningful predictions of dissipation.

Overall, we can expect MQME-D+TSS will accurately decompose the dissipated energy when MQME qualitatively accounts for the dynamics of state populations (large ΔEΔ𝐸\Delta Eroman_Δ italic_E or ΛΛ\Lambdaroman_Λ), and TSS leads to an additional increase in the accuracy at the quantitative level.

IV.2.2 Spin-boson Model with Brownian Oscillator Bath

We now test how MQME-D+TSS performs for the Hamiltonian models involving Brownian oscillator BSD JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) [Eq. (8)]. We examine the dissipation induced by an underdamped bath mode whose frequency is tuned to achieve resonance with the subsystem. By modulating the reorganization energy ΛΛ\Lambdaroman_Λ and dam** parameter γ𝛾\gammaitalic_γ, we conduct a systematic investigation on how the strength of subsystem-bath interaction and memory time of the bath affect the accuracy of our method.

Refer to caption
Figure 9: Shapes of JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) used in the simulations of spin-boson model, for different values of the dam** strength γ𝛾\gammaitalic_γ. The inset in panel (c) shows how TSS splits the BSD into Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) (dashed red) and Jfast(ω)subscript𝐽fast𝜔J_{\text{fast}}(\omega)italic_J start_POSTSUBSCRIPT fast end_POSTSUBSCRIPT ( italic_ω ) (dashed blue) when γ=1.0𝛾1.0\gamma=1.0italic_γ = 1.0.
Table 2: Summary of the simulation conditions used for the spin-boson model with Brownian oscillator bath. The first parameter is for both MQME and HEOM calculations, while the rest are specifically for HEOM. Each of the 3 simulation conditions in the table was combined with three different values of the dam** constant ωcsubscript𝜔c\omega_{\text{c}}italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT to yield 9 different conditions in total.
Simulation condition (i) (ii) (iii)
Reorganization energy (ΛΛ\Lambdaroman_Λ) 0.05 0.25 1.0
Maximum time step (ΔtmaxΔsubscript𝑡max\Delta t_{\text{max}}roman_Δ italic_t start_POSTSUBSCRIPT max end_POSTSUBSCRIPT) 0.01 0.05 0.05
Number of hierarchy tiers (Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT) 5 7 12
Secondary Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT for ωc=0.05subscript𝜔c0.05\omega_{\text{c}}=0.05italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 0.05 10 15 25
H-R factor of the probe mode (sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT) 2×1062superscript1062\times 10^{-6}2 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 1×1051superscript1051\times 10^{-5}1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT

We simulate the dynamics of a spin-boson Hamiltonian defined by Eqs. (3), (20), and (21). The subsystem parameters are E=2𝐸2E=2italic_E = 2 and V=0.25𝑉0.25V=0.25italic_V = 0.25. For the BSD, we use three different values of ΛΛ\Lambdaroman_Λ as 0.05, 0.25 and 1.0 as listed in Table 2, and vary the dam** strength γ𝛾\gammaitalic_γ as 0.05, 0.25, and 1.0 (Fig. 9) for each value of ΛΛ\Lambdaroman_Λ to make a total of 9 simulation conditions. The characteristic frequency of the BSD was set as ω0=2.062subscript𝜔02.062\omega_{0}=2.062italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2.062 to match the difference between the eigenenergies of H^subsubscript^𝐻sub\hat{H}_{\text{sub}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT. The temperature of the bath was kept constant as T=1𝑇1T=1italic_T = 1 for all simulation conditions.

For MQME and MQME-D simulations, the rate constants for population transfer and dissipation were calculated according to Eqs. (10) and (13) by setting d±j=±djsubscript𝑑plus-or-minus𝑗plus-or-minussubscript𝑑𝑗d_{\pm j}=\pm d_{j}italic_d start_POSTSUBSCRIPT ± italic_j end_POSTSUBSCRIPT = ± italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT in Eq. (4). The BSD was discretized into 5000 harmonic oscillator modes based on the scheme described in Appendix B for all simulation conditions in Table 2 except the case of Λ=0.05Λ0.05\Lambda=0.05roman_Λ = 0.05 and γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05, which required 20000 oscillators to guarantee numerical convergence. The upper limit of frequency ωmaxsubscript𝜔max\omega_{\text{max}}italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT and the integration scheme were the same as what we used for the dimer model (Sec. III.1). The TSS was applied by using the same cutoff ω=0.2superscript𝜔0.2\omega^{*}=0.2italic_ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 as in the Drude-Lorentz BSDs in the dimer model (Sec. IV.2.1), while the scaling factor η𝜂\etaitalic_η was reduced from 0.99 to 0.6 due to the increased difficulty of achieving detailed balance condition for JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ). The number of individual noise trajectories was always kept as 1000.

For HEOM and HEOM-D simulations, we implemented the Brownian oscillator BSD based on the efficient framework reported in Ref. 14 and combined it with the perturbative low-temperature correction[20] with NMatsu=30subscript𝑁Matsu30N_{\text{Matsu}}=30italic_N start_POSTSUBSCRIPT Matsu end_POSTSUBSCRIPT = 30. The depth of hierarchy Nhiersubscript𝑁hierN_{\text{hier}}italic_N start_POSTSUBSCRIPT hier end_POSTSUBSCRIPT was adjusted depending on the reorganization energy as listed in Table 2. When γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05, deeper hierarchy was needed for the numerical convergence near ω=ω0𝜔subscript𝜔0\omega=\omega_{0}italic_ω = italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT due to the strong resonance arising from the subsystem-bath interaction. We scanned the frequency of the probe mode in the range of [0.2, 1.9) and (2.2, 3.0] with a constant spacing of 0.05, while a finer grid of 0.005 was used for the range of [1.9, 2.2] for capturing the detailed structure of (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ) near the resonance.

Refer to caption
Figure 10: Comparison between the evolution of population inversion σ^zdelimited-⟨⟩subscript^𝜎𝑧\langle\hat{\sigma}_{z}\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ for the spin-boson model, calculated by MQME (solid black), MQME+TSS (solid gray), and HEOM (dashed purple).
Refer to caption
Figure 11: Steady-state cumulative dissipation density (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) calculated for the spin-boson model by using MQME-D (solid black), MQME-D+TSS (solid gray), and HEOM-D (dashed purple). For the conditions with γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05, we increased the hierarchy depth for HEOM-D (dashed red) in the frequency domain [1.9, 2.2] to guarantee the convergence under strong resonance between the subsystem and bath modes.

In Fig. 10, we have presented the calculated σ^z(t)delimited-⟨⟩subscript^𝜎𝑧𝑡\langle\hat{\sigma}_{z}(t)\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_t ) ⟩ for all 9 simulation conditions listed in Table 2. The results for γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05 [Figs. 10(a)–(c)] show that it is challenging for MQME and MQME+TSS to describe highly non-Markovian character of the bath originating from the small dam**. However, the agreement becomes much better as γ𝛾\gammaitalic_γ increases, and show nearly quantitative match for [Figs. 10(d)–(i)]. Meanwhile, in contrast to the molecular dimer model coupled to Drude-Lorentz BSD (Sec. IV.2.1), there is almost no visible difference between MQME and MQME+TSS. This is because JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) takes relatively small value near ω=0𝜔0\omega=0italic_ω = 0, which makes Jslow(ω)subscript𝐽slow𝜔J_{\text{slow}}(\omega)italic_J start_POSTSUBSCRIPT slow end_POSTSUBSCRIPT ( italic_ω ) have only a minute contribution to the overall BSD [Fig. 9(c)] in our simulation conditions.

We now examine Fig. 11 and discuss how the dissipation by the Brownian oscillator BSD looks like. For small (γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05) and intermediate (γ=0.25𝛾0.25\gamma=0.25italic_γ = 0.25) dam** [Figs. 11(a)–(f)], HEOM-D results show that most of the dissipation occurs through the resonant channel around ω=2𝜔2\omega=2italic_ω = 2. What is interesting is that (ω,)𝜔\mathcal{E}(\omega,\infty)caligraphic_E ( italic_ω , ∞ ) for γ=0.05𝛾0.05\gamma=0.05italic_γ = 0.05 [Figs. 11(a)–(c)] does not form a single peak as in the Brownian BSD (Fig. 9) but instead a pair of peaks closely lying together. Such a structure arises from the interaction between the upper subsystem state |+ket\ket{+}| start_ARG + end_ARG ⟩ and the first excited state of the underdamped bath mode, as in the formation of a polaritonic state pair.[24, 25] Nevertheless, this behavior soon disappears as the effect of resonance is diluted due to the increased dam** [Figs. 11(d)–(i)]. For all panels in Fig. 11, both MQME-D and MQME-D+TSS qualitatively reproduce the results from HEOM-D calculations. The predictability becomes better with increasing ΛΛ\Lambdaroman_Λ and γ𝛾\gammaitalic_γ which enhances the adequacy of second-order perturbation and Markov approximation, respectively.

In contrast to the population dynamics for which TSS had virtually no effect, averaging over the TSS noise trajectories removes rapid oscillations in the dissipation, which appears near the resonance frequency (ω2𝜔2\omega\approx 2italic_ω ≈ 2) in the bare MQME-D [Figs. 11(b)–(e)]. These oscillations arise from incomplete numerical convergence, and it is possible to mitigate them to some extent by extending the upper limit of integration in Eq. (13). However, we found out that the convergence without TSS was extremely slow and the oscillations were still prevalent even after integrating up to 5×1055superscript1055\times 10^{5}5 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT (100 times the limit used in the main calculations). Therefore, TSS offers a convenient way to achieve converged results under strong subsystem-bath resonance, which can be also straightforwardly parallelized by distributing the propagation and averaging procedure over multiple processors.

As seen from Figs. 11(d), (e), (g), and (h), MQME-D+TSS tends to overestimate the dissipation near the region where subsystem-bath resonance occurs. This is likely because the dissipated energy cannot return to the subsystem in MQME due to the second-order approximation, although such re-absorption of energy by subsystem does occur when the underdamped bath mode exerts strong subsystem-bath resonance.[5]

V Computational Efficiency of MQME-D

In this paper, we obtained exact decompositions of dissipation in our Hamiltonian models by combining the numerically exact HEOM method[9, 14] with a technique for extracting the statistics of a particular bath mode.[8] However, such computations become exponentially more costly as the number of subsystem DOFs increases. Most of the other numerically exact simulation methods exhibit similar exponential scaling for propagating the subsystem RDM, although we are aware of a recently reported method that could potentially overcome this issue.[26]

In addition, the computational cost of HEOM-D also depends on the temperature of the bath and the characteristic frequency of the harmonic bath mode whose dissipation we want to calculate. As explained in Appendix A, HEOM-D extracts the dissipation by merging the subsystem with an extra bath mode which acts as a probe for monitoring the dynamics of the bath mode of interest. The probe mode must have the same frequency as the mode we want to monitor, and should be described by large enough number of quantum states to represent the thermal properties during the dynamics. Therefore, the number of required vibrational quantum states drastically increases as we reduce the energy spacing ωPlanck-constant-over-2-pi𝜔\hbar\omegaroman_ℏ italic_ω below kBTsubscript𝑘B𝑇k_{\text{B}}Titalic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T and move toward zero. If n𝑛nitalic_n quantum states are used to describe the probe mode, the cost of propagating the reduced density matrix and auxiliary density matrices of the extended subsystem would approximately depend on O(n3)𝑂superscript𝑛3O(n^{3})italic_O ( italic_n start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ), as it involves matrix-matrix multiplications arising from commutators and anti-commutators. Moreover, the perturbative low-temperature correction described in Ref. 20 requires diagonalizations of the super-operators represented in the Liouville space, whose cost would exhibit the dependence of O(n6)𝑂superscript𝑛6O(n^{6})italic_O ( italic_n start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ). Such a steep growth of the computational burden was the reason why we always needed to terminate HEOM-D calculations at a certain lower limit for ω𝜔\omegaitalic_ω.

On the other hand, the aforementioned issues pose much less difficulty for MQME-D. Namely, it is less problematic to apply MQME-D to large systems as the cost for evaluation of the dissipation rate constants and time propagation only grows as O(N2)𝑂superscript𝑁2O(N^{2})italic_O ( italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) where N𝑁Nitalic_N is the dimension of the subsystem, if we assume that all subsystem DOFs are coupled to an identical number of bath modes. In addition, MQME-D can be trivially parallelized by distributing the load of evaluating the rate constants and propagating TSS-based noise trajectories across multiple processors. This is in contrast to HEOM[27, 28] and tensor-train-based simulation methods[29, 30] whose equations-of-motion are usually densely coupled and therefore require substantial amounts of communication for propagation. Furthermore, the cost of evaluating the dissipation rate constants based on Eq. (13) does not depend on ω𝜔\omegaitalic_ω, which allows us to conveniently access the dissipation in low-frequency region without any additional burden.

Finally, when one employs numerically exact simulation methods to calculate the dissipation into individual bath modes, only one mode is usually monitored at a time to keep the dimension of the subsystem RDM within a viable extent. As a result, the formation and propagation of the extended subsystem needs to be repeated for every bath mode to obtain a complete decomposition of the dissipation. On the other hand, provided that the rate constants have been already computed, MQME-D captures all information regarding the dissipation in a single propagation. Based on these observations, we expect MQME-D with TSS to have a promising utility in studying the role of individual bath modes to the quantum dynamics in large molecular systems, whose details cannot be easily accessed by numerically exact simulation methods.

VI Conclusion

In this paper, we investigated the accuracy of MQME-D, a theoretical method that enables us to decompose the dissipation under MQME dynamics into contributions from individual bath modes. The theory was applied to multiple types of Hamiltonian models and the outcomes from the simulations were compared against numerically exact results provided by HEOM. We have demonstrated that the dissipation calculated by MQME-D offers a qualitatively correct view of the dissipated energy by individual bath modes. However, it can quantitatively differ from the exact results even in the limit where MQME exhibits accurate population dynamics. We have provided detailed arguments that support Markovian origin of the observed discrepancies, and demonstrated that the accuracy of the calculation is indeed significantly increased by inclusion of non-Markovianity via TSS. In the end, despite the inherent limitation arising from second-order perturbation approximation, MQME-D combined with TSS offers an efficient way to obtain semi-quantitative decompositions of the dissipation in wide range of subsystem-bath couplings and temperatures. Even for the Brownian oscillator bath for which TSS does not significantly affect the BSD, TSS offered a useful way to improve the numerical convergence of MQME-D. However, TSS could not reproduce the asymmetry between the dissipation by two molecules, which becomes more pronounced toward the low-frequency region. We expect the asymmetry may be realized in our method by using separate scaling functions for different BSDs or extending our method to include non-equilibrium motion of the bath modes.[31]

MQME-D shows quadratic scaling of the computational cost with the size of the system, and its parallelization across multiple processors is also straightforward. Moreover, the cost of MQME-D does not depend on the characteristic frequency of the bath mode, in contrast of HEOM-D which shows rapid increase of the burden as ω𝜔\omegaitalic_ω decreases. We therefore expect the usefulness of MQME-D to grow with larger systems for which the computational costs of numerically exact methods become expensive due to their exponential scaling and challenges in parallelization.

We anticipate that applying the framework outlined in Paper I[1] to a range of quantum master equations would lead us to corresponding dissipation theories in the near future. Upon rigorous validation as demonstrated in this paper, these theories can be integrated with realistic Hamiltonian models extracted by using state-of-the-art experimental and computational techniques. [32, 33, 34, 35, 36] We envisage that such efforts would lead us to deeper understandings on the quantum dynamics in a wide range of systems including photosynthetic complexes,[37, 38, 39] artificial excitonic systems,[40, 41, 42] plasmonic systems,[43, 44] and molecular and solid-state qubits.[45, 46, 47]

Supplementary Material

See the Supplementary Material for comprehensive discussion and analysis on the artificial drift in the cumulative dissipation density (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ), which demonstrate that (ω,t)𝜔𝑡\mathcal{E}(\omega,t)caligraphic_E ( italic_ω , italic_t ) from HEOM can be reliably used as a quantitative benchmark after calibrating the drift.

Acknowledgements.
CWK was financially supported by the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT (MSIT) of Korea (Grant Number: 2022R1F1A1074027, 2023M3K5A1094813, and RS-2023-00218219). IF is supported by the National Science Foundation under Grant No. CHE-2102386 and PHY-2310657.

Data Availability

The data and computer programs that support the findings of this study are available from the corresponding author upon reasonable request.

Appendix A A Brief Introduction to HEOM-D Method

In this section, we present a minimal explanation about the motivation and formulation of HEOM-D,[8] a technique we used for calculating the dissipation into a specific bath component in HEOM simulations. In HEOM, the RDM of the subsystem σ(t)𝜎𝑡\sigma(t)italic_σ ( italic_t ) is propagated by coupled equations of motion which connect σ(t)𝜎𝑡\sigma(t)italic_σ ( italic_t ) to the hierarchy of ADMs. It was reported that ADMs encode the consequence of subsystem-bath interaction during the dynamics, and therefore one can extract the statistics related to the bath from the ADMs.[48, 49, 50] However, such methods could only handle the bath modes in the entire BSD collectively, and does not allow isolation of the information regarding a single bath mode. Moreover, the widely used approach which re-classifies the bath mode of interest as the subsystem[51, 16, 17] is not allowed for HEOM. This is because the subtraction of a bath mode from a BSD converts its analytical quantum time correlation function to a form which cannot be handled by HEOM without drastically increasing the complexity of the calculation.[52]

The HEOM-D method[8] overcomes this challenge by introducing an extra bath mode (“probe mode”) that weakly couples to the subsystem through the same channel as the bath mode under examination (“target mode”). Under such a setting, the dynamical information regarding the target mode can be elucidated from that of the probe mode. To formulate the method, we first divide the full Hamiltonian [Eq. (1)] into contributions of the target mode and the rest,

H^=H^bt+H^rest,^𝐻subscript^𝐻btsubscript^𝐻rest\hat{H}=\hat{H}_{\text{bt}}+\hat{H}_{\text{rest}},over^ start_ARG italic_H end_ARG = over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT rest end_POSTSUBSCRIPT , (28)

respectively, where H^btsubscript^𝐻bt\hat{H}_{\text{bt}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT can be factorized as s^b^tensor-product^𝑠^𝑏\hat{s}\otimes\hat{b}over^ start_ARG italic_s end_ARG ⊗ over^ start_ARG italic_b end_ARG with s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG and b^^𝑏\hat{b}over^ start_ARG italic_b end_ARG representing the subsystem and bath part of H^btsubscript^𝐻bt\hat{H}_{\text{bt}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT, respectively. For example, if we want to examine the k𝑘kitalic_k-th bath mode in the spin-boson Hamiltonian [Eqs. (20)–(21)], s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG and b^^𝑏\hat{b}over^ start_ARG italic_b end_ARG needs to be set as

s^=|++|||,b^=p^k22+ωk22(x^kdk)2.formulae-sequence^𝑠ketbraketbra^𝑏superscriptsubscript^𝑝𝑘22superscriptsubscript𝜔𝑘22superscriptsubscript^𝑥𝑘subscript𝑑𝑘2\hat{s}=\ket{+}\bra{+}-\ket{-}\bra{-},\quad\hat{b}=\frac{\hat{p}_{k}^{2}}{2}+% \frac{\omega_{k}^{2}}{2}(\hat{x}_{k}-d_{k})^{2}.over^ start_ARG italic_s end_ARG = | start_ARG + end_ARG ⟩ ⟨ start_ARG + end_ARG | - | start_ARG - end_ARG ⟩ ⟨ start_ARG - end_ARG | , over^ start_ARG italic_b end_ARG = divide start_ARG over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + divide start_ARG italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_d start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (29)

We now modify the system Hamiltonian according to

H^=H^+H^bp,superscript^𝐻^𝐻subscript^𝐻bp\hat{H}^{\prime}=\hat{H}+\hat{H}_{\text{bp}},over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = over^ start_ARG italic_H end_ARG + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT , (30)

where we added a new Hamiltonian component H^bp=αH^btsubscript^𝐻bp𝛼subscript^𝐻bt\hat{H}_{\text{bp}}=\alpha\hat{H}_{\text{bt}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT = italic_α over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT for the probe mode. The scaling constant α𝛼\alphaitalic_α is the ratio between the coupling strengths of the target and probe modes, whose value should be small enough to ensure that the dynamics under H^superscript^𝐻\hat{H}^{\prime}over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT remains almost identical to that under H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG. As H^bpsubscript^𝐻bp\hat{H}_{\text{bp}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT was not a part of the original BSD and therefore does not alter the structure of HEOM, it can now be freely included in the subsystem and monitored over time. At the start of the dynamics, the density matrix ρ^bp(0)subscript^𝜌bp0\hat{\rho}_{\text{bp}}(0)over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( 0 ) for the probe mode is constructed as the same thermal equilibrium associated with the target mode. Then, the initial density σ^(0)superscript^𝜎0\hat{\sigma}^{\prime}(0)over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) for the extended subsystem H^sub+H^bpsubscript^𝐻subsubscript^𝐻bp\hat{H}_{\text{sub}}+\hat{H}_{\text{bp}}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT sub end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT is set to be σ^(0)ρ^bptensor-product^𝜎0subscript^𝜌bp\hat{\sigma}(0)\otimes\hat{\rho}_{\text{bp}}over^ start_ARG italic_σ end_ARG ( 0 ) ⊗ over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT and propagated by the same structure of HEOM used for the original subsystem RDM σ^(t)^𝜎𝑡\hat{\sigma}(t)over^ start_ARG italic_σ end_ARG ( italic_t ). Practically, ρ^bpsubscript^𝜌bp\hat{\rho}_{\text{bp}}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT is implemented by using a finite number n𝑛nitalic_n of bath quantum states which faithfully represent the thermal equilibrium. The value of n𝑛nitalic_n should be chosen carefully so as not to excessively increase the computational burden, as the time spent for applying perturbative low-temperature correction and RDM propagation increases with the order of O(n6)𝑂superscript𝑛6O(n^{6})italic_O ( italic_n start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ) and O(n3)𝑂superscript𝑛3O(n^{3})italic_O ( italic_n start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ), respectively (Sec. V). After propagating σ^(t)superscript^𝜎𝑡\hat{\sigma}^{\prime}(t)over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) for a certain amount of time, the time-dependent dissipation ΔEbp(t)Δsubscript𝐸bp𝑡\Delta E_{\text{bp}}(t)roman_Δ italic_E start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( italic_t ) induced by the probe mode can be calculated by

ΔEbp(t)=Tr[H^bp{σ^(t)σ^(0)}].Δsubscript𝐸bp𝑡Trdelimited-[]subscript^𝐻bpsuperscript^𝜎𝑡superscript^𝜎0\Delta E_{\text{bp}}(t)=\text{Tr}\big{[}\hat{H}_{\text{bp}}\big{\{}\hat{\sigma% }^{\prime}(t)-\hat{\sigma}^{\prime}(0)\big{\}}\big{]}.roman_Δ italic_E start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( italic_t ) = Tr [ over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT { over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) - over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) } ] . (31)

In Ref. 8, we proved that ΔEbp(t)Δsubscript𝐸bp𝑡\Delta E_{\text{bp}}(t)roman_Δ italic_E start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( italic_t ) is related to the dissipation ΔEbt(t)Δsubscript𝐸bt𝑡\Delta E_{\text{bt}}(t)roman_Δ italic_E start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT ( italic_t ) induced by the target mode via

ΔEbt(t)=limα0ΔEbp(t)α2.Δsubscript𝐸bt𝑡subscript𝛼0Δsubscript𝐸bp𝑡superscript𝛼2\Delta E_{\text{bt}}(t)=\lim_{\alpha\rightarrow 0}\frac{\Delta E_{\text{bp}}(t% )}{\alpha^{2}}.roman_Δ italic_E start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT ( italic_t ) = roman_lim start_POSTSUBSCRIPT italic_α → 0 end_POSTSUBSCRIPT divide start_ARG roman_Δ italic_E start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (32)

For harmonic oscillator bath, Eq. (32) can be converted to the dissipation density by combining it with the definition of the BSD [Eq. (5)],

(ωbt,t)=limsbp0J(ωbp)ωbp2sbpΔEbp(t),subscript𝜔bt𝑡subscriptsubscript𝑠bp0𝐽subscript𝜔bpPlanck-constant-over-2-pisuperscriptsubscript𝜔bp2subscript𝑠bpΔsubscript𝐸bp𝑡\mathcal{E}(\omega_{\text{bt}},t)=\lim_{s_{\text{bp}}\rightarrow 0}\frac{J(% \omega_{\text{bp}})}{\hbar\omega_{\text{bp}}^{2}s_{\text{bp}}}\Delta E_{\text{% bp}}(t),caligraphic_E ( italic_ω start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT , italic_t ) = roman_lim start_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT divide start_ARG italic_J ( italic_ω start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ) end_ARG start_ARG roman_ℏ italic_ω start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT end_ARG roman_Δ italic_E start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT ( italic_t ) , (33)

where ωbt=ωbpsubscript𝜔btsubscript𝜔bp\omega_{\text{bt}}=\omega_{\text{bp}}italic_ω start_POSTSUBSCRIPT bt end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT is the frequency of the target and probe modes, and sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT is the H-R factor of the probe mode

sbp=ωbpdbp22.subscript𝑠bpsubscript𝜔bpsuperscriptsubscript𝑑bp22Planck-constant-over-2-pis_{\text{bp}}=\frac{\omega_{\text{bp}}d_{\text{bp}}^{2}}{2\hbar}.italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT = divide start_ARG italic_ω start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 roman_ℏ end_ARG . (34)

with the corresponding PES displacement dbpsubscript𝑑bpd_{\text{bp}}italic_d start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT. Equation (33) states that sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT must be sufficiently small to mimic the sbp0subscript𝑠bp0s_{\text{bp}}\rightarrow 0italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT → 0 limit. In practice, however, reducing the value of sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT too much negatively affects the accuracy of the calculation due to the limited machine precision. Therefore it is required to seek the balance between these two aspects by checking the convergence of the calculation with different values of sbpsubscript𝑠bps_{\text{bp}}italic_s start_POSTSUBSCRIPT bp end_POSTSUBSCRIPT.

As a final remark, we note that the applicability of the approach illustrated in this Appendix is not limited to HEOM and is compatible with any numerically exact simulation methods for open quantum system dynamics,[53, 54, 55, 26, 56, 12] whenever it is not possible to construct the extended subsystem by directly using the target mode.

Appendix B Discretization of the Bath Spectral Densities

In this Appendix, we elaborate on how the BSDs are discretized for the MQME and MQME-D simulations. For the Drude-Lorentz spectral density [Eq. (7)], we follow Ref. 19, where the individual bath modes are placed at the frequencies

ωj=j2N2ωmax,j=1, 2,,N,formulae-sequencesubscript𝜔𝑗superscript𝑗2superscript𝑁2subscript𝜔max𝑗12𝑁\omega_{j}=\frac{j^{2}}{N^{2}}\omega_{\text{max}},\quad j=1,\>2,\>\cdots,\>N,italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = divide start_ARG italic_j start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT , italic_j = 1 , 2 , ⋯ , italic_N , (35)

with N𝑁Nitalic_N being the total number of bath modes in a spectral density, and ωmaxsubscript𝜔max\omega_{\text{max}}italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT the upper limit of the frequency. Equation (35) makes the bath modes more densely packed in the low-frequency region to reflect the increase of the reorganization energy density JDL(ω)/ωsubscript𝐽DL𝜔𝜔J_{\text{DL}}(\omega)/\omegaitalic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) / italic_ω therein. If we now define a function fDL(ω)subscript𝑓DL𝜔f_{\text{DL}}(\omega)italic_f start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) which connects the discretized [Eq. (5)] and continuous [Eq. (7)] forms of the BSD via

ωj3dj22=JDL(ωj)fDL(ωj),superscriptsubscript𝜔𝑗3superscriptsubscript𝑑𝑗22subscript𝐽DLsubscript𝜔𝑗subscript𝑓DLsubscript𝜔𝑗\frac{\omega_{j}^{3}d_{j}^{2}}{2}=\frac{J_{\text{DL}}(\omega_{j})}{f_{\text{DL% }}(\omega_{j})},divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG = divide start_ARG italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG start_ARG italic_f start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG , (36)

its explicit expression becomes

fDL(ω)=N2ωωmax,subscript𝑓DL𝜔𝑁2𝜔subscript𝜔maxf_{\text{DL}}(\omega)=\frac{N}{2\sqrt{\omega\omega_{\text{max}}}},italic_f start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG italic_N end_ARG start_ARG 2 square-root start_ARG italic_ω italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT end_ARG end_ARG , (37)

which renders the reorganization energies of the individual bath modes as

λj=ωj2dj22=4Λjπωcωjωj2+ωc2.subscript𝜆𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝑑𝑗224Λ𝑗𝜋subscript𝜔csubscript𝜔𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝜔c2\lambda_{j}=\frac{\omega_{j}^{2}d_{j}^{2}}{2}=\frac{4\Lambda}{j\pi}\frac{% \omega_{\text{c}}\omega_{j}}{\omega_{j}^{2}+\omega_{\text{c}}^{2}}.italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG = divide start_ARG 4 roman_Λ end_ARG start_ARG italic_j italic_π end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (38)

Meanwhile, the reorganization energy arising from the corresponding region of the continuous spectral density is

(ωj1+ωj)/2(ωj+ωj+1)/2JDL(ω)ω𝑑ωJDL(ωj)ωj[(ωj+ωj+12)(ωj1+ωj2)]=4Λjπωcωjωj2+ωc2,superscriptsubscriptsubscript𝜔𝑗1subscript𝜔𝑗2subscript𝜔𝑗subscript𝜔𝑗12subscript𝐽DL𝜔𝜔differential-d𝜔subscript𝐽DLsubscript𝜔𝑗subscript𝜔𝑗delimited-[]subscript𝜔𝑗subscript𝜔𝑗12subscript𝜔𝑗1subscript𝜔𝑗24Λ𝑗𝜋subscript𝜔csubscript𝜔𝑗superscriptsubscript𝜔𝑗2superscriptsubscript𝜔c2\begin{split}\int_{(\omega_{j-1}+\omega_{j})/2}^{(\omega_{j}+\omega_{j+1})/2}% \frac{J_{\text{DL}}(\omega)}{\omega}\>d\omega&\approx\frac{J_{\text{DL}}(% \omega_{j})}{\omega_{j}}\bigg{[}\bigg{(}\frac{\omega_{j}+\omega_{j+1}}{2}\bigg% {)}-\bigg{(}\frac{\omega_{j-1}+\omega_{j}}{2}\bigg{)}\bigg{]}\\ &=\frac{4\Lambda}{j\pi}\frac{\omega_{\text{c}}\omega_{j}}{\omega_{j}^{2}+% \omega_{\text{c}}^{2}},\end{split}start_ROW start_CELL ∫ start_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ) / 2 end_POSTSUPERSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) end_ARG start_ARG italic_ω end_ARG italic_d italic_ω end_CELL start_CELL ≈ divide start_ARG italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG [ ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) - ( divide start_ARG italic_ω start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = divide start_ARG 4 roman_Λ end_ARG start_ARG italic_j italic_π end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , end_CELL end_ROW (39)

where the approximation becomes exact at the limit of N𝑁N\rightarrow\inftyitalic_N → ∞. The equality between the last expressions of Eqs. (38) and (39) indicates the validity of fDL(ω)subscript𝑓DL𝜔f_{\text{DL}}(\omega)italic_f start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) defined by Eq. (37).

For the Brownian oscillator [Eq. (8)], we set ω0<ωmaxsubscript𝜔0subscript𝜔max\omega_{0}<\omega_{\text{max}}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT and calculate the frequency ΩΩ\Omegaroman_Ω where the reorganization energy density JBO/ωsubscript𝐽BO𝜔J_{\text{BO}}/\omegaitalic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT / italic_ω is maximized within [0,ωmax]0subscript𝜔max[0,\omega_{\text{max}}][ 0 , italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ], which is

Ω=max[0,ω022γ2].Ωmax0superscriptsubscript𝜔022superscript𝛾2\Omega=\sqrt{\text{max}[0,\omega_{0}^{2}-2\gamma^{2}]}.roman_Ω = square-root start_ARG max [ 0 , italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG . (40)

If Ω=0Ω0\Omega=0roman_Ω = 0, we can apply the same scheme as the Drude-Lorentz BSD [Eqs. (35)–(37)] with JDL(ω)subscript𝐽DL𝜔J_{\text{DL}}(\omega)italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ) in Eq. (36) replaced by JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ). Otherwise, we divide the frequency domain into two separate windows [0,Ω]0Ω[0,\Omega][ 0 , roman_Ω ] and (Ω,ωmax]Ωsubscript𝜔max(\Omega,\omega_{\text{max}}]( roman_Ω , italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ] and describe each region by using half of the bath modes under separate discretization schemes. This is achieved by setting the bath frequencies {ω1,j}subscript𝜔1𝑗\{\omega_{1,j}\}{ italic_ω start_POSTSUBSCRIPT 1 , italic_j end_POSTSUBSCRIPT } ({ω2,j}subscript𝜔2𝑗\{\omega_{2,j}\}{ italic_ω start_POSTSUBSCRIPT 2 , italic_j end_POSTSUBSCRIPT }) and the connecting function fBO1(ω)subscript𝑓BO1𝜔f_{\text{BO1}}(\omega)italic_f start_POSTSUBSCRIPT BO1 end_POSTSUBSCRIPT ( italic_ω ) [fBO2(ω)subscript𝑓BO2𝜔f_{\text{BO2}}(\omega)italic_f start_POSTSUBSCRIPT BO2 end_POSTSUBSCRIPT ( italic_ω )] for the former (latter) window as

ω1,j=[1(12jN)2]Ω,fBO1(ω)=N(Ωω)Ω,j=1, 2,,N21,formulae-sequencesubscript𝜔1𝑗delimited-[]1superscript12𝑗𝑁2Ωformulae-sequencesubscript𝑓BO1𝜔𝑁Ω𝜔Ω𝑗12𝑁21\omega_{1,j}=\bigg{[}1-\bigg{(}1-\frac{2j}{N}\bigg{)}^{2}\bigg{]}\Omega,\quad f% _{\text{BO1}}(\omega)=\frac{N}{\sqrt{(\Omega-\omega)\Omega}},\quad j=1,\>2,\>% \cdots,\>\frac{N}{2}-1,italic_ω start_POSTSUBSCRIPT 1 , italic_j end_POSTSUBSCRIPT = [ 1 - ( 1 - divide start_ARG 2 italic_j end_ARG start_ARG italic_N end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] roman_Ω , italic_f start_POSTSUBSCRIPT BO1 end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG italic_N end_ARG start_ARG square-root start_ARG ( roman_Ω - italic_ω ) roman_Ω end_ARG end_ARG , italic_j = 1 , 2 , ⋯ , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1 , (41a)
ω2,j=Ω+4j2N2(ωmaxΩ),fBO2(ω)=N(ωΩ)(ωmaxΩ),j=1, 2,,N2.formulae-sequencesubscript𝜔2𝑗Ω4superscript𝑗2superscript𝑁2subscript𝜔maxΩformulae-sequencesubscript𝑓BO2𝜔𝑁𝜔Ωsubscript𝜔maxΩ𝑗12𝑁2\omega_{2,j}=\Omega+\frac{4j^{2}}{N^{2}}(\omega_{\text{max}}-\Omega),\quad f_{% \text{BO2}}(\omega)=\frac{N}{\sqrt{(\omega-\Omega)(\omega_{\text{max}}-\Omega)% }},\quad j=1,\>2,\>\cdots,\>\frac{N}{2}.italic_ω start_POSTSUBSCRIPT 2 , italic_j end_POSTSUBSCRIPT = roman_Ω + divide start_ARG 4 italic_j start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT - roman_Ω ) , italic_f start_POSTSUBSCRIPT BO2 end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG italic_N end_ARG start_ARG square-root start_ARG ( italic_ω - roman_Ω ) ( italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT - roman_Ω ) end_ARG end_ARG , italic_j = 1 , 2 , ⋯ , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG . (41b)

Equation (41) lacks the case of ω=Ω𝜔Ω\omega=\Omegaitalic_ω = roman_Ω where both fBO1(ω)subscript𝑓BO1𝜔f_{\text{BO1}}(\omega)italic_f start_POSTSUBSCRIPT BO1 end_POSTSUBSCRIPT ( italic_ω ) and fBO2(ω)subscript𝑓BO2𝜔f_{\text{BO2}}(\omega)italic_f start_POSTSUBSCRIPT BO2 end_POSTSUBSCRIPT ( italic_ω ) diverge. Nevertheless, we can solve this issue by setting the reorganization energy of the discrete bath mode at ω=Ω𝜔Ω\omega=\Omegaitalic_ω = roman_Ω equal to that calculated from the continuous BSD JBO(ω)subscript𝐽BO𝜔J_{\text{BO}}(\omega)italic_J start_POSTSUBSCRIPT BO end_POSTSUBSCRIPT ( italic_ω ) [Eq. (8)] in the frequency window [(Ωω1,N/21)/2,(ω2,1Ω)/2]Ωsubscript𝜔1𝑁212subscript𝜔21Ω2[(\Omega-\omega_{1,N/2-1})/2,\>(\omega_{2,1}-\Omega)/2][ ( roman_Ω - italic_ω start_POSTSUBSCRIPT 1 , italic_N / 2 - 1 end_POSTSUBSCRIPT ) / 2 , ( italic_ω start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT - roman_Ω ) / 2 ]. As a result, we get

λω=Ω=2ΛπN2ωmaxω02γ(ω02γ2).subscript𝜆𝜔Ω2Λ𝜋superscript𝑁2subscript𝜔maxsuperscriptsubscript𝜔02𝛾superscriptsubscript𝜔02superscript𝛾2\lambda_{\omega=\Omega}=\frac{2\Lambda}{\pi N^{2}}\frac{\omega_{\text{max}}% \omega_{0}^{2}}{\gamma(\omega_{0}^{2}-\gamma^{2})}.italic_λ start_POSTSUBSCRIPT italic_ω = roman_Ω end_POSTSUBSCRIPT = divide start_ARG 2 roman_Λ end_ARG start_ARG italic_π italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_ω start_POSTSUBSCRIPT max end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_γ ( italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG . (42)

The discretization scheme defined by Eqs. (41) and (42) makes the bath modes more concentrated around ω=Ω𝜔Ω\omega=\Omegaitalic_ω = roman_Ω compared to the rest of the frequency domain. As a result, more emphasis is put on the region which contributes larger to the overall subsystem-bath coupling, as we did for JDL(ω)subscript𝐽DL𝜔J_{\text{DL}}(\omega)italic_J start_POSTSUBSCRIPT DL end_POSTSUBSCRIPT ( italic_ω ).

References

  • Kim and Franco [2024] C. W. Kim and I. Franco, “General Framework for Quantifying Dissipation Pathways in Open Quantum Systems. I. Theoretical Formulation,” J. Chem. Phys. 160, 214111 (2024).
  • Nakajima [1958] S. Nakajima, “On Quantum Theory of Transport Phenomena: Steady Diffusion,” Prog. Theor. Phys. 20, 948–959 (1958).
  • Zwanzig [1960] R. Zwanzig, “Ensemble Method in the Theory of Irreversibility,” J. Chem. Phys. 33, 1338–1341 (1960).
  • Kim and Franco [2021] C. W. Kim and I. Franco, “Theory of Dissipation Pathways in Open Quantum Systems,” J. Chem. Phys. 154, 084109 (2021).
  • Kim and Rhee [2020] C. W. Kim and Y. M. Rhee, “Toward Monitoring the Dissipative Vibrational Energy Flows in Open Quantum Systems by Mixed Quantum-Classical Simulations,” J. Chem. Phys. 152, 244109 (2020).
  • Nassimi, Bonella, and Kapral [2010] A. Nassimi, S. Bonella, and R. Kapral, “Analysis of the Quantum-Classical Liouville Equation in the Map** Basis,” J. Chem. Phys. 133, 134115 (2010).
  • Kim and Rhee [2014] H. W. Kim and Y. M. Rhee, “Improving Long Time Behavior of Poisson Bracket Map** Equation: A Non-Hamiltonian Approach,” J. Chem. Phys. 140, 184106 (2014).
  • Kim [2022] C. W. Kim, “Extracting Bath Information from Open-Quantum-System Dynamics with the Hierarchical Equations-of-Motion Method,” Phys. Rev. A 106, 042223 (2022).
  • Tanimura [1990] Y. Tanimura, “Nonperturbative Expansion Method for a Quantum System Coupled to a Harmonic-Oscillator Bath,” Phys. Rev. A 41, 6676–6687 (1990).
  • Tanimura [2020] Y. Tanimura, “Numerically “Exact” Approach to Open Quantum Dynamics: The Hierarchical Equations of Motion (HEOM),” J. Chem. Phys. 153, 020901 (2020).
  • Topaler and Makri [1993] M. Topaler and N. Makri, “Quasi-Adiabatic Propagator Path Integral Methods. Exact Quantum Rate Constants for Condensed Phase Reactions,” Chem. Phys. Lett. 210, 285–293 (1993).
  • Kundu and Makri [2023] S. Kundu and N. Makri, “PathSum: A C++ and Fortran suite of fully quantum mechanical real-time path integral methods for (multi-)system + bath dynamics,” J. Chem. Phys. 158, 224801 (2023).
  • Montoya-Castillo, Berkelbach, and Reichman [2015] A. Montoya-Castillo, T. C. Berkelbach, and D. R. Reichman, “Extending the Applicability of Redfield Theories into Highly Non-Markovian Regimes,” J. Chem. Phys. 143, 194108 (2015).
  • Ikeda and Scholes [2020] T. Ikeda and G. D. Scholes, “Generalization of the hierarchical equations of motion theory for efficient calculations with arbitrary correlation functions,” J. Chem. Phys. 152, 204101 (2020).
  • O’Reilly and Olaya-Castro [2014] E. J. O’Reilly and A. Olaya-Castro, “Non-Classicality of the Molecular Vibrations Assisting Exciton Energy Transfer at Room Temperature,” Nat. Commun. 5, 3012 (2014).
  • Novoderezhkin et al. [2017] V. I. Novoderezhkin, E. Romero, J. Prior, and R. van Grondelle, “Exciton-vibrational resonance and dynamics of charge separation in the photosystem ii reaction center,” Phys. Chem. Chem. Phys. 19, 5195–5208 (2017).
  • Bennett et al. [2018] D. I. G. Bennett, P. Malý, C. Kreisbeck, R. van Grondelle, and A. Aspuru-Guzik, “Mechanistic Regimes of Vibronic Transport in a Heterodimer and the Design Principle of Incoherent Vibronic Transport in Phycobiliproteins,” J. Phys. Chem. Lett. 9, 2665–2670 (2018).
  • Förster [1959] T. Förster, “10th Spiers Memorial Lecture. Transfer mechanisms of electronic excitation,” Discuss. Faraday Soc. 27, 7–17 (1959).
  • Wang et al. [1999] H. Wang, X. Song, D. Chandler, and W. H. Miller, “Semiclassical Study of Electronically Nonadiabatic Dynamics in the Condensed-Phase: Spin-Boson Problem with Debye Spectral Sensity,” J. Chem. Phys. 110, 4828–4840 (1999).
  • Fay [2022] T. P. Fay, “A Simple Improved Low Temperature Correction for the Hierarchical Equations of Motion,” J. Chem. Phys. 157, 054108 (2022).
  • Fehlberg [1969] E. Fehlberg, Low-order classical Runge-Kutta formulas with stepsize control and their application to some heat transfer problems, Vol. 315 (National Aeronautics and Space Administration, Huntsville, 1969).
  • Tanimura [2014] Y. Tanimura, “Reduced Hierarchical Equations of Motion in Real and Imaginary Time: Correlated Initial States and Thermodynamic Quantities,” J. Chem. Phys. 141, 044114 (2014).
  • Berkelbach, Markland, and Reichman [2012] T. C. Berkelbach, T. E. Markland, and D. R. Reichman, “Reduced Density Matrix Hybrid Approach: Application to Electronic Energy Transfer,” J. Chem. Phys. 136, 084104 (2012).
  • Ribeiro et al. [2018] R. F. Ribeiro, L. A. Martínez-Martínez, M. Du, J. Campos-Gonzalez-Angulo, and J. Yuen-Zhou, “Polariton chemistry: controlling molecular dynamics with optical cavities,” Chem. Sci. 9, 6325–6339 (2018).
  • Ebbesen, Rubio, and Scholes [2023] T. W. Ebbesen, A. Rubio, and G. D. Scholes, “Introduction: Polaritonic chemistry,” Chem. Rev. 123, 12037–12038 (2023).
  • Varvelo, Lynd, and Bennett [2021] L. Varvelo, J. K. Lynd, and D. I. G. Bennett, “Formally exact simulations of mesoscale exciton dynamics in molecular materials,” Chem. Sci. 12, 9704–9711 (2021).
  • Strümpfer and Schulten [2012] J. Strümpfer and K. Schulten, “Open Quantum Dynamics Calculations with the Hierarchy Equations of Motion on Parallel Computers,” J. Chem. Theory Comput. 8, 2808–2816 (2012).
  • Noack et al. [2018] M. Noack, A. Reinefeld, T. Kramer, and T. Steinke, “DM-HEOM: A Portable and Scalable Solver-Framework for the Hierarchical Equations of Motion,” in 2018 IEEE International Parallel and Distributed Processing Symposium Workshops (IPDPSW) (2018) pp. 947–956.
  • Urbanek and Soldán [2016] M. Urbanek and P. Soldán, “Parallel implementation of the time-evolving block decimation algorithm for the Bose–Hubbard model,” Comput. Phys. Commun. 199, 170–177 (2016).
  • Strathearn et al. [2018] A. Strathearn, P. Kirton, D. Kilda, J. Keeling, and B. W. Lovett, “Efficient non-markovian quantum dynamics using time-evolving matrix product operators,” Nat. Commun. 9, 3322 (2018).
  • Jang, Jung, and Silbey [2002] S. Jang, Y. Jung, and R. J. Silbey, “Nonequilibrium Generalization of Förster–Dexter Theory for Excitation Energy Transfer,” Chem. Phys. 275, 319–332 (2002).
  • Lee, Bravaya, and Coker [2017] M. K. Lee, K. B. Bravaya, and D. F. Coker, “First-Principles Models for Biological Light-Harvesting: Phycobiliprotein Complexes from Cryptophyte Algae,” J. Am. Chem. Soc. 139, 7803–7814 (2017).
  • Mennucci and Corni [2019] B. Mennucci and S. Corni, “Multiscale Modelling of Photoinduced Processes in Composite Systems,” Nat. Rev. Chem. 3, 315–330 (2019).
  • Connors et al. [2019] E. J. Connors, J. Nelson, H. Qiao, L. F. Edge, and J. M. Nichol, “Low-frequency charge noise in Si/SiGe quantum dots,” Phys. Rev. B 100, 165305 (2019).
  • Kim et al. [2023] Y. Kim, Z. Mitchell, J. Lawrence, D. Morozov, S. Savikhin, and L. V. Slipchenko, “Predicting Mutation-Induced Changes in the Electronic Properties of Photosynthetic Proteins from First Principles: The Fenna–Matthews–Olson Complex Example,” J. Phys. Chem. Lett. 14, 7038–7044 (2023).
  • Gustin et al. [2023] I. Gustin, C. W. Kim, D. W. McCamant, and I. Franco, “Map** Electronic Decoherence Pathways in Molecules,” Proc. Natl. Acad. Sci. U. S. A. 120, e2309987120 (2023).
  • Wilkins and Dattani [2015] D. M. Wilkins and N. S. Dattani, “Why Quantum Coherence Is Not Important in the Fenna–Matthews–Olsen Complex,” J. Chem. Theory Comput. 11, 3411–3419 (2015).
  • Blau et al. [2018] S. M. Blau, D. I. G. Bennett, C. Kreisbeck, G. D. Scholes, and A. Aspuru-Guzik, “Local Protein Solvation Drives Direct Down-Conversion in Phycobiliprotein PC645 via Incoherent Vibronic Transport,” Proc. Natl. Acad. Sci. U. S. A. 115, E3342–E3350 (2018).
  • Cardoso Ramos et al. [2019] F. Cardoso Ramos, M. Nottoli, L. Cupellini, and B. Mennucci, “The Molecular Mechanisms of Light Adaption in Light-Harvesting Complexes of Purple Bacteria Revealed by a Multiscale Modeling,” Chem. Sci. 10, 9650–9662 (2019).
  • Bolzonello, Fassioli, and Collini [2016] L. Bolzonello, F. Fassioli, and E. Collini, “Correlated Fluctuations and Intraband Dynamics of J-Aggregates Revealed by Combination of 2DES Schemes,” J. Phys. Chem. Lett. 7, 4996–5001 (2016).
  • Yang and Jang [2020] L. Yang and S. J. Jang, “Theoretical Investigation of Non-Förster Exciton Transfer Mechanisms in Perylene Diimide Donor, Phenylene Bridge, and Terrylene Diimide Acceptor Systems,” J. Chem. Phys. 153, 144305 (2020).
  • Bialas and Spano [2022] A. L. Bialas and F. C. Spano, “A Holstein–Peierls Approach to Excimer Spectra: The Evolution from Vibronically Structured to Unstructured Emission,” J. Phys. Chem. C 126, 4067–4081 (2022).
  • Hsu, Ding, and Schatz [2017] L.-Y. Hsu, W. Ding, and G. C. Schatz, “Plasmon-Coupled Resonance Energy Transfer,” J. Phys. Chem. Lett. 8, 2357–2367 (2017).
  • Bai et al. [2021] P. Bai, S. ter Huurne, E. van Heijst, S. Murai, and J. Gómez Rivas, “Evolutionary Optimization of Light-Matter Coupling in Open Plasmonic Cavities,” J. Chem. Phys. 154, 134110 (2021).
  • Gertler et al. [2021] J. M. Gertler, B. Baker, J. Li, S. Shirol, J. Koch, and C. Wang, “Protecting a Bosonic Qubit with Autonomous Quantum Error Correction,” Nature 590, 243–248 (2021).
  • Harrington, Mueller, and Murch [2022] P. M. Harrington, E. J. Mueller, and K. W. Murch, “Engineered Dissipation for Quantum Information Science,” Nat. Rev. Phys. 4, 660–671 (2022).
  • Chiesa et al. [2023] A. Chiesa, A. Privitera, E. Macaluso, M. Mannini, R. Bittl, R. Naaman, M. R. Wasielewski, R. Sessoli, and S. Carretta, “Chirality-induced spin selectivity: An enabling technology for quantum applications,” Adv. Mater. 35, 2300472 (2023).
  • Zhu et al. [2012] L. Zhu, H. Liu, W. Xie, and Q. Shi, “Explicit system-bath correlation calculated using the hierarchical equations of motion method,” J. Chem. Phys. 137, 194106 (2012).
  • Kato and Tanimura [2016] A. Kato and Y. Tanimura, “Quantum heat current under non-perturbative and non-Markovian conditions: Applications to heat machines,” J. Chem. Phys. 145, 224105 (2016).
  • Ikeda and Nakayama [2022] T. Ikeda and A. Nakayama, “Collective bath coordinate map** of “hierarchy” in hierarchical equations of motion,” J. Chem. Phys. 156, 104104 (2022).
  • Womick and Moran [2011] J. M. Womick and A. M. Moran, “Vibronic enhancement of exciton sizes and energy transport in photosynthetic complexes,” J. Phys. Chem. B 115, 1347–1356 (2011).
  • Yan et al. [2021] Y. Yan, M. Xu, T. Li, and Q. Shi, “Efficient propagation of the hierarchical equations of motion using the Tucker and hierarchical Tucker tensors,” J. Chem. Phys. 154, 194104 (2021).
  • Diósi, Gisin, and Strunz [1998] L. Diósi, N. Gisin, and W. T. Strunz, “Non-markovian quantum state diffusion,” Phys. Rev. A 58, 1699–1712 (1998).
  • Yan [2014] Y. Yan, “Theory of open quantum systems with bath of electrons and phonons and spins: Many-dissipaton density matrixes approach,” J. Chem. Phys. 140, 054105 (2014).
  • Tamascelli et al. [2019] D. Tamascelli, A. Smirne, J. Lim, S. F. Huelga, and M. B. Plenio, “Efficient simulation of finite-temperature open quantum systems,” Phys. Rev. Lett. 123, 090402 (2019).
  • Cygorek et al. [2022] M. Cygorek, M. Cosacchi, A. Vagov, V. M. Axt, B. W. Lovett, J. Keeling, and E. M. Gauger, “Simulation of open quantum systems by automated compression of arbitrary environments,” Nat. Phys. 18, 662–668 (2022).