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

Quasi-Lindblad pseudomode theory for open quantum systems

Gunhee Park (박건희) Division of Engineering and Applied Science, California Institute of Technology, Pasadena, California 91125, USA    Zhen Huang Department of Mathematics, University of California, Berkeley, California 94720, USA    Yuanran Zhu Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Chao Yang Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Garnet Kin-Lic Chan Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA    Lin Lin Department of Mathematics, University of California, Berkeley, California 94720, USA Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA Division of Engineering and Applied Science, California Institute of Technology, Pasadena, California 91125, USA Department of Mathematics, University of California, Berkeley, California 94720, USA Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA Department of Mathematics, University of California, Berkeley, California 94720, USA Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA
Abstract

We introduce a new framework to study the dynamics of open quantum systems with linearly coupled Gaussian baths. Our approach replaces the continuous bath with an auxiliary discrete set of pseudomodes with dissipative dynamics, but we further relax the complete positivity requirement in the Lindblad master equation and formulate a quasi-Lindblad pseudomode theory. We show that this quasi-Lindblad pseudomode formulation directly leads to a representation of the bath correlation function in terms of a complex weighted sum of complex exponentials, an expansion that is known to be rapidly convergent in practice and thus leads to a compact set of pseudomodes. The pseudomode representation is not unique and can differ by a gauge choice. When the global dynamics can be simulated exactly, the system dynamics is unique and independent of the specific pseudomode representation. However, the gauge choice may affect the stability of the global dynamics, and we provide an analysis of why and when the global dynamics can retain stability despite losing positivity. We showcase the performance of this formulation across various spectral densities in both bosonic and fermionic problems, finding significant improvements over conventional pseudomode formulations.

I Introduction

Open quantum systems (OQS) are central to many fields, including quantum optics, condensed matter physics, chemical physics, and quantum information science [1, 2, 3, 4]. The most widely studied case corresponds to a system coupled to a bath with continuous (bosonic or fermionic) degrees of freedom with the bath dynamics governed by a Hamiltonian quadratic in the bath field operators (a Gaussian bath) and with a system-bath coupling linear in the bath field operators. The primary task in such setups is to compute the system’s dynamical observables.

There are two main classes of approaches to simulate this problem. The first approach integrates out the Gaussian bath dynamics to yield a path integral for the reduced system dynamics [5], which can then be evaluated using various approximations [6, 7, 8, 9, 10, 11, 12, 13]. The second approach, which is the focus of this work, replaces the continuous physical bath with a discrete mode representation [14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25], i.e., an auxiliary bath, allowing the global state of the system and auxiliary bath to be computationally propagated in time. In both approaches, the influence of the bath on the reduced system dynamics is entirely determined by the bath correlation function (BCF) [5, 2]. As long as the auxiliary BCF matches the original BCF, the reduced system dynamics can be shown to be identical [20]. Therefore, the primary consideration in this second approach is to begin with a compact auxiliary bath representation of the BCF.

The conventional discrete mode representation is to directly discretize the continuous Hamiltonian [14, 15, 16] without introducing any dissipation. However, this discretization requires a large number of modes in the long-time limit, which scales linearly with the propagated time [14]. In most physical cases, the continuous bath provides dissipation to the system. It is thus natural to introduce an ansatz where the auxiliary bath dynamics is dissipative, in which case, the bath modes are referred to as pseudomodes [18, 19, 20, 26, 27, 28]. The Liouvillian for the system and auxiliary bath can then be chosen to satisfy the constraint of physical dynamics, taking a Lindblad form and generating a completely positive trace-preserving (CPTP) dynamics. The corresponding BCF (assuming independent pseudomodes) is then a sum of complex exponential functions weighted by positive numbers. This BCF exactly represents the continuous spectral density as a positive weighted sum of Lorentzian functions. Unfortunately, many studies have found that this natural Lorentzian pseudomode expansion is inefficient, achieving only a slow rate of convergence as the number of pseudomodes increases [26, 25, 29]. Improving this convergence rate remains an active area of research, with approaches including the introduction of additional intra-bath Hamiltonian terms [26, 29, 30, 31] or non-Hermitian pseudomode formulations [32, 33, 34, 35].

In the current work, we demonstrate that by relaxing the complete positivity (CP) of the global dynamics of the system and pseudomodes, we gain the flexibility to use a broader set of functions to approximate the BCF, significantly enhancing the convergence rate of the pseudomode expansion. The generator of the dynamics is a slight modification of the Lindblad form, which we refer to as a quasi-Lindblad pseudomode representation. The BCF is expressed as a sum of complex exponentials weighted by complex numbers. This ansatz is well-known in the signal processing and applied mathematics communities for yielding an exponentially fast convergence rate for many functions [36, 37, 38, 39], a result we also observe in our numerical tests. Additionally, this ansatz for the BCF has been adopted in non-pseudomode approaches to OQS dynamics, such as hierarchical equations of motion (HEOM) [40, 41, 42, 43, 44] and tensor network influence functionals [45, 46], providing a connection among these various methods.

It is worth noting that the BCF alone does not uniquely determine the pseudomode formulation. Meanwhile, the system dynamics is uniquely determined by the BCF and is independent of the pseudomode formulation as long as the global dynamics can be simulated exactly (i.e., without further numerical approximations such as truncating the number of bosonic modes or even finite precision arithmetic operations). However, certain pseudomode formulations can be unstable if the corresponding Liouvillian has eigenvalues with positive real parts, leading to exponentially growing modes. This phenomenon has also been observed in other contexts, such as HEOM [47, 48, 49, 50]. Therefore, in practice, the stability of the pseudomode formulation can indeed impact the quality of system dynamics. Nevertheless, with a careful choice of the pseudomode formulation that minimally violates the CP condition, we observe that the global dynamics can be stable across a range of physically relevant OQS in our simulations. We attribute this stability to the mixing of the divergent and dissipative parts of the global dynamics via the coherent terms in the Liouvillian.

Refer to caption
Figure 1: Pseudomode theory defines a discrete auxiliary dissipative bath AA to reproduce a system-reduced density operator. (a) The conventional pseudomode assumes bath-only dissipators, 𝒟A\mathcal{D}_{A}. (b) Our quasi-Lindblad pseudomode relaxes the complete positivity condition by introducing an additional system-bath Lindblad coupling with dissipators, 𝒟SA\mathcal{D}_{SA}.

II Quasi-Lindblad pseudomode construction

II.1 Problem setup: original unitary OQS

We consider a quantum system SS linearly coupled to a continuous Gaussian bath BB, either bosonic or fermionic, evolving under the following Hamiltonian,

H^phy\displaystyle\hat{H}_{\text{phy}} =H^S+H^B+H^SB\displaystyle=\hat{H}_{S}+\hat{H}_{B}+\hat{H}_{SB}
=H^S+𝑑ωωc^ωc^ω+jS^jB^j,\displaystyle=\hat{H}_{S}+\int\!d\omega\ \omega\ \hat{c}_{\omega}^{\dagger}\hat{c}_{\omega}+\sum_{j}\hat{S}_{j}\hat{B}_{j}, (1)

where H^S\hat{H}_{S} denotes the system-only Hamiltonian and =1\hbar=1. c^ω\hat{c}_{\omega}^{\dagger} denotes the creation operator for the bath mode with energy ω\omega. In the system-bath coupling term, H^SB\hat{H}_{SB}, S^j\hat{S}_{j} and B^j\hat{B}_{j} are operators within the system and bath, respectively, and can be assumed to be Hermitian operators, without loss of generality [1]. The bath coupling operator is given by B^j=𝑑ωgj(ω)c^ω+gj(ω)c^ω\hat{B}_{j}=\int d\omega\ g_{j}(\omega)\hat{c}_{\omega}+g_{j}^{*}(\omega)\hat{c}_{\omega}^{\dagger}. We assume the initial state to take a factorized form, ρ^SB(0)=ρ^S(0)ρ^B(0)\hat{\rho}_{SB}(0)=\hat{\rho}_{S}(0)\otimes\hat{\rho}_{B}(0), where ρ^B(0)\hat{\rho}_{B}(0) is a Gaussian state. We will also assume that ρ^B(0)\hat{\rho}_{B}(0) is a number-conserving operator (and also for all initial bath reduced density operators henceforth) for simplicity.

Time-evolution of the density operator is described by the von Neumann equation, tρ^SB=i[H^phy,ρ^SB]\partial_{t}\hat{\rho}_{SB}=-i[\hat{H}_{\text{phy}},\hat{\rho}_{SB}], and the system-reduced density operator is obtained by tracing out the bath, ρ^S(t)=TrB[ρ^SB(t)]\hat{\rho}_{S}(t)=\operatorname{Tr}_{B}[\hat{\rho}_{SB}(t)]. The influence of the Gaussian bath on the system-reduced density operator is determined by the bath correlation function (BCF),

CjjB(t,t)=TrB[B^j(t)B^j(t)ρ^B(0)],C_{jj^{\prime}}^{B}(t,t^{\prime})=\operatorname{Tr}_{B}\left[\hat{B}_{j}(t)\hat{B}_{j^{\prime}}(t^{\prime})\hat{\rho}_{B}(0)\right], (2)

for tt0t\geq t^{\prime}\geq 0 with B^j(t)=eiH^BtB^jeiH^Bt\hat{B}_{j}(t)=e^{i\hat{H}_{B}t}\hat{B}_{j}e^{-i\hat{H}_{B}t}. When ρ^B(0)\hat{\rho}_{B}(0) is stationary, i.e., eiH^Btρ^B(0)eiH^Bt=ρ^B(0)e^{-i\hat{H}_{B}t}\hat{\rho}_{B}(0)e^{i\hat{H}_{B}t}=\hat{\rho}_{B}(0), the BCF only depends on the time difference ttt-t^{\prime}, i.e., CjjB(t,t)=CjjB(tt)C_{jj}^{B}(t,t^{\prime})=C_{jj}^{B}(t-t^{\prime}).

II.2 Conventional pseudomode

In many cases of interest, the above unitary dynamics using a continuous bath leads to a dissipative bath correlation function, i.e., a decaying CjjB(tt)C_{jj}^{B}(t-t^{\prime}) with limttCjjB(tt)0\lim_{t-t^{\prime}\to\infty}C_{jj}^{B}(t-t^{\prime})\to 0. To reproduce the associated dynamics of ρ^S(t)\hat{\rho}_{S}(t) with a discrete bath, we now introduce a pseudomode description. First, we review the conventional pseudomode formulation [20], which introduces an auxiliary bosonic bath AA with a Lindblad dissipator 𝒟A\mathcal{D}_{A} applied within the bath AA. The density operator across the system and auxiliary bath, ρ^SA\hat{\rho}_{SA}, evolves under the following Lindblad master equation,

tρ^SA=i[H^aux,ρ^SA]+𝒟Aρ^SA,\partial_{t}\hat{\rho}_{SA}=-i\left[\hat{H}_{\text{aux}},\hat{\rho}_{SA}\right]+\mathcal{D}_{A}\hat{\rho}_{SA}, (3)

where the Hamiltonian H^aux=H^S+H^A+H^SA\hat{H}_{\text{aux}}=\hat{H}_{S}+\hat{H}_{A}+\hat{H}_{SA} is composed of the system (bath)-only Hamiltonian H^S\hat{H}_{S} (H^A\hat{H}_{A}) and coupling Hamiltonian H^SA=jS^jA^j\hat{H}_{SA}=\sum_{j}\hat{S}_{j}\hat{A}_{j}. 𝒟A\mathcal{D}_{A} has a Lindblad form

𝒟A=μL^μL^μ12{L^μL^μ,},\mathcal{D}_{A}\bullet=\sum_{\mu}\hat{L}_{\mu}\bullet\hat{L}_{\mu}^{\dagger}-\frac{1}{2}\{\hat{L}_{\mu}^{\dagger}\hat{L}_{\mu},\bullet\}, (4)

which guarantees the CPTP property of the global dynamics consisting of the system and the pseudomodes. Here, L^μ\hat{L}_{\mu} can be an arbitrary operator within the bath AA. The Lindblad dissipator can be expressed in terms of a pseudomode basis,

𝒟A=2kkΓkk(F^kF^k12{F^kF^k,}),\mathcal{D}_{A}\bullet=2\sum_{kk^{\prime}}\Gamma_{kk^{\prime}}\left(\hat{F}_{k^{\prime}}\bullet\hat{F}_{k}^{\dagger}-\frac{1}{2}\{\hat{F}_{k}^{\dagger}\hat{F}_{k^{\prime}},\bullet\}\right), (5)

where F^k\hat{F}_{k} is an operator within the kkth pseudomode, and Γkk\Gamma_{kk^{\prime}} is a positive semidefinite (PSD) matrix from the CPTP condition. H^A\hat{H}_{A} and 𝒟A\mathcal{D}_{A} are quadratic in the creation and annihilation operators of the pseudomode, and Aj^\hat{A_{j}} and F^k\hat{F}_{k} are linear, to maintain the Gaussian properties of the bath. 𝒟A\mathcal{D}_{A} has negative eigenvalues, and thus, we can interpret it as adding dissipation to the reduced system dynamics.

The BCF for the pseudomode can be defined using the bath-only Liouvillian A=i[H^A,]+𝒟A\mathcal{L}_{A}\bullet=-i[\hat{H}_{A},\bullet]+\mathcal{D}_{A}\bullet and initial pseudomode reduced density operator ρ^A(0)\hat{\rho}_{A}(0),

CjjA(t,t)=TrA[A^jeA(tt)A^jeAtρ^A(0)].C_{jj^{\prime}}^{A}(t,t^{\prime})=\operatorname{Tr}_{A}\left[\hat{A}_{j}e^{\mathcal{L}_{A}(t-t^{\prime})}\hat{A}_{j^{\prime}}e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)\right]. (6)

Ref. [20] established the equivalence condition between the unitary bath dynamics and the pseudomode dynamics in terms of the BCF: if CjjB(t,t)=CjjA(t,t)C^{B}_{jj^{\prime}}(t,t^{\prime})=C^{A}_{jj^{\prime}}(t,t^{\prime}) for tt0\forall t\geq t^{\prime}\geq 0, then TrB[ρ^SB(t)]=TrA[ρ^SA(t)]\operatorname{Tr}_{B}[\hat{\rho}_{SB}(t)]=\operatorname{Tr}_{A}[\hat{\rho}_{SA}(t)] 111In the original formulation in [20], the additional conditions on the one-point bath expectation values Tr[B^j(t)ρ^B(0)]\operatorname{Tr}[\hat{B}_{j}(t)\hat{\rho}_{B}(0)] were imposed. In the main text, we further assumed that the initial bath reduced density operators, ρ^B(0)\hat{\rho}_{B}(0) and ρ^A(0)\hat{\rho}_{A}(0), are number-conserving operators. It leads to the one-point bath expectation values becoming zero for both the original unitary and pseudomode formulations.. As in the unitary case in Sec. II.1, when ρ^A(0)\hat{\rho}_{A}(0) is stationary, i.e., eAtρ^A(0)=ρ^A(0)e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)=\hat{\rho}_{A}(0), the BCF depends solely on ttt-t^{\prime}.

II.3 Quasi-Lindblad pseudomode

As described in the introduction, the conventional pseudomode description can reproduce a dissipative bath correlation function within a finite discretization, but it typically requires a large number of pseudomodes for an accurate approximation to the BCF. To remedy this, we now provide a more general pseudomode description that introduces an additional system-bath coupling 𝒟SA\mathcal{D}_{SA}, expressed in terms of Lindblad dissipators,

𝒟SA=jL^jS^j+S^jL^j12{S^jL^j+L^jS^j,},\mathcal{D}_{SA}\bullet=\sum_{j}\hat{L}^{\prime}_{j}\bullet\hat{S}_{j}+\hat{S}_{j}\bullet\hat{L}^{\prime{\dagger}}_{j}-\frac{1}{2}\{\hat{S}_{j}\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j}\hat{S}_{j},\bullet\}, (7)

where L^j=k2MjkF^k\hat{L}^{\prime}_{j}=\sum_{k}2M_{jk}\hat{F}_{k}. We call this coupling a Lindblad coupling, as illustrated in Fig. 1b. The total dissipator 𝒟=𝒟A+𝒟SA\mathcal{D}=\mathcal{D}_{A}+\mathcal{D}_{SA} can be compactly written as

𝒟=2pqΓ~pq(F^qF^p12{F^pF^q,}),\mathcal{D}\bullet=2\sum_{pq}\widetilde{\Gamma}_{pq}\left(\hat{F}_{q}\bullet\hat{F}^{\dagger}_{p}-\frac{1}{2}\{\hat{F}^{\dagger}_{p}\hat{F}_{q},\bullet\}\right), (8)

where the index pp and qq refer to both the coupling index jj and the pseudomode bath index kk, i.e., p{j}{k}p\in\{j\}\cup\{k\}, and if p=jp=j, F^j=S^j\hat{F}_{j}=\hat{S}_{j}. The coefficient Γ~pq\widetilde{\Gamma}_{pq} is

𝚪~=(𝟎𝑴𝑴𝚪),\widetilde{\bm{\Gamma}}=\left(\begin{array}[]{cc}\bm{0}&\bm{M}\\ \bm{M}^{\dagger}&\bm{\Gamma}\end{array}\right), (9)

where the bold font denotes matrices. We note that the matrix Γ~pq\widetilde{\Gamma}_{pq} is not PSD in general, unlike the PSD matrix Γpq\Gamma_{pq}, and hence, this pseudomode equation of motion violates the CP condition in the Lindblad master equation. Therefore, we call it a quasi-Lindblad pseudomode representation.

To take into account both the Hamiltonian and Lindblad coupling in the BCF, we define the system-bath Liouvillian SA=i[H^SA,]+𝒟SA\mathcal{L}_{SA}\bullet=-i[\hat{H}_{SA},\bullet]+\mathcal{D}_{SA}\bullet and further decompose it in the following form,

SA=ij𝒮jj+ij𝒮~j~j,\mathcal{L}_{SA}=-i\sum_{j}\mathcal{S}_{j}\mathcal{F}_{j}+i\sum_{j}\widetilde{\mathcal{S}}_{j}\widetilde{\mathcal{F}}_{j}, (10)

where the system superoperators, 𝒮j\mathcal{S}_{j} and 𝒮~j\widetilde{\mathcal{S}}_{j} are defined as 𝒮j=S^j\mathcal{S}_{j}\bullet=\hat{S}_{j}\bullet and 𝒮~j=S^j\widetilde{\mathcal{S}}_{j}\bullet=\bullet\hat{S}_{j}. The superoperators, j\mathcal{F}_{j} and ~j\widetilde{\mathcal{F}}_{j}, are obtained as,

j=A^j+iL^ji2(L^j+L^j),~j=A^jiL^j+i2(L^j+L^j).\begin{gathered}\mathcal{F}_{j}\bullet=\hat{A}_{j}\bullet+\bullet i\hat{L}^{\prime{\dagger}}_{j}-\frac{i}{2}(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j})\bullet,\\ \widetilde{\mathcal{F}}_{j}\bullet=\bullet\hat{A}_{j}-i\hat{L}^{\prime}_{j}\bullet+\bullet\frac{i}{2}(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j}).\end{gathered} (11)

The BCF is defined with the superoperators j\mathcal{F}_{j},

CjjA(t,t)=TrA[jeA(tt)jeAtρ^A(0)].C_{jj^{\prime}}^{A}(t,t^{\prime})=\operatorname{Tr}_{A}\left[\mathcal{F}_{j}e^{\mathcal{L}_{A}(t-t^{\prime})}\mathcal{F}_{j^{\prime}}e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)\right]. (12)

We remark that j\mathcal{F}_{j} reduces to A^j\hat{A}_{j} when L^j=0\hat{L}^{\prime}_{j}=0, and then the BCF in Eq. 12 also reduces to the BCF in Eq. 6. Similar to the pseudomode description in Sec. II.2, the above BCF leads to the equivalence of the reduced system dynamics to that generated by the original unitary formulation under the equivalence of the BCFs. We provide the derivation of the equivalence condition in the Supplemental Material (SM) [52].

In this framework, violation of the CP condition leads the dissipator Liouvillian 𝒟\mathcal{D} to have eigenvalues with positive real parts. Thus, in the absence of any coherent term (i.e., the commutator term of the form i[H^aux,]-i[\hat{H}_{\rm aux},\bullet] for some Hamiltonian H^aux\hat{H}_{\rm aux}), this would lead to unstable global dynamics with modes growing exponentially in time. Note that when simulations are performed exactly, the reduced system dynamics, after tracing out the bath, can still be perfectly reproduced from the BCF equivalence condition. However, this cannot be guaranteed in practical simulations involving finite precision arithmetic operations and physical approximations, such as truncating the number of bosonic modes. Therefore, the stability of the Liouvillian is an important issue to study.

Interestingly, we find that the Liouvillian, including both the coherent and dissipator terms, can generate stable dynamics despite violating the CP condition. We will analyze this effect in more detail in Sec. IV.

II.4 Ansatz for quasi-Lindblad pseudomode

We next introduce an ansatz for the quasi-Lindblad pseudomode that yields a simple form for the BCF, facilitating numerical fitting. We describe this ansatz for both bosonic and fermionic baths. We mainly target the reproduction of the BCF of the unitary dynamics in Sec. II.1 with the initial bath being a stationary state, such as a thermal state ρ^B(0)eβH^B\hat{\rho}_{B}(0)\propto e^{-\beta\hat{H}_{B}}, where β\beta represents the inverse temperature.

We start with a bosonic bath, setting the initial reduced density operator to a vacuum state, ρ^A(0)=|𝟎𝟎|\hat{\rho}_{A}(0)=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}. The dissipator 𝒟A\mathcal{D}_{A} is set as F^k=d^k\hat{F}_{k}=\hat{d}_{k}, where d^k\hat{d}_{k} is the bosonic annihilation operator of the kkth pseudomode, and the Hamiltonian H^A=kkHkkAd^kd^k\hat{H}_{A}=\sum_{kk^{\prime}}H^{A}_{kk^{\prime}}\hat{d}_{k}^{\dagger}\hat{d}_{k^{\prime}}. This ansatz satisfies the stationary condition Aρ^A(0)=0\mathcal{L}_{A}\hat{\rho}_{A}(0)=0. Then, eAtρ^A(0)=ρ^A(0)e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)=\hat{\rho}_{A}(0) and CjjA(t,t)C^{A}_{jj^{\prime}}(t,t^{\prime}) is only dependent on the time difference CjjA(t,t)=CjjA(tt)=CjjA(Δt)C^{A}_{jj^{\prime}}(t,t^{\prime})=C^{A}_{jj^{\prime}}(t-t^{\prime})=C^{A}_{jj^{\prime}}(\Delta t), where Δt=tt\Delta t=t-t^{\prime}. The system-bath coupling is set as A^j=kVjkd^k+Vjkd^k\hat{A}_{j}\ =\sum_{k}V_{jk}\hat{d}_{k}+V_{jk}^{*}\hat{d}_{k}^{\dagger}.

We obtain the BCF based on the above ansatz:

𝑪A(Δt)=(𝑽i𝑴)e𝒁Δt(𝑽+i𝑴),\bm{C}^{A}(\Delta t)=(\bm{V}-i\bm{M})e^{-\bm{Z}\Delta t}(\bm{V}+i\bm{M})^{\dagger}, (13)

where 𝒁=𝚪+i𝑯A\bm{Z}=\bm{\Gamma}+i\bm{H}^{A}. This BCF expression has a gauge redundancy: 𝒁𝑮𝒁𝑮1\bm{Z}\rightarrow\bm{G}\bm{Z}\bm{G}^{-1}, 𝑽i𝑴(𝑽i𝑴)𝑮1\bm{V}-i\bm{M}\rightarrow(\bm{V}-i\bm{M})\bm{G}^{-1}, 𝑽+i𝑴(𝑽+i𝑴)𝑮\bm{V}+i\bm{M}\rightarrow(\bm{V}+i\bm{M})\bm{G}^{\dagger}, for arbitrary invertible matrix 𝑮\bm{G}. Therefore, we focus on a diagonal matrix 𝒁=diag({zk})\bm{Z}=\mathrm{diag}(\{z_{k}\}), which can be transformed to a more general diagonalizable 𝒁\bm{Z} through a unitary transformation 222In principle, we have a general expression by considering non-diagonalizable 𝒁\bm{Z}. However, numerically, fitting with arbitrary 𝒁\bm{Z} yields the same optimal solution as with diagonalizable 𝒁\bm{Z}.. In this case, we define H^A=kEkd^kd^k\hat{H}_{A}=\sum_{k}E_{k}\hat{d}^{\dagger}_{k}\hat{d}_{k} and Γkk=γkδkk\Gamma_{kk^{\prime}}=\gamma_{k}\delta_{kk^{\prime}}, and zk=γk+iEkz_{k}=\gamma_{k}+iE_{k}. Even with a diagonal matrix 𝒁\bm{Z}, 𝑽\bm{V} and 𝑴\bm{M} still have a gauge redundancy associated with a diagonal 𝑮\bm{G}, since 𝒁\bm{Z} remains invariant under the transformation 𝑮𝒁𝑮1=𝒁\bm{G}\bm{Z}\bm{G}^{-1}=\bm{Z}. It is important to note that while different choices of gauge yield the same BCF and thus the same reduced system dynamics (in the absence of simulation errors), the total system-bath Liouvillian differs, leading to different global system-bath dynamics.

The case of a fermionic bath can be introduced similarly. In particular, we focus on number-conserving fermionic impurity models where the coupling Hamiltonian is given by H^SB=ja^jB^j+B^ja^j,B^j=𝑑ωgj(ω)c^ω\hat{H}_{SB}=\sum_{j}\hat{a}_{j}^{\dagger}\hat{B}_{j}+\hat{B}_{j}^{\dagger}\hat{a}_{j},\>\hat{B}_{j}=\int d\omega\ g_{j}(\omega)\hat{c}_{\omega} (a^j\hat{a}_{j}^{\dagger} is the creation operator of the impurity with index jj, which can involve both spin and orbital degrees of freedom). The number-conserving property of the fermionic impurity model reduces the BCFs to two different BCFs, namely the greater and lesser BCFs: Cjj>BC^{>B}_{jj^{\prime}} and Cjj<BC^{<B}_{jj^{\prime}}. For any fermionic Gaussian state ρ^B(0)\hat{\rho}_{B}(0), Cjj>BC^{>B}_{jj^{\prime}} and Cjj<BC^{<B}_{jj^{\prime}} are defined as,

Cjj>B(t,t)=TrB[B^j(t)B^j(t)ρ^B(0)],Cjj<B(t,t)=TrB[B^j(t)B^j(t)ρ^B(0)].\begin{gathered}C^{>B}_{jj^{\prime}}(t,t^{\prime})=\operatorname{Tr}_{B}\left[\hat{B}_{j}(t)\hat{B}^{{\dagger}}_{j^{\prime}}(t^{\prime})\hat{\rho}_{B}(0)\right],\\ C^{<B}_{jj^{\prime}}(t,t^{\prime})=\operatorname{Tr}_{B}\left[\hat{B}^{\dagger}_{j}(t)\hat{B}_{j^{\prime}}(t^{\prime})\hat{\rho}_{B}(0)\right].\end{gathered} (14)

Corresponding to these two BCFs, we use two different auxiliary baths with an initial reduced density operator ρ^A(0)=ρ^A1(0)ρ^A2(0)=|𝟎𝟎||𝟏𝟏|\hat{\rho}_{A}(0)=\hat{\rho}_{A_{1}}(0)\otimes\hat{\rho}_{A_{2}}(0)=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}\otimes\mathinner{|{\bm{1}}\rangle}\!\!\mathinner{\langle{\bm{1}}|}. The bath-only Hamiltonian and dissipator are set to be diagonal without loss of generality, as discussed in the bosonic case. The Hamiltonian is H^A=k1Ek1d^k1d^k1+k2Ek2d^k2d^k2\hat{H}_{A}=\sum_{k_{1}}E_{k_{1}}\hat{d}^{\dagger}_{k_{1}}\hat{d}_{k_{1}}+\sum_{k_{2}}E_{k_{2}}\hat{d}^{\dagger}_{k_{2}}\hat{d}_{k_{2}} where k1k_{1} and k2k_{2} refer to the pseudomode index in each bath, respectively, and the dissipator is set to be

𝒟A1=k12γk1(d^k1d^k112{d^k1d^k1,}),𝒟A2=k22γk2(d^k2d^k212{d^k2d^k2,}),\begin{gathered}\mathcal{D}_{A_{1}}\bullet=\sum_{k_{1}}2\gamma_{k_{1}}\left(\hat{d}_{k_{1}}\bullet\hat{d}^{{\dagger}}_{k_{1}}-\frac{1}{2}\{\hat{d}^{{\dagger}}_{k_{1}}\hat{d}_{k_{1}},\bullet\}\right),\\ \mathcal{D}_{A_{2}}\bullet=\sum_{k_{2}}2\gamma_{k_{2}}\left(\hat{d}^{\dagger}_{k_{2}}\bullet\hat{d}_{k_{2}}-\frac{1}{2}\{\hat{d}_{k_{2}}\hat{d}^{\dagger}_{k_{2}},\bullet\}\right),\end{gathered} (15)

which satisfies 𝒟Aiρ^Ai(0)=0\mathcal{D}_{A_{i}}\hat{\rho}_{A_{i}}(0)=0 for both i=1,2i=1,2. The system-bath Hamiltonian coupling is H^SA=a^jA^j+A^ja^j\hat{H}_{SA}=\hat{a}_{j}^{\dagger}\hat{A}_{j}+\hat{A}_{j}^{\dagger}\hat{a}_{j}, A^j=k1(𝑽1)jk1d^k1+k2(𝑽2)jk2d^k2\hat{A}_{j}=\sum_{k_{1}}(\bm{V}_{1})_{jk_{1}}\hat{d}_{k_{1}}+\sum_{k_{2}}(\bm{V}_{2})_{jk_{2}}\hat{d}_{k_{2}} and the system-bath Lindblad coupling is set to be

𝒟SA1=ja^jL^1j+L^1ja^j12{L^1ja^j+a^jL^1j,},𝒟SA2=ja^jL^2j+L^2ja^j12{L^2ja^j+a^jL^2j,},\begin{gathered}\mathcal{D}_{SA_{1}}\bullet=\sum_{j}\hat{a}_{j}\bullet\hat{L}_{1j}^{\dagger}+\hat{L}_{1j}\bullet\hat{a}_{j}^{\dagger}-\frac{1}{2}\{\hat{L}_{1j}^{\dagger}\hat{a}_{j}+\hat{a}_{j}^{\dagger}\hat{L}_{1j},\bullet\},\\ \mathcal{D}_{SA_{2}}\bullet=\sum_{j}\hat{a}^{\dagger}_{j}\bullet\hat{L}_{2j}+\hat{L}^{\dagger}_{2j}\bullet\hat{a}_{j}-\frac{1}{2}\{\hat{L}_{2j}\hat{a}^{\dagger}_{j}+\hat{a}_{j}\hat{L}^{\dagger}_{2j},\bullet\},\end{gathered} (16)

where L^1j=k12(𝑴1)jk1d^k1\hat{L}_{1j}=\sum_{k_{1}}2(\bm{M}_{1})_{jk_{1}}\hat{d}_{k_{1}} and L^2j=k22(𝑴2)jk2d^k2\hat{L}_{2j}=\sum_{k_{2}}2(\bm{M}_{2})_{jk_{2}}\hat{d}_{k_{2}}. The BCF with this ansatz becomes

𝑪>A(Δt)=(𝑽1i𝑴1)e𝒁1Δt(𝑽1+i𝑴1),𝑪<A(Δt)=(𝑽2i𝑴2)e𝒁2Δt(𝑽2+i𝑴2)T,\begin{gathered}\bm{C}^{>A}(\Delta t)=(\bm{V}_{1}-i\bm{M}_{1})e^{-\bm{Z}_{1}\Delta t}(\bm{V}_{1}+i\bm{M}_{1})^{\dagger},\\ \bm{C}^{<A}(\Delta t)=(\bm{V}_{2}-i\bm{M}_{2})^{*}e^{-\bm{Z}_{2}\Delta t}(\bm{V}_{2}+i\bm{M}_{2})^{T},\end{gathered} (17)

where 𝒁1\bm{Z}_{1} and 𝒁2\bm{Z}_{2} are diagonal matrices with elements zk1=γk1+iEk1z_{k_{1}}=\gamma_{k_{1}}+iE_{k_{1}} and zk2=γk2iEk2z_{k_{2}}=\gamma_{k_{2}}-iE_{k_{2}}. We note that the baths A1A_{1} and A2A_{2} only contribute to the BCF 𝑪>A\bm{C}^{>A} and 𝑪<A\bm{C}^{<A}, respectively. We remark that the BCF from the fermionic bath is also expressed as a complex weighted sum of complex exponential functions and has the same gauge redundancy as the bosonic bath.

It is instructive to consider a model with a single coupling term (with only j=1j=1). In this case, the BCF simplifies to a scalar function without indices:

CA(Δt)=kwkezkΔt,C^{A}(\Delta t)=\sum_{k}w_{k}e^{-z_{k}\Delta t}, (18)

where wk=(VkiMk)(VkiMk)w_{k}=(V_{k}-iM_{k})(V_{k}^{*}-iM_{k}^{*}). This expression is a sum of complex exponential functions with complex weights. In contrast, the conventional pseudomode description corresponds to Mk=0M_{k}=0, resulting in positive weights wk=|Vk|2w_{k}=|V_{k}|^{2}. Physically, this exactly represents the BCF from a positive weighted sum of the Lorentzian spectrum, and thus, we also refer to the conventional pseudomode as a Lorentzian pseudomode. We see, however, that the quasi-Lindblad pseudomode provides for a more flexible and general representation, whose influence on the compactness of the pseudomode description we will examine in the numerics below.

In addition to the enhanced flexibility, the use of complex exponential expressions also simplifies the numerical fitting of the BCF. In this work, we utilize the ESPRIT algorithm [37, 38], which has previously been reported as one of the most competitive methods for BCF fitting [44], to determine the complex exponents, zkz_{k}. We then perform a least-squares fitting of the complex weights wkw_{k}. We note that this simple unconstrained least-squares approach is possible because we do not restrict wk0w_{k}\geq 0 as would be required in the conventional pseudomode approach. Once wkw_{k} is determined, we choose the gauge for VkV_{k} and MkM_{k}. In this limit, the gauge choice is parameterized as VkiMk=κkwk,Vk+iMk=κk1wkV_{k}-iM_{k}=\kappa_{k}\sqrt{w_{k}},V_{k}+iM_{k}=\kappa_{k}^{*-1}\sqrt{w^{*}_{k}} where κk\kappa_{k}\in\mathbb{C}. Here, we follow a simple heuristic of minimizing the magnitude of MkM_{k}, as this term determines the extent of CP condition violation. We describe the result for the bosonic bath expression in Eq. 13 for 𝑽\bm{V} and 𝑴\bm{M}, and its adaptation to the fermionic BCF expression is straightforward. The solution of VkV_{k} and MkM_{k} with minimum |Mk||M_{k}| is then given by VkiMk=wkV_{k}-iM_{k}=\sqrt{w_{k}} with Vk,MkV_{k},M_{k}\in\mathbb{R}, (κk=1\kappa_{k}=1) which we provide its proof in the SM [52]. We will examine other choices of VkV_{k} and MkM_{k} and their effect on the stability of the system-bath dynamics in Sec. IV.

There are several approaches to relax the CPTP condition in pseudomode representations. One approach involves a non-Hermitian pseudomode formulation [32, 33, 34, 35], where the system-bath coupling is given by a non-Hermitian Hamiltonian (without Lindblad coupling). This formulation results in real weights, wkw_{k}\in\mathbb{R} [32, 35], or complex weights from a pair of pseudomodes [35]. Another approach, presented in [41], is a Lindblad-like pseudomode formulation derived from the HEOM with a similarity transformation. While this approach allows for complex weights, it is important to emphasize that the Lindblad-like equation in [41] represents one gauge choice for κk\kappa_{k} 333The corresponding κk\kappa_{k} for the Lindblad-like equation in [41] is κk=Re[wk]/wk\kappa_{k}=\sqrt{\operatorname{Re}[w_{k}]/w_{k}}. within our quasi-Lindblad pseudomode framework.

Refer to caption
Figure 2: Exponential function fitting of BCF for (a,d) subohmic, (b,e) semicircular, (c,f) and two-site impurity spectral densities. The real part of the fitted BCF is shown in (a,b,c) from the complex-weighted fit to complex exponentials (red solid, quasi-Lindblad pseudomode formulation in this work), positive-weighted fit to complex exponentials (blue dash-dotted, conventional Lorentzian pseudomode formulation), and positive-weighted fit to real exponentials (green dotted, discrete Hamiltonian) compared to the exact BCF (black dashed). The inset shows the zoomed-in data at t>1t>1 (only the results from complex exponential fits). The number of exponential functions, NexpN_{\exp}, used for the fitting was Nexp=4N_{\exp}=4 and 66 for (a) and (b,c), respectively. (d,e,f) Fitting errors, fit\mathcal{E}_{\text{fit}}, as a function of NexpN_{\exp} from the fitting with complex weights (red dots), positive weights (blue stars), and discrete Hamiltonian (green crosses).

For general couplings with multiple jj, a similar strategy can be adopted. In this case, we first find a sum of exponential functions with a matrix-valued weight 𝑪A(Δt)k𝑾kezkΔt\bm{C}^{A}(\Delta t)\approx\sum_{k}\bm{W}_{k}e^{-z_{k}\Delta t}. We determine zkz_{k} by fitting a scalar quantity, such as the sum of the BCF entries, jjCjjA(Δt)k(jj(𝑾k)jj)ezkΔt\sum_{jj^{\prime}}C^{A}_{jj^{\prime}}(\Delta t)\approx\sum_{k}\left(\sum_{jj^{\prime}}(\bm{W}_{k})_{jj^{\prime}}\right)e^{-z_{k}\Delta t}, or the trace jCjjA(Δt)k(j(𝑾k)jj)ezkΔt\sum_{j}C^{A}_{jj}(\Delta t)\approx\sum_{k}\left(\sum_{j}(\bm{W}_{k})_{jj}\right)e^{-z_{k}\Delta t}. The matrix-valued complex weights 𝑾k\bm{W}_{k} are then obtained through the least-squares fitting.

Finding 𝑽\bm{V} and 𝑴\bm{M} from 𝑾k\bm{W}_{k} involves additional matrix decompositions. However, the representation from Eq. 13 has a rank-1 matrix factorization form, (𝑾k)jj=(𝑽i𝑴)jk(𝑽+i𝑴)jk(\bm{W}_{k})_{jj^{\prime}}=(\bm{V}-i\bm{M})_{jk}(\bm{V}+i\bm{M})^{*}_{j^{\prime}k}, and 𝑾k\bm{W}_{k} from the least-squares fitting is generally not a rank-1 matrix, but of rank NSN_{S}, after denoting the range of jj as NSN_{S}. To address this, we represent the rank-NSN_{S} 𝑾k\bm{W}_{k} by indexing the pseudomode with n=(k,s)n=(k,s) where the index ss is an additional index with a range of NSN_{S}. Then, we assign the diagonal elements of 𝒁\bm{Z} as zn=z(k,s)=zkz_{n}=z_{(k,s)}=z_{k}. This gives the following expression:

CjjA(Δt)=n=(k,s)(𝑽i𝑴)jn(𝑽+i𝑴)jnezkΔt,C^{A}_{jj^{\prime}}(\Delta t)=\sum_{n=(k,s)}(\bm{V}-i\bm{M})_{jn}(\bm{V}+i\bm{M})^{*}_{j^{\prime}n}e^{-z_{k}\Delta t}, (19)

where (𝑾k)jj=s(𝑽i𝑴)jn(𝑽+i𝑴)jn(\bm{W}_{k})_{jj^{\prime}}=\sum_{s}(\bm{V}-i\bm{M})_{jn}(\bm{V}+i\bm{M})^{*}_{j^{\prime}n}. Thus, given NexpN_{\exp} exponents zkz_{k} and NSN_{S} coupling operators, NSNexpN_{S}N_{\exp} pseudomodes are used in the BCF fitting. With this setting, we decompose 𝑾k\bm{W}_{k} with a singular value decomposition (SVD), 𝑾k=𝑼k𝚺k𝑼~k\bm{W}_{k}=\bm{U}_{k}\bm{\Sigma}_{k}\widetilde{\bm{U}}_{k}^{\dagger}, where 𝚺k\bm{\Sigma}_{k} is a diagonal matrix. One possible choice for 𝑽\bm{V} and 𝑴\bm{M} is (𝑽i𝑴)jn=(𝑼k𝚺k1/2)js,(𝑽+i𝑴)jn=(𝑼~k𝚺k1/2)js(\bm{V}-i\bm{M})_{jn}=(\bm{U}_{k}\bm{\Sigma}^{1/2}_{k})_{js},(\bm{V}+i\bm{M})_{jn}=(\widetilde{\bm{U}}_{k}{\bm{\Sigma}_{k}}^{1/2})_{js}, which reduces to the gauge choice above for the single coupling limit 444The SVD representation has a phase redundancy in 𝑼k\bm{U}_{k}, 𝑼k𝑼kei𝚯k\bm{U}_{k}\rightarrow\bm{U}_{k}e^{i\bm{\Theta}_{k}}, and also in 𝑴k\bm{M}_{k}, 𝑴k𝑴kei𝚯k\bm{M}_{k}\rightarrow\bm{M}_{k}e^{i\bm{\Theta}_{k}}, where 𝚯k\bm{\Theta}_{k} is a real diagonal matrix. However, it does not affect the overall norm of 𝑴k\bm{M}_{k}..

III Numerical results

III.1 Fitting the bath correlation function

The quasi-Lindblad pseudomode formulation provides the BCF as a complex weighted sum of complex exponential functions. This type of BCF representation is widely used, for example, in HEOM [40, 42, 41, 43, 44] and tensor network influence functionals [45, 46]. It is known in many applications to show a fast exponential convergence rate with the number of pseudomodes. In this section, we focus on comparing the performance of BCF fitting using the quasi-Lindblad ansatz to that achieved with the conventional Lorentzian pseudomode ansatz and a discrete Hamiltonian ansatz obtained from direct discretization.

Given a spectral density Jjj(ω)=gj(ω)gj(ω)J_{jj^{\prime}}(\omega)=g_{j}(\omega)g_{j^{\prime}}^{*}(\omega), we will fit the associated zero-temperature BCF [2],

Cjj(t)=Jjj(ω)eiωt𝑑ω,C_{jj^{\prime}}(t)=\int J_{jj^{\prime}}(\omega)e^{-i\omega t}d\omega, (20)

using the three different ansatz. We examine the performance of these ansatz for three spectral densities, as illustrated in Fig. 2. The first case is a subohmic spectral density,

J(ω)=α2ωc1sωseω/ωc,ω[0,),J(\omega)=\frac{\alpha}{2}\omega_{c}^{1-s}\omega^{s}e^{-\omega/\omega_{c}},\quad\omega\in[0,\infty), (21)

the second, a semicircular spectral density,

J(ω)=Γπ1ω2W2,ω[0,W],J(\omega)=\frac{\Gamma}{\pi}\sqrt{1-\frac{\omega^{2}}{W^{2}}},\quad\omega\in[0,W], (22)

and the third corresponds to a two-site spectral density,

Jjj(ω)=(1r(ω)r(ω)1)J(ω),J_{jj^{\prime}}(\omega)=\begin{pmatrix}1&r(\omega)\\ r^{*}(\omega)&1\end{pmatrix}J(\omega), (23)

where r(ω)=exp(iω/2W)r(\omega)=\exp(-i\omega/2W) and J(ω)J(\omega) is the semicircular spectral density specified above. For the subohmic case, we use s=0.5s=0.5 and α=ωc=1\alpha=\omega_{c}=1, and for the semicircular case, we set Γ=1\Gamma=1 and W=10ΓW=10\Gamma.

Figs. 2 (a–c) compare the fitted BCF to the exact expressions shown by the black dashed lines. Three different fits are compared: (1) complex exponential function fits with complex weights (red solid), (2) with positive (or PSD for the multi-site case) weights (blue dash-dotted), and (3) a discrete Hamiltonian (green dotted). Fitting scheme (2) represents the Lorentzian pseudomode. Here, the parameters are optimized through numerical minimization of the fitting error. For fitting (3), we used a discretization scheme based on Gaussian quadrature using Legendre polynomials [14]. Details of these fitting schemes are in the SM [52]. The fits to the quasi-Lindblad ansatz with Nexp=4N_{\exp}=4 (Nexp=6N_{\exp}=6) for Fig. 2a (Fig. 2b and Fig. 2c), respectively, show high accuracy already with a small number of exponentials. In Fig. 2d, Fig. 2e, and 2f, the maximum fitting error as a function of NexpN_{\text{exp}} is illustrated. This comparison shows the clear advantage of the quasi-Lindblad pseudomode compared to the Lorentzian pseudomode and direct discretization schemes and illustrates an exponential decrease of error with respect to NexpN_{\text{exp}} in both cases.

Refer to caption
Figure 3: Time evolution of coherence σ^x\langle\hat{\sigma}_{x}\rangle from the quench dynamics of the spin-boson model. The analytical result from the continuous bath with s=0.5s=0.5 subohmic spectral density is compared to results from a two-mode discretized bath using quasi-Lindblad pseudomodes, Lorentzian pseudomodes, and a discrete Hamiltonian. (inset) Errors of coherence compared to the analytical result.

III.2 Quench dynamics simulation

We then test the accuracy of the quasi-Lindblad pseudomode formulation in simulating the quench dynamics of a spin-boson model and a fermionic impurity model. First, we demonstrate the effectiveness of this approach in the spin-boson model. We choose the Hamiltonian H^S=ω0σ^z/2\hat{H}_{S}=\omega_{0}\hat{\sigma}_{z}/2 with ω0=4\omega_{0}=4 and S^j=1=σ^z\hat{S}_{j=1}=\hat{\sigma}_{z} and the initial system-reduced density operator ρ^S(0)=|++|\hat{\rho}_{S}(0)=\mathinner{|{+}\rangle}\!\!\mathinner{\langle{+}|}, with |+=1/2(|0+|1)\mathinner{|{+}\rangle}=1/\sqrt{2}(\mathinner{|{0}\rangle}+\mathinner{|{1}\rangle}). While the analytical solution for this setup is known [2, 4], obtaining accurate numerical results remains nontrivial [56, 26, 57], particularly due to the requirement for a precise description of the BCF.

We choose the s=0.5s=0.5 subohmic spectral density as above (α=ωc=1\alpha=\omega_{c}=1) and consider a zero temperature bath initial condition. The BCF is fitted using two bosonic modes, and we compare the quasi-Lindblad pseudomode, Lorentzian pseudomode, and a discrete Hamiltonian from Gaussian quadrature. Fig. 3 shows the time evolution of the coherence σ^x(t)\langle\hat{\sigma}_{x}(t)\rangle using these three different bath discretizations. We see that the time evolution from the discrete Hamiltonian shows unphysical oscillations in the long-time limit, while the Lorentzian pseudomode qualitatively captures the dissipative dynamics but at a much lower accuracy than the quasi-Lindblad pseudomode representation.

Refer to caption
Figure 4: Quench dynamics of spinless (a) single-site and (b) two-site impurity models with H^S=εjn^j\hat{H}_{S}=\varepsilon\sum_{j}\hat{n}_{j} with ε=4Γ\varepsilon=-4\Gamma, (c) spinful SIAM and (d) spinless two-site impurity model with H^S=Un^0n^1+εjn^j\hat{H}_{S}=U\hat{n}_{0}\hat{n}_{1}+\varepsilon\sum_{j}\hat{n}_{j} with U=2ε=8ΓU=-2\varepsilon=8\Gamma. (a, b) Time-dependence of impurity occupations, n^j(t)\langle\hat{n}_{j}(t)\rangle, computed from Gaussian calculations from our quasi-Lindblad pseudomode, Lorentzian pseudomode, discrete Hamiltonian, and exact unitary dynamics with Nexp=4N_{\text{exp}}=4. (inset) A maximum error of n^j(t)\langle\hat{n}_{j}(t)\rangle from pseudomode dynamics as a function of NexpN_{\text{exp}} for three different fitting schemes in Fig. 2. (c, d) n^j(t)\langle\hat{n}_{j}(t)\rangle of our pseudomode compared to the original unitary dynamics with large bath discretization, computed from tdDMRG.

We next study the dynamics of the spinless noninteracting single-site and two-site models with H^S=εjn^j\hat{H}_{S}=\varepsilon\sum_{j}\hat{n}_{j} (n^j=a^ja^j\hat{n}_{j}=\hat{a}_{j}^{\dagger}\hat{a}_{j}). Here, we can carry out efficient Gaussian calculations, which we do for both a continuous unitary bath and for the Lindblad pseudomode dynamics [22, 58]. We choose the semicircular spectral density in Eq. 22 and Eq. 23 with W=10ΓW=10\Gamma, but with ω[W,W]\omega\in[-W,W], ε=4Γ\varepsilon=-4\Gamma, and consider the initial bath as a fully occupied (empty) state for ω<0\omega<0 (ω>0\omega>0). We fit the two BCFs, Cjj>(t)C^{>}_{jj^{\prime}}(t) and Cjj<(t)C^{<}_{jj^{\prime}}(t), with NexpN_{\text{exp}} exponential functions each and represent the bath with in total 2Nexp2N_{\text{exp}} (4Nexp4N_{\text{exp}}) pseudomodes for the single-site (two-site) impurity model, respectively. The initial impurity state is assumed to be an empty state, ρ^S(0)=|00|\hat{\rho}_{S}(0)=\mathinner{|{0}\rangle}\!\!\mathinner{\langle{0}|}, for the single-site model, and ρ^S(0)=|1010|\hat{\rho}_{S}(0)=\mathinner{|{10}\rangle}\!\!\mathinner{\langle{10}|} for the two-site model. Fig. 4a and 4b show the time-dependence of the impurity occupation, n^j(t)\langle\hat{n}_{j}(t)\rangle, with Nexp=4N_{\text{exp}}=4 for the exact dynamics, the quasi-Lindblad pseudomode, Lorentzian pseudomode, and discrete Hamiltonian dynamics. These results clearly show the improved accuracy of the quasi-Lindblad pseudomode relative to the Lorentzian pseudomode, which reaches the wrong steady state. The inset in Fig. 4a and 4b shows the maximum error of the impurity occupation as a function of NexpN_{\text{exp}}, which also shows an exponential convergence in NexpN_{\text{exp}}.

Refer to caption
Figure 5: Stability behavior of the spin-boson model for different gauge choices (for the same BCF) in the quasi-Lindblad formulation (a,d) κ=1\kappa=1, (b,e) κ=2\kappa=2, and (c,f) κ=0.5\kappa=0.5, (details in the main text). (a,b,c) Time evolution of the excited state occupation, P1(t)=1|ρ^S(t)|1P_{1}(t)=\mathinner{\langle{1}|}\hat{\rho}_{S}(t)\mathinner{|{1}\rangle}. (d,e,f) Real and imaginary parts of eigenvalues λ\lambda of Liouvillian =+𝒟\mathcal{L}=\mathcal{H}+\mathcal{D} at Nmax=40N_{\max}=40. The red vertical line marks the point where the real part of the eigenvalue is zero. A positive real part in the eigenvalues implies instability.

Next, we carry out a test with interacting models - a spinful single-site impurity model, known as the single-impurity Anderson model (SIAM), with H^S=Un^n^+ε(n^+n^)\hat{H}_{S}=U\hat{n}_{\uparrow}\hat{n}_{\downarrow}+\varepsilon(\hat{n}_{\uparrow}+\hat{n}_{\downarrow}), and a spinless two-site impurity model with H^S=Un^0n^1+ε(n^0+n^1)\hat{H}_{S}=U\hat{n}_{0}\hat{n}_{1}+\varepsilon(\hat{n}_{0}+\hat{n}_{1}) - and choose U=2ε=8ΓU=-2\varepsilon=8\Gamma. For the SIAM model, we assume that there is the same bath for each spin with a semicircular spectral density without any coupling between the baths. The initial bath state is assumed to be the same as in the noninteracting case. As a reference for the exact unitary dynamics, we use a large bath discretization obtained from Gaussian quadrature with Legendre polynomials [14] - 80 (200) bath modes for the single-site (two-site) impurity model, respectively. For both the unitary reference and pseudomode Lindblad dynamics, we use the time-dependent density matrix renormalization group (tdDMRG) to propagate the dynamics [59, 60] with bond dimensions of up to 250, using the Block2 [61, 62] package. Fig. 4c and 4d show the time-dependence of n^j(t)\langle\hat{n}_{j}(t)\rangle (j=j=\uparrow for the SIAM, j=0,1j=0,1 for the two-site impurity). The quasi-Lindblad ansatz was constructed from Nexp=4N_{\text{exp}}=4 (2Nexp=82N_{\text{exp}}=8 pseudomodes per each spin) for SIAM and Nexp=8N_{\text{exp}}=8 (4Nexp=324N_{\text{exp}}=32 pseudomodes) for the two-site impurity model. We see that the quasi-Lindblad pseudomode accurately reproduces the original unitary dynamics, including the long-lived oscillation in the two-site impurity model.

IV Stability of Quasi-Lindblad dynamics

We now discuss the stability of quasi-Lindblad global dynamics. As mentioned above, the quasi-Lindblad master equation relaxes the CP condition on the Lindblad dissipator forms, and this implies that the dissipator for the quasi-Lindbladian, 𝒟=𝒟A+𝒟SA\mathcal{D}=\mathcal{D}_{A}+\mathcal{D}_{SA}, always has an eigenvalue with a positive real part. However, we observe that adding the Hamiltonian part of the Liouvillian, =i[H^aux,]\mathcal{H}\bullet=-i[\hat{H}_{\text{aux}},\bullet], can induce a stable dynamics (i.e., the real parts of all eigenvalues of =+𝒟\mathcal{L}=\mathcal{H}+\mathcal{D} are nonpositive). We call this Hamiltonian-induced stability. For some intuition, we can consider the Trotter decomposition of the time evolution operator et=limN(et/Ne𝒟t/N)Ne^{\mathcal{L}t}=\lim_{N\rightarrow\infty}\left(e^{\mathcal{H}t/N}e^{\mathcal{D}t/N}\right)^{N}. Although the time evolution e𝒟t/Ne^{\mathcal{D}t/N} contains both dissipative and divergent components, the Hamiltonian part et/Ne^{\mathcal{H}t/N} can mix these different components so that the overall dynamics may become stable.

As a concrete first example, consider the following quasi-Lindblad dynamics for a single spin

tρ^=i[H^,ρ^]+γxσ^x(ρ)γzσ^z(ρ^).\partial_{t}\hat{\rho}=-i[\hat{H},\hat{\rho}]+\gamma_{x}\mathcal{L}_{\hat{\sigma}_{x}}(\rho)-\gamma_{z}\mathcal{L}_{\hat{\sigma}_{z}}(\hat{\rho}). (24)

Here σ^α=σ^ασ^α\mathcal{L}_{\hat{\sigma}_{\alpha}}\bullet=\hat{\sigma}_{\alpha}\bullet\hat{\sigma}_{\alpha}-\bullet where σ^α\hat{\sigma}_{\alpha} is a Pauli matrix. Without the Hamiltonian part, the dynamics is clearly unstable for any γx,γz>0\gamma_{x},\gamma_{z}>0. However, if we choose H^=νσ^y\hat{H}=\nu\hat{\sigma}_{y} and solve Eq. (24) using the Bloch sphere representation ρ^=12(I^+𝒂𝝈^)\hat{\rho}=\frac{1}{2}(\hat{I}+{\bm{a}}\cdot\bm{\hat{\sigma}}), we can write down a set of closed equations for the coefficients 𝒂(t){\bm{a}}(t) as

t𝒂(t)=2(γz0ν0γx+γz0ν0γx)𝒂(t).\partial_{t}{\bm{a}}(t)=2\begin{pmatrix}\gamma_{z}&0&\nu\\ 0&-\gamma_{x}+\gamma_{z}&0\\ -\nu&0&-\gamma_{x}\end{pmatrix}{\bm{a}}(t). (25)

Therefore, when γx>γz>0,|ν|2>γxγz\gamma_{x}>\gamma_{z}>0,|\nu|^{2}>\gamma_{x}\gamma_{z}, all eigenvalues have nonpositive real parts, and Eq. (24) has a unique fixed point which is the maximally mixed state. Note that the Hamiltonian-induced stability is contingent upon both the form of the Hamiltonian (in this case, choosing H^σ^y\hat{H}\propto\hat{\sigma}_{y}) and the parameter choices. This dependency also persists in the more complex examples discussed below. A more general formulation of the Hamiltonian-induced stability may be captured by the hypocoercivity theory, which was originally formulated in the context of classical kinetic theory [63] and has recently been applied in the context of open quantum systems described by Lindblad equations [64].

Next, we go beyond single spin dynamics and illustrate the form of Hamiltonian-induced stability arising in a noninteracting fermionic impurity model, assuming a system-only Hamiltonian H^S=jj(𝑯S)jja^ja^j\hat{H}_{S}=\sum_{jj^{\prime}}(\bm{H}_{S})_{jj^{\prime}}\hat{a}_{j}^{\dagger}\hat{a}_{j^{\prime}}. The noninteracting nature of this fermionic model allows us to describe its dynamics using a Gaussian fermionic formalism, where we formulate the equation of motion in terms of a one-particle reduced density matrix (1-RDM) Ppq(t)=Tr[c^qc^pρ^SA(t)]P_{pq}(t)=\operatorname{Tr}[\hat{c}_{q}^{\dagger}\hat{c}_{p}\hat{\rho}_{SA}(t)]. The equation of motion for Ppq(t)P_{pq}(t) is given by the continuous-time differential Lyapunov equation [22, 58]:

t𝑷=𝑿𝑷+𝑷𝑿+𝒀,\partial_{t}\bm{P}=\bm{X}\bm{P}+\bm{P}\bm{X}^{\dagger}+\bm{Y}, (26)

where

𝑿=(i𝑯Si𝑽1𝑴1i𝑽2𝑴2i𝑽1𝑴1𝒁1𝟎i𝑽2𝑴2𝟎𝒁2),𝒀=(𝟎𝟎2𝑴2𝟎𝟎𝟎2𝑴2𝟎2𝚪2).\begin{gathered}\bm{X}=\left(\begin{array}[]{ccc}-i\boldsymbol{H}_{S}&-i\bm{V}_{1}-\bm{M}_{1}&-i\bm{V}_{2}-\bm{M}_{2}\\ -i\bm{V}_{1}^{\dagger}-\bm{M}_{1}^{\dagger}&-\bm{Z}_{1}&\bm{0}\\ -i\bm{V}_{2}^{\dagger}-\bm{M}_{2}^{\dagger}&\bm{0}&-\bm{Z}_{2}\end{array}\right),\\ \bm{Y}=\left(\begin{array}[]{ccc}\bm{0}&\bm{0}&2\bm{M}_{2}\\ \bm{0}&\bm{0}&\bm{0}\\ 2\bm{M}_{2}^{\dagger}&\bm{0}&2\bm{\Gamma}_{2}\end{array}\right).\end{gathered} (27)

The Lyapunov equation in Eq. (26) is asymptotically stable if and only if the real parts of all eigenvalues of 𝑿\bm{X} are strictly negative [65]. For simplicity, we now assume a single-site system, i.e., H^S=ha^a^\hat{H}_{S}=h\hat{a}^{\dagger}\hat{a} where hh\in\mathbb{R}, and 𝑿\bm{X} is

𝑿=(ihi𝑽𝑴i𝑽𝑴𝒁),\bm{X}=\left(\begin{array}[]{cc}-ih&-i\bm{V}-\bm{M}\\ -i\bm{V}^{\dagger}-\bm{M}^{\dagger}&-\bm{Z}\end{array}\right), (28)

where we do not separate baths 1 and 2 since there is no difference in the expression for 𝑿\bm{X}. In this setting, and assuming without loss of generality, transformation to a diagonal 𝐙\mathbf{Z} with elements zk=γk+iEkz_{k}=\gamma_{k}+iE_{k} (see discussion around Eq. (13)) we can prove that when 𝑴<12minγk\|\bm{M}\|<\frac{1}{2}\min\gamma_{k}, if the magnitude of the Hamiltonian term 𝑽\|\bm{V}\| is sufficiently large, then the dynamics in Eq. 26 with 𝑿\bm{X} being Eq. 28 is asymptotically stable (see the proof in the SM [52]).

The noninteracting fermionic impurity model has a special feature that the eigenvalue spectrum of 𝑿\bm{X}, which determines the asymptotic stability of dynamics, is dependent on the weight wk=(VkiMk)(Vk+iMk)w_{k}=(V_{k}-iM_{k})(V_{k}+iM_{k})^{*}, but not on the gauge choice for each VkV_{k} and MkM_{k}. In contrast, in non-integrable systems, such as the ones studied earlier in this work, we observe different stability behavior depending on the gauge degrees of freedom. Take a single-mode spin-boson problem, for instance. In HEOM simulations, this problem has been observed to have instabilities [47] after a bosonic Hilbert space truncation. The problem with the reported instability is defined by H^S=Δσ^x/2\hat{H}_{S}=\Delta\hat{\sigma}_{x}/2, S^j=1=σ^z\hat{S}_{j=1}=\hat{\sigma}_{z}, and the BCF C(t)=weztC(t)=we^{-zt}, with Δ=4.0\Delta=4.0, w=50.02.5iw=50.0-2.5i, and z=1.0z=1.0, and the initial reduced density operator is ρ^S(0)=|11|\hat{\rho}_{S}(0)=\mathinner{|{1}\rangle}\!\!\mathinner{\langle{1}|}. The Hamiltonian and quasi-Lindblad parameters, V,MV,M\in\mathbb{C}, satisfy the constraint (ViM)(V+iM)=w(V-iM)(V+iM)^{*}=w. We vary VV and MM through the parameterization ViM=κw,V+iM=κ1wV-iM=\kappa\sqrt{w},V+iM=\kappa^{-1}\sqrt{w^{*}}. Fig. 5 shows the time evolution of the excited state occupation, P1(t)=1|ρ^S(t)|1P_{1}(t)=\mathinner{\langle{1}|}\hat{\rho}_{S}(t)\mathinner{|{1}\rangle} at κ=1\kappa=1, κ=2\kappa=2, and κ=0.5\kappa=0.5. We recall that κ=1\kappa=1 is the parameter that minimizes |M||M|. We impose a Hilbert space truncation by limiting the maximum number of bosonic excitations, nNmaxn\leq N_{\max}. At κ=1\kappa=1, the results converge with increasing NmaxN_{\max} without any propagation instabilities, whereas at κ=2\kappa=2 and κ=0.5\kappa=0.5, the results show instabilities in the long time limit, similar to the HEOM results in [47].

Figs. 5 (d–f) show the eigenvalues λ\lambda of the Liouvillian =+𝒟\mathcal{L}=\mathcal{H}+\mathcal{D} with Nmax=40N_{\max}=40. The maximum real part of the eigenvalues is 0 at κ=1\kappa=1, but it is positive at κ=2\kappa=2 and κ=0.5\kappa=0.5, which shows the origin of the differing stability behavior. In Fig. 6, the maximum real part of the eigenvalues is plotted between κ=1\kappa=1 and κ=2\kappa=2 (Fig. 6a) and between 1/κ=11/\kappa=1 and 1/κ=21/\kappa=2 (Fig. 6b). This clearly shows a transition from stable to unstable dynamics at κ>1.3\kappa>1.3 and 1/κ>1.21/\kappa>1.2.

Refer to caption
Figure 6: The maximum real parts of eigenvalues of \mathcal{L} as a function of κ\kappa with H^S=2σ^x\hat{H}_{S}=2\hat{\sigma}_{x} (black dots) and H^S=0\hat{H}_{S}=0 (red stars). κ=1\kappa=1 is the recommended choice in this work and is found to be numerically stable.

This result shows that our choice κ=1\kappa=1 that minimizes |M||M| provides a stable solution for this BCF. However, we are not yet aware of a parameterization that can guarantee the stability of the dynamics for an arbitrary BCF, and the verification becomes more challenging in general multi-mode cases.

A practical recipe is to guess κ\kappa from a related problem where the stability can be easily determined. Here, we propose a feasible procedure for a multi-mode boson model. In particular, the spin-boson model has an analytical solution when the system Hamiltonian H^S\hat{H}_{S} commutes with the coupling operator S^\hat{S}. For simplicity, we take H^S=0\hat{H}_{S}=0, with a single spin-site, and the limit of a single-coupling. Then, the system-reduced density operator in the coupling operator eigenbasis, ρ^S=|ss|\hat{\rho}_{S}=\mathinner{|{s}\rangle}\!\!\mathinner{\langle{s^{\prime}}|} and S^|s=s|s\hat{S}\mathinner{|{s}\rangle}=s\mathinner{|{s}\rangle}, the density operator ρ^=ρ^Sρ^B\hat{\rho}=\hat{\rho}_{S}\otimes\hat{\rho}_{B} forms an invariant subspace of the Liouvillian \mathcal{L},

(ρ^Sρ^B)=ρ^Ss,s(ρ^B),\mathcal{L}(\hat{\rho}_{S}\otimes\hat{\rho}_{B})=\hat{\rho}_{S}\otimes\mathcal{L}_{s,s^{\prime}}(\hat{\rho}_{B}), (29)

where s,s\mathcal{L}_{s,s^{\prime}} is an effective Liouvillian within the auxiliary bath depending on ρ^S\hat{\rho}_{S} (ss and ss^{\prime}). s,s\mathcal{L}_{s,s^{\prime}} is explicitly expressed as,

s,s\displaystyle\mathcal{L}_{s,s^{\prime}}\bullet =𝒟AisA^+isA^\displaystyle=\mathcal{D}_{A}\bullet-is\hat{A}\bullet+is^{\prime}\bullet\hat{A}
+k((2ss)Mkd^ksMkd^k)\displaystyle+\sum_{k}\left((2s^{\prime}-s)M_{k}\hat{d}_{k}-sM_{k}^{*}\hat{d}_{k}^{\dagger}\right)\bullet
+k(sMkd^k+(2ss)Mkd^k),\displaystyle+\sum_{k}\bullet\left(-s^{\prime}M_{k}\hat{d}_{k}+(2s-s^{\prime})M_{k}^{*}\hat{d}_{k}^{\dagger}\right), (30)

with A^=kVkd^k+Vkd^k\hat{A}=\sum_{k}V_{k}\hat{d}_{k}+V_{k}^{*}\hat{d}_{k}^{\dagger}. We note that if 𝒟A\mathcal{D}_{A} is given in the diagonal basis (diagonal 𝒁\bm{Z} in Eq. 13), the Liouvillian becomes a sum of independent Liouvillians, s,s=ks,s,k\mathcal{L}_{s,s^{\prime}}=\sum_{k}\mathcal{L}_{s,s^{\prime},k}. In this case, we can determine the stability of multi-mode spin-boson problems from single-mode spin-boson calculations with s,s,k\mathcal{L}_{s,s^{\prime},k}.

Although the above calculation is only applicable in the analytically solvable case, we can imagine that the stability is still related to that under general H^S\hat{H}_{S} dynamics. To assess this correlation, we compute the maximum real part of the eigenvalues of \mathcal{L} as a function of κ\kappa with H^S=0\hat{H}_{S}=0, and with the other parameters set to be the same as in the previous setup in Fig. 6. This calculation is done by computing the eigenvalues of s,s\mathcal{L}_{s,s^{\prime}} (single kk in this case) with s,s=±1s,s^{\prime}=\pm 1 and taking the maximum value over them. We also observe that all the eigenvalues with positive real parts are from s=1,s=1s=1,s^{\prime}=-1, implying that the origin of instability is from the nondiagonal elements of the system density operator. For κ<1\kappa<1, the computed eigenvalues match remarkably well between the analytically tractable H^S=0\hat{H}_{S}=0 dynamics and the case of H^S=2σ^x\hat{H}_{S}=2\hat{\sigma}_{x} considered previously. For κ>1\kappa>1, the eigenvalues overestimate the region of stability in κ\kappa but nonetheless correctly predicts that a region around κ=1\kappa=1 is stable.

V Conclusions

We have presented the quasi-Lindblad pseudomode theory, which provides a flexible framework to incorporate a highly compact representation of the bath correlation function (BCF) of an open quantum system. Specifically, the BCF in the quasi-Lindblad pseudomode theory takes the form of a sum of complex exponential functions with complex weights, in contrast to positive weights in the conventional Lindblad pseudomode theory. By adopting an efficient numerical algorithm for the complex exponential function fitting procedure, we showed that the number of pseudomodes can be significantly reduced, and only a small number of exponential functions (<10<10) is sufficient for both fitting the BCF and describing the associated reduced system dynamics.

The quasi-Lindblad pseudomode theory achieves this representation of the BCF by relaxing the complete positivity (CP) of the global dynamics consisting of the system and the pseudomodes. Although the total system-bath density operator along the dynamics remains trace-preserving throughout the dynamics, it is generally not positive. Nonetheless, after tracing out the bath, the quasi-Lindblad pseudomode theory can still produce an accurate system-reduced density operator, which can be very close to a positive operator, provided the error in the BCF relative to the original unitary problem is sufficiently small. The dependence of the error of the system-reduced density operator on the BCF has been rigorously analyzed in the context of unitary dynamics for spin-boson systems [66, 45, 67], and work is in progress towards extending such results for fermionic systems and for the Lindblad and quasi-Lindblad theories [68].

The pseudomode formulation provides an alternative framework to other numerical approaches to open quantum system dynamics, which have also used a complex-weighted sum of exponential functions to represent the BCF, such as the HEOM [40, 42, 41]. An important strength is that once the pseudomodes are determined, there exists a wide variety of numerical methods that can be used to propagate the resulting master equation [56, 22, 23, 21, 30, 31, 69, 70, 71].

A crucial theoretical question raised by the quasi-Lindblad formulation is the stability of the global dynamics. Despite relaxing the CP condition, we found that the combination of the Hamiltonian and dissipator terms can lead to dynamics which is stable, a phenomenon we refer to as Hamiltonian-induced stability. Further, because the quasi-Lindblad formulation contains a set of gauge degrees of freedom, we can take advantage of them to create a stable global dynamics for a particular system, as we demonstrated in the case of the single-mode spin-boson model. In this example, adjusting the gauge parameters allowed us to resolve an instability issue that is known to occur when using the HEOM method [47]. Nonetheless, although such instabilities are a formal possibility within our theory and should be studied further, the widespread use of HEOM in open quantum system simulations also suggests that they may not be relevant to a range of physical studies.

We anticipate that this theoretical development can facilitate accurate simulations of open quantum systems within more complex environments, such as those encountered in photosynthetic systems [72], nonequilibrium Kondo problems [73], and quantum control [74].

Acknowledgements.– This work is an equal collaboration between two SciDAC teams “Real-time dynamics of driven correlated electrons in quantum materials” and “Traversing the ‘death valley’ separating short and long times in non-equilibrium quantum dynamical simulations of real materials”, supported by the U.S. Department of Energy, Office of Science, Office of Advanced Scientific Computing Research and Office of Basic Energy Sciences, Scientific Discovery through Advanced Computing (SciDAC) program under Award Number DE–SC0022088 (G.P., C.Y., G.K.C.) and DE–SC0022198 (L.L., C.Y., Y.Z.). The work by Z.H. was supported by the Simons Targeted Grants in Mathematics and Physical Sciences on Moiré Materials Magic. G.K.C. and L.L. are Simons Investigators. We thank Zhiyan Ding, Joonho Lee, David Limmer, Vojtěch Vlček, Erika Ye, Lexing Ying, and Neill Lambert for helpful discussions.

Author contributions.– G.P., Z.H., G.K.C., L.L. conceived the original study. G.P., Z.H., Y.Z., and L.L. carried out theoretical analysis to support the study. G.P. and Z.H. carried out numerical calculations to support the study. All authors, G.P., Z.H., Y.Z., C.Y., G.K.C., and L.L., discussed the results of the manuscript and contributed to the writing of the manuscript.

Note added.– During submission of this manuscript, a related paper [75] appeared, which discusses analytical numerical scaling and provides different numerical fitting schemes.

References

  • Rivas and Huelga [2012] A. Rivas and S. F. Huelga, Open quantum systems (Springer, 2012).
  • Breuer and Petruccione [2007] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, 2007).
  • Berkelbach and Thoss [2020] T. C. Berkelbach and M. Thoss, Special topic on dynamics of open quantum systems, The Journal of Chemical Physics 152, 020401 (2020).
  • Leggett et al. [1987] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, Dynamics of the dissipative two-state system, Rev. Mod. Phys. 59, 1 (1987).
  • Feynman and Vernon [2000] R. Feynman and F. Vernon, The theory of a general quantum system interacting with a linear dissipative system, Annals of Physics 281, 547 (2000).
  • Makri and Makarov [1995] N. Makri and D. E. Makarov, Tensor propagator for iterative quantum time evolution of reduced density matrices. i. theory, The Journal of Chemical Physics 102, 4600 (1995).
  • 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, Nature Communications 910.1038/s41467-018-05617-3 (2018).
  • Ye and Chan [2021] E. Ye and G. K.-L. Chan, Constructing tensor network influence functionals for general quantum dynamics, The Journal of Chemical Physics 155, 044104 (2021).
  • Ng et al. [2023] N. Ng, G. Park, A. J. Millis, G. K.-L. Chan, and D. R. Reichman, Real-time evolution of anderson impurity models via tensor network influence functionals, Phys. Rev. B 107, 125103 (2023).
  • Mühlbacher and Rabani [2008] L. Mühlbacher and E. Rabani, Real-time path integral approach to nonequilibrium many-body quantum systems, Phys. Rev. Lett. 100, 176403 (2008).
  • Cohen et al. [2015] G. Cohen, E. Gull, D. R. Reichman, and A. J. Millis, Taming the dynamical sign problem in real-time evolution of quantum many-body problems, Phys. Rev. Lett. 115, 266802 (2015).
  • Núñez Fernández et al. [2022] Y. Núñez Fernández, M. Jeannin, P. T. Dumitrescu, T. Kloss, J. Kaye, O. Parcollet, and X. Waintal, Learning feynman diagrams with tensor trains, Phys. Rev. X 12, 041018 (2022).
  • Ivander et al. [2024] F. Ivander, L. P. Lindoy, and J. Lee, Unified framework for open quantum dynamics with memory, Nature Communications 15, 8087 (2024).
  • de Vega et al. [2015] I. de Vega, U. Schollwöck, and F. A. Wolf, How to discretize a quantum bath for real-time evolution, Phys. Rev. B 92, 155126 (2015).
  • Prior et al. [2010] J. Prior, A. W. Chin, S. F. Huelga, and M. B. Plenio, Efficient simulation of strong system-environment interactions, Phys. Rev. Lett. 105, 050404 (2010).
  • Woods et al. [2015] M. P. Woods, M. Cramer, and M. B. Plenio, Simulating bosonic baths with error bars, Phys. Rev. Lett. 115, 130401 (2015).
  • Nusspickel and Booth [2020] M. Nusspickel and G. H. Booth, Efficient compression of the environment of an open quantum system, Phys. Rev. B 102, 165107 (2020).
  • Garraway [1997] B. M. Garraway, Nonperturbative decay of an atomic system in a cavity, Phys. Rev. A 55, 2290 (1997).
  • Dalton et al. [2001] B. J. Dalton, S. M. Barnett, and B. M. Garraway, Theory of pseudomodes in quantum optical processes, Phys. Rev. A 64, 053813 (2001).
  • Tamascelli et al. [2018] D. Tamascelli, A. Smirne, S. F. Huelga, and M. B. Plenio, Nonperturbative treatment of non-markovian dynamics of open quantum systems, Phys. Rev. Lett. 120, 030402 (2018).
  • Schwarz et al. [2016] F. Schwarz, M. Goldstein, A. Dorda, E. Arrigoni, A. Weichselbaum, and J. von Delft, Lindblad-driven discretized leads for nonequilibrium steady-state transport in quantum impurity models: Recovering the continuum limit, Phys. Rev. B 94, 155142 (2016).
  • Lotem et al. [2020] M. Lotem, A. Weichselbaum, J. von Delft, and M. Goldstein, Renormalized lindblad driving: A numerically exact nonequilibrium quantum impurity solver, Phys. Rev. Res. 2, 043052 (2020).
  • Brenes et al. [2020] M. Brenes, J. J. Mendoza-Arenas, A. Purkayastha, M. T. Mitchison, S. R. Clark, and J. Goold, Tensor-network method to simulate strongly interacting quantum thermal machines, Phys. Rev. X 10, 031040 (2020).
  • Elenewski et al. [2017] J. E. Elenewski, D. Gruss, and M. Zwolak, Communication: Master equations for electron transport: The limits of the Markovian limit, The Journal of Chemical Physics 147, 151101 (2017).
  • Trivedi et al. [2021] R. Trivedi, D. Malz, and J. I. Cirac, Convergence guarantees for discrete mode approximations to non-markovian quantum baths, Phys. Rev. Lett. 127, 250404 (2021).
  • Mascherpa et al. [2020] F. Mascherpa, A. Smirne, A. D. Somoza, P. Fernández-Acebal, S. Donadi, D. Tamascelli, S. F. Huelga, and M. B. Plenio, Optimized auxiliary oscillators for the simulation of general open quantum systems, Phys. Rev. A 101, 052108 (2020).
  • Chen et al. [2019] F. Chen, E. Arrigoni, and M. Galperin, Markovian treatment of non-markovian dynamics of open fermionic systems, New Journal of Physics 21, 123035 (2019).
  • Li [2021] X. Li, Markovian embedding procedures for non-markovian stochastic schrödinger equations, Physics Letters A 387, 127036 (2021).
  • Lednev et al. [2024] M. Lednev, F. J. García-Vidal, and J. Feist, Lindblad master equation capable of describing hybrid quantum systems in the ultrastrong coupling regime, Phys. Rev. Lett. 132, 106902 (2024).
  • Dorda et al. [2014] A. Dorda, M. Nuss, W. von der Linden, and E. Arrigoni, Auxiliary master equation approach to nonequilibrium correlated impurities, Phys. Rev. B 89, 165105 (2014).
  • Dorda et al. [2015] A. Dorda, M. Ganahl, H. G. Evertz, W. von der Linden, and E. Arrigoni, Auxiliary master equation approach within matrix product states: Spectral properties of the nonequilibrium anderson impurity model, Phys. Rev. B 92, 125145 (2015).
  • Lambert et al. [2019] N. Lambert, S. Ahmed, M. Cirio, and F. Nori, Modelling the ultra-strongly coupled spin-boson model with unphysical modes, Nature Communications 10, 3721 (2019).
  • Pleasance et al. [2020] G. Pleasance, B. M. Garraway, and F. Petruccione, Generalized theory of pseudomodes for exact descriptions of non-markovian quantum processes, Phys. Rev. Res. 2, 043058 (2020).
  • Cirio et al. [2023] M. Cirio, N. Lambert, P. Liang, P.-C. Kuo, Y.-N. Chen, P. Menczel, K. Funo, and F. Nori, Pseudofermion method for the exact description of fermionic environments: From single-molecule electronics to the kondo resonance, Phys. Rev. Res. 5, 033011 (2023).
  • Menczel et al. [2024] P. Menczel, K. Funo, M. Cirio, N. Lambert, and F. Nori, Non-hermitian pseudomodes for strongly coupled open quantum systems: Unravelings, correlations, and thermodynamics, Phys. Rev. Res. 6, 033237 (2024).
  • Beylkin and Monzón [2005] G. Beylkin and L. Monzón, On approximation of functions by exponential sums, Applied and Computational Harmonic Analysis 19, 17 (2005).
  • Paulraj et al. [1986] A. Paulraj, R. Roy, and T. Kailath, A subspace rotation approach to signal parameter estimation, Proceedings of the IEEE 74, 1044 (1986).
  • Li et al. [2020] W. Li, W. Liao, and A. Fannjiang, Super-resolution limit of the esprit algorithm, IEEE Transactions on Information Theory 66, 4593 (2020).
  • Nakatsukasa et al. [2018] Y. Nakatsukasa, O. Sète, and L. N. Trefethen, The aaa algorithm for rational approximation, SIAM Journal on Scientific Computing 40, A1494 (2018).
  • Xu et al. [2022] M. Xu, Y. Yan, Q. Shi, J. Ankerhold, and J. T. Stockburger, Taming quantum noise for efficient low temperature simulations of open quantum systems, Phys. Rev. Lett. 129, 230601 (2022).
  • Xu et al. [2023] M. Xu, V. Vadimov, M. Krug, J. T. Stockburger, and J. Ankerhold, A universal framework for quantum dissipation:minimally extended state space and exact time-local dynamics (2023), arXiv:2307.16790 [quant-ph] .
  • Dan et al. [2023] X. Dan, M. Xu, J. T. Stockburger, J. Ankerhold, and Q. Shi, Efficient low-temperature simulations for fermionic reservoirs with the hierarchical equations of motion method: Application to the anderson impurity model, Phys. Rev. B 107, 195429 (2023).
  • Chen et al. [2022] Z.-H. Chen, Y. Wang, X. Zheng, R.-X. Xu, and Y. Yan, Universal time-domain prony fitting decomposition for optimized hierarchical quantum master equations, The Journal of Chemical Physics 156, 221102 (2022).
  • Takahashi et al. [2024] H. Takahashi, S. Rudge, C. Kaspar, M. Thoss, and R. Borrelli, High accuracy exponential decomposition of bath correlation functions for arbitrary and structured spectral densities: Emerging methodologies and new approaches, The Journal of Chemical Physics 160, 204105 (2024).
  • Vilkoviskiy and Abanin [2024] I. Vilkoviskiy and D. A. Abanin, Bound on approximating non-markovian dynamics by tensor networks in the time domain, Phys. Rev. B 109, 205126 (2024).
  • Guo and Chen [2024] C. Guo and R. Chen, Efficient construction of the Feynman-Vernon influence functional as matrix product states, SciPost Phys. Core 7, 063 (2024).
  • Krug and Stockburger [2023] M. Krug and J. Stockburger, On stability issues of the heom method, The European Physical Journal Special Topics 232, 3219 (2023).
  • Dunn et al. [2019] I. S. Dunn, R. Tempelaar, and D. R. Reichman, Removing instabilities in the hierarchical equations of motion: Exact and approximate projection approaches, The Journal of Chemical Physics 150, 184109 (2019).
  • Yan et al. [2020] Y. Yan, T. Xing, and Q. Shi, A new method to improve the numerical stability of the hierarchical equations of motion for discrete harmonic oscillator modes, The Journal of Chemical Physics 153, 204109 (2020).
  • Li et al. [2022] T. Li, Y. Yan, and Q. Shi, A low-temperature quantum Fokker-Planck equation that improves the numerical stability of the hierarchical equations of motion for the Brownian oscillator spectral density, The Journal of Chemical Physics 156, 064107 (2022).
  • Note [1] In the original formulation in [20], the additional conditions on the one-point bath expectation values Tr[B^j(t)ρ^B(0)]\operatorname{Tr}[\hat{B}_{j}(t)\hat{\rho}_{B}(0)] were imposed. In the main text, we further assumed that the initial bath reduced density operators, ρ^B(0)\hat{\rho}_{B}(0) and ρ^A(0)\hat{\rho}_{A}(0), are number-conserving operators. It leads to the one-point bath expectation values becoming zero for both the original unitary and pseudomode formulations.
  • [52] See the Supplemental Material for the proof of the equivalence condition, the BCF expression for the spin-boson and the fermionic impurity model, proofs of the stability condition, and details of fitting procedures and DMRG calculations, which includes Refs. [20, 1, 2, 76, 77, 78, 79, 30, 80, 22, 58, 37, 81, 82, 83, 38, 84, 85, 86, 38, 87, 23, 24, 25, 61, 62, 88, 89, 90].
  • Note [2] In principle, we have a general expression by considering non-diagonalizable 𝒁\bm{Z}. However, numerically, fitting with arbitrary 𝒁\bm{Z} yields the same optimal solution as with diagonalizable 𝒁\bm{Z}.
  • Note [3] The corresponding κk\kappa_{k} for the Lindblad-like equation in [41] is κk=Re[wk]/wk\kappa_{k}=\sqrt{\operatorname{Re}[w_{k}]/w_{k}}.
  • Note [4] The SVD representation has a phase redundancy in 𝑼k\bm{U}_{k}, 𝑼k𝑼kei𝚯k\bm{U}_{k}\rightarrow\bm{U}_{k}e^{i\bm{\Theta}_{k}}, and also in 𝑴k\bm{M}_{k}, 𝑴k𝑴kei𝚯k\bm{M}_{k}\rightarrow\bm{M}_{k}e^{i\bm{\Theta}_{k}}, where 𝚯k\bm{\Theta}_{k} is a real diagonal matrix. However, it does not affect the overall norm of 𝑴k\bm{M}_{k}.
  • Somoza et al. [2019] A. D. Somoza, O. Marty, J. Lim, S. F. Huelga, and M. B. Plenio, Dissipation-assisted matrix product factorization, Phys. Rev. Lett. 123, 100502 (2019).
  • Cygorek et al. [2024] M. Cygorek, J. Keeling, B. W. Lovett, and E. M. Gauger, Sublinear scaling in non-markovian open quantum systems simulations, Phys. Rev. X 14, 011010 (2024).
  • Barthel and Zhang [2022] T. Barthel and Y. Zhang, Solving quasi-free and quadratic lindblad master equations for open fermionic and bosonic systems, Journal of Statistical Mechanics: Theory and Experiment 2022, 113101 (2022).
  • White and Feiguin [2004] S. R. White and A. E. Feiguin, Real-time evolution using the density matrix renormalization group, Phys. Rev. Lett. 93, 076401 (2004).
  • Paeckel et al. [2019] S. Paeckel, T. Köhler, A. Swoboda, S. R. Manmana, U. Schollwöck, and C. Hubig, Time-evolution methods for matrix-product states, Annals of Physics 411, 167998 (2019).
  • Zhai and Chan [2021] H. Zhai and G. K.-L. Chan, Low communication high performance ab initio density matrix renormalization group algorithms, The Journal of Chemical Physics 15410.1063/5.0050902 (2021), 224116.
  • Zhai et al. [2023] H. Zhai, H. R. Larsson, S. Lee, Z.-H. Cui, T. Zhu, C. Sun, L. Peng, R. Peng, K. Liao, J. Tölle, J. Yang, S. Li, and G. K.-L. Chan, Block2: A comprehensive open source framework to develop and apply state-of-the-art DMRG algorithms in electronic structure and beyond, The Journal of Chemical Physics 159, 234801 (2023).
  • Villani [2007] C. Villani, Hypocoercive diffusion operators, Bollettino dellUnione Matematica Italiana 10-B, 257 (2007).
  • Fang et al. [2024] D. Fang, J. Lu, and Y. Tong, Mixing time of open quantum systems via hypocoercivity (2024), arXiv:2404.11503 [quant-ph] .
  • Simoncini [2016] V. Simoncini, Computational methods for linear matrix equations, SIAM Review 58, 377 (2016).
  • Mascherpa et al. [2017] F. Mascherpa, A. Smirne, S. F. Huelga, and M. B. Plenio, Open systems with error bounds: Spin-boson model with spectral density variations, Phys. Rev. Lett. 118, 100401 (2017).
  • Liu and Lu [2024] K. Liu and J. Lu, Error bounds for open quantum systems with harmonic bosonic bath (2024), arXiv:2408.04009 .
  • Huang et al. [2024] Z. Huang, L. Lin, G. Park, and Y. Zhu, Unified analysis of non-markovian open quantum systems in gaussian environment using superoperator formalism (2024), arXiv:2411.08741 [quant-ph] .
  • Werner et al. [2023] D. Werner, J. Lotze, and E. Arrigoni, Configuration interaction based nonequilibrium steady state impurity solver, Phys. Rev. B 107, 075119 (2023).
  • Werner et al. [2024] D. Werner, R. Žitko, and E. Arrigoni, Auxiliary master equation approach to the anderson-holstein impurity problem out of equilibrium, Phys. Rev. B 109, 075156 (2024).
  • Luo et al. [2023] S. Luo, N. Lambert, P. Liang, and M. Cirio, Quantum-classical decomposition of gaussian quantum environments: A stochastic pseudomode model, PRX Quantum 4, 030316 (2023).
  • Ishizaki and Fleming [2009] A. Ishizaki and G. R. Fleming, Theoretical examination of quantum coherence in a photosynthetic system at physiological temperature, Proceedings of the National Academy of Sciences 106, 17255 (2009).
  • Hewson [1993] A. C. Hewson, The Kondo Problem to Heavy Fermions, Cambridge Studies in Magnetism (Cambridge University Press, 1993).
  • Lambert et al. [2023] N. Lambert, T. Raheja, S. Cross, P. Menczel, S. Ahmed, A. Pitchford, D. Burgarth, and F. Nori, Qutip-bofin: A bosonic and fermionic numerical hierarchical-equations-of-motion library with applications in light-harvesting, quantum control, and single-molecule electronics, Phys. Rev. Res. 5, 013181 (2023).
  • Thoenniss et al. [2024] J. Thoenniss, I. Vilkoviskiy, and D. A. Abanin, Efficient pseudomode representation and complexity of quantum impurity models (2024), arXiv:2409.08816 [cond-mat.str-el] .
  • Aurell et al. [2020] E. Aurell, R. Kawai, and K. Goyal, An operator derivation of the feynman-vernon theory, with applications to the generating function of bath energy changes and to an-harmonic baths, Journal of Physics A: Mathematical and Theoretical 53, 275303 (2020).
  • Cirio et al. [2022] M. Cirio, P.-C. Kuo, Y.-N. Chen, F. Nori, and N. Lambert, Canonical derivation of the fermionic influence superoperator, Phys. Rev. B 105, 035121 (2022).
  • Schmutz [1978] M. Schmutz, Real-time green’s functions in many body problems, Zeitschrift für Physik B Condensed Matter 30, 97 (1978).
  • Dzhioev and Kosov [2011] A. A. Dzhioev and D. S. Kosov, Super-fermion representation of quantum kinetic equations for the electron transport problem, The Journal of Chemical Physics 13410.1063/1.3548065 (2011), 044121.
  • Park et al. [2024] G. Park, N. Ng, D. R. Reichman, and G. K.-L. Chan, Tensor network influence functionals in the continuous-time limit: Connections to quantum embedding, bath discretization, and higher-order time propagation, Phys. Rev. B 110, 045104 (2024).
  • Roy and Kailath [1989] R. Roy and T. Kailath, Esprit-estimation of signal parameters via rotational invariance techniques, IEEE Transactions on Acoustics, Speech, and Signal Processing 37, 984 (1989).
  • Rouquette and Najim [2001] S. Rouquette and M. Najim, Estimation of frequencies and damping factors by two-dimensional esprit type methods, IEEE Transactions on Signal Processing 49, 237 (2001).
  • Shahbazpanahi et al. [2001] S. Shahbazpanahi, S. Valaee, and M. Bastani, Distributed source localization using esprit algorithm, IEEE Transactions on Signal Processing 49, 2169 (2001).
  • Ding et al. [2024] Z. Ding, E. N. Epperly, L. Lin, and R. Zhang, The ESPRIT algorithm under high noise: Optimal error scaling and noisy super-resolution, in 2024 Symposium on Foundations of Computer Science (FOCS 2024) (2024) arXiv:2404.03885 .
  • Stroeks et al. [2022] M. E. Stroeks, J. Helsen, and B. M. Terhal, Spectral estimation for hamiltonians: a comparison between classical imaginary-time evolution and quantum real-time evolution, New Journal of Physics 24, 103024 (2022).
  • Li et al. [2023] H. Li, H. Ni, and L. Ying, Adaptive low-depth quantum algorithms for robust multiple-phase estimation, Phys. Rev. A 108, 062408 (2023).
  • Huang et al. [2023] Z. Huang, E. Gull, and L. Lin, Robust analytic continuation of green’s functions via projection, pole estimation, and semidefinite relaxation, Phys. Rev. B 107, 075151 (2023).
  • Saberi et al. [2008] H. Saberi, A. Weichselbaum, and J. von Delft, Matrix-product-state comparison of the numerical renormalization group and the variational formulation of the density-matrix renormalization group, Phys. Rev. B 78, 035124 (2008).
  • Rams and Zwolak [2020] M. M. Rams and M. Zwolak, Breaking the entanglement barrier: Tensor network simulation of quantum transport, Phys. Rev. Lett. 124, 137701 (2020).
  • Wójtowicz et al. [2020] G. Wójtowicz, J. E. Elenewski, M. M. Rams, and M. Zwolak, Open-system tensor networks and kramers’ crossover for quantum transport, Phys. Rev. A 101, 050301 (2020).

Supplemental Material for “Quasi-Lindblad pseudomode theory for open quantum systems”

Gunhee Park (박건희) Zhen Huang Yuanran Zhu Chao Yang Garnet Kin-Lic Chan Lin Lin

SM-I Proof of the equivalence condition for the bosonic Gaussian bath

In this section, we provide the derivation of the equivalence condition for the bosonic Gaussian bath based on the bath correlation function (BCF). Our strategy here is different from that in [20], which is only applicable to the Lindblad equation and relies on the dilation of the Lindblad system into an infinite dimensional unitary system.

We start from the following system-bath Liouvillian superoperator form introduced in the main text,

SA=ij𝒮jj+ij𝒮~j~j.\mathcal{L}_{SA}=-i\sum_{j}\mathcal{S}_{j}\mathcal{F}_{j}+i\sum_{j}\widetilde{\mathcal{S}}_{j}\widetilde{\mathcal{F}}_{j}. (S1)

For example, in the case when there is only Hamiltonian coupling, H^SA=jS^jA^j\hat{H}_{SA}=\sum_{j}\hat{S}_{j}\hat{A}_{j}, j=A^j\mathcal{F}_{j}\bullet=\hat{A}_{j}\bullet and ~j=A^j\widetilde{\mathcal{F}}_{j}\bullet=\bullet\hat{A}_{j}. In the case of the quasi-Lindblad pseudomode with both Hamiltonian and Lindblad couplings with 𝒟SA=jL^jS^j+S^jL^j12{S^jL^j+L^jS^j,}\mathcal{D}_{SA}\bullet=\sum_{j}\hat{L}^{\prime}_{j}\bullet\hat{S}_{j}+\hat{S}_{j}\bullet\hat{L}^{\prime{\dagger}}_{j}-\frac{1}{2}\{\hat{S}_{j}\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j}\hat{S}_{j},\bullet\},

j=A^j+iL^ji2(L^j+L^j),~j=A^jiL^j+i2(L^j+L^j).\begin{gathered}\mathcal{F}_{j}\bullet=\hat{A}_{j}\bullet+\bullet i\hat{L}^{\prime{\dagger}}_{j}-\frac{i}{2}(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j})\bullet,\\ \widetilde{\mathcal{F}}_{j}\bullet=\bullet\hat{A}_{j}-i\hat{L}^{\prime}_{j}\bullet+\bullet\frac{i}{2}(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j}).\end{gathered} (S2)

For brevity, we will not explicitly distinguish between the non-tilde and tilde superoperators, and denote SA=ij𝒮jj\mathcal{L}_{SA}=-i\sum_{j}\mathcal{S}_{j}\mathcal{F}_{j}, treating 𝒮~j=𝒮jmax+j\widetilde{\mathcal{S}}_{j}=\mathcal{S}_{j_{\max}+j} and ~j=jmax+j\widetilde{\mathcal{F}}_{j}=-\mathcal{F}_{j_{\max}+j}. We will return to the explicit tilde superoperator notation when the distinction is necessary.

In the interaction picture, ρ~^SA(t)=e(S+A)tρ^SA(t)\hat{\widetilde{\rho}}_{SA}(t)=e^{-(\mathcal{L}_{S}+\mathcal{L}_{A})t}\hat{\rho}_{SA}(t), 𝒮j(t)=eSt𝒮jeSt\mathcal{S}_{j}(t)=e^{-\mathcal{L}_{S}t}\mathcal{S}_{j}e^{\mathcal{L}_{S}t}, j(t)=eAtjeAt\mathcal{F}_{j}(t)=e^{-\mathcal{L}_{A}t}\mathcal{F}_{j}e^{\mathcal{L}_{A}t}, and SA(t)=ij𝒮j(t)j(t)\mathcal{L}_{SA}(t)=-i\sum_{j}\mathcal{S}_{j}(t)\mathcal{F}_{j}(t), the von Neumann equation of motion for ρ~^SA(t)\hat{\widetilde{\rho}}_{SA}(t) is,

tρ~^SA(t)=SA(t)ρ~^SA(t).\partial_{t}\hat{\widetilde{\rho}}_{SA}(t)=\mathcal{L}_{SA}(t)\hat{\widetilde{\rho}}_{SA}(t). (S3)

Note that the backward evolution operators such as eAte^{-\mathcal{L}_{A}t} are introduced mainly to simplify the notation, and these operators do not appear explicitly after applying the time-ordering operations below. The formal solution of ρ~^SA(t)\hat{\widetilde{\rho}}_{SA}(t) can be obtained with the Dyson series expansion [1],

ρ~^SA(t)=m=01m!𝒯(0t0tSA(t1)SA(tm)𝑑t1𝑑tm)ρ^SA(0),\hat{\widetilde{\rho}}_{SA}(t)=\sum_{m=0}^{\infty}\frac{1}{m!}\mathcal{T}\left(\int_{0}^{t}\cdots\int_{0}^{t}\mathcal{L}_{SA}(t_{1})\cdots\mathcal{L}_{SA}(t_{m})dt_{1}\cdots dt_{m}\right)\hat{\rho}_{SA}(0), (S4)

where 𝒯\mathcal{T} refers to the time-ordering superoperator. We trace out the bath to obtain the system-reduced density operator,

ρ~^S(t)=m=0(i)mm!0t0tj1jmTrA[𝒯A(j1(t1)jm(tm))ρ^A(0)]𝒯S(𝒮j1(t1)𝒮jm(tm))ρ^S(0).\hat{\widetilde{\rho}}_{S}(t)=\sum_{m=0}^{\infty}\frac{(-i)^{m}}{m!}\int_{0}^{t}\cdots\int_{0}^{t}\sum_{j_{1}}\cdots\sum_{j_{m}}\operatorname{Tr}_{A}\left[\mathcal{T}_{A}(\mathcal{F}_{j_{1}}(t_{1})\cdots\mathcal{F}_{j_{m}}(t_{m}))\hat{\rho}_{A}(0)\right]\mathcal{T}_{S}(\mathcal{S}_{j_{1}}(t_{1})\cdots\mathcal{S}_{j_{m}}(t_{m}))\hat{\rho}_{S}(0). (S5)

We separate the time-ordering superoperator into the system time-ordering superoperator 𝒯S\mathcal{T}_{S} and the bath time-ordering superoperator 𝒯A\mathcal{T}_{A} to be applied within each Liouville operator space. We denote Cj1,,jm(t1,,tm)=TrA[𝒯A(j1(t1)jm(tm))ρ^A(0)]C_{j_{1},\cdots,j_{m}}(t_{1},\cdots,t_{m})=\operatorname{Tr}_{A}\left[\mathcal{T}_{A}(\mathcal{F}_{j_{1}}(t_{1})\cdots\mathcal{F}_{j_{m}}(t_{m}))\hat{\rho}_{A}(0)\right]. From Wick’s theorem, if mm is odd, it is zero, and if mm is even (m=2nm=2n),

Cj1,,j2n(t1,,t2n)=σΠ2ni=1nCjσ(2i1),jσ(2i)(tσ(2i1),tσ(2i)).C_{j_{1},\cdots,j_{2n}}(t_{1},\cdots,t_{2n})=\sum_{\sigma\in\Pi_{2n}}\prod_{i=1}^{n}C_{j_{\sigma(2i-1)},j_{\sigma(2i)}}(t_{\sigma(2i-1)},t_{\sigma(2i)}). (S6)

After inserting it into Eq. S5,

ρ~^S(t)=n=0(1)n(2n)!0t0tj1jnσΠ2n𝒯S(i=1nCjσ(2i1),jσ(2i)(tσ(2i1),tσ(2i))𝒮j1(t1)𝒮jn(tn))ρ^S(0)dt1dtn\hat{\widetilde{\rho}}_{S}(t)=\sum_{n=0}^{\infty}\frac{(-1)^{n}}{(2n)!}\int_{0}^{t}\cdots\int_{0}^{t}\sum_{j_{1}}\cdots\sum_{j_{n}}\sum_{\sigma\in\Pi_{2n}}\mathcal{T}_{S}\left(\prod_{i=1}^{n}C_{j_{\sigma(2i-1)},j_{\sigma(2i)}}(t_{\sigma(2i-1)},t_{\sigma(2i)})\mathcal{S}_{j_{1}}(t_{1})\cdots\mathcal{S}_{j_{n}}(t_{n})\right)\hat{\rho}_{S}(0)dt_{1}\cdots dt_{n} (S7)

Thanks to the system time ordering superoperator, 𝒯S\mathcal{T}_{S}, 𝒯S(i=1nCjσ(2i1),jσ(2i)(tσ(2i1),tσ(2i))𝒮j1(t1)𝒮jn(tn))\mathcal{T}_{S}\left(\prod_{i=1}^{n}C_{j_{\sigma(2i-1)},j_{\sigma(2i)}}(t_{\sigma(2i-1)},t_{\sigma(2i)})\mathcal{S}_{j_{1}}(t_{1})\cdots\mathcal{S}_{j_{n}}(t_{n})\right) becomes the same operator for all σΠ2n\sigma\in\Pi_{2n}. The number of elements of Π2n\Pi_{2n} is (2n)!!(2n)!!, so Eq. S7 is reduced to,

ρ~^S(t)\displaystyle\hat{\widetilde{\rho}}_{S}(t) =n=0(1)n2nn!0t0tj1jn𝒯S(i=1nCj2i1,j2i(t2i1,t2i)𝒮j1(t1)𝒮jn(tn))ρ^S(0)dt1dtn\displaystyle=\sum_{n=0}^{\infty}\frac{(-1)^{n}}{2^{n}n!}\int_{0}^{t}\cdots\int_{0}^{t}\sum_{j_{1}}\cdots\sum_{j_{n}}\mathcal{T}_{S}\left(\prod_{i=1}^{n}C_{j_{2i-1},j_{2i}}(t_{2i-1},t_{2i})\mathcal{S}_{j_{1}}(t_{1})\cdots\mathcal{S}_{j_{n}}(t_{n})\right)\hat{\rho}_{S}(0)dt_{1}\cdots dt_{n} (S8)
=𝒯Sexp(120t0tj1,j2Cj1,j2(t1,t2)𝒮j1(t1)𝒮j2(t2)dt1dt2)ρ^S(0)\displaystyle=\mathcal{T}_{S}\exp\left(-\frac{1}{2}\int_{0}^{t}\int_{0}^{t}\sum_{j_{1},j_{2}}C_{j_{1},j_{2}}(t_{1},t_{2})\mathcal{S}_{j_{1}}(t_{1})\mathcal{S}_{j_{2}}(t_{2})dt_{1}dt_{2}\right)\hat{\rho}_{S}(0) (S9)
=𝒯Sexp(0t0t1j1,j2Cj1,j2(t1,t2)𝒮j1(t1)𝒮j2(t2)dt1dt2)ρ^S(0).\displaystyle=\mathcal{T}_{S}\exp\left(-\int_{0}^{t}\int_{0}^{t_{1}}\sum_{j_{1},j_{2}}C_{j_{1},j_{2}}(t_{1},t_{2})\mathcal{S}_{j_{1}}(t_{1})\mathcal{S}_{j_{2}}(t_{2})dt_{1}dt_{2}\right)\hat{\rho}_{S}(0). (S10)

From the second to the third line, we fix the ordering between t1t_{1} and t2t_{2} as t1>t2t_{1}>t_{2}. After returning to the original frame and parameterizing the integrand in the time non-local exponential superoperator,

ρ^S(t)=eSt𝒯Sexp(0t0t1𝒲(t1,t2)𝑑t1𝑑t2)ρ^S(0),\hat{\rho}_{S}(t)=e^{\mathcal{L}_{S}t}\mathcal{T}_{S}\exp\left(\int_{0}^{t}\int_{0}^{t_{1}}\mathcal{W}(t_{1},t_{2})dt_{1}dt_{2}\right)\hat{\rho}_{S}(0), (S11)

which is termed the influence superoperator [2, 76, 77]. The expression for 𝒲(t1,t2)\mathcal{W}(t_{1},t_{2}) after reviving the tilde superoperators becomes,

𝒲(t1,t2)=\displaystyle\mathcal{W}(t_{1},t_{2})= jj(j(t1)j(t2)A𝒮j(t1)𝒮j(t2)j(t1)~j(t2)A𝒮j(t1)𝒮~j(t2)\displaystyle-\sum_{jj^{\prime}}\Big{(}\langle\mathcal{F}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A}\mathcal{S}_{j}(t_{1})\mathcal{S}_{j^{\prime}}(t_{2})-\langle\mathcal{F}_{j}(t_{1})\widetilde{\mathcal{F}}_{j^{\prime}}(t_{2})\rangle_{A}\mathcal{S}_{j}(t_{1})\widetilde{\mathcal{S}}_{j^{\prime}}(t_{2})
~j(t1)j(t2)A𝒮~j(t1)𝒮j(t2)+~j(t1)~j(t2)A𝒮~j(t1)𝒮~j(t2)),\displaystyle-\langle\widetilde{\mathcal{F}}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A}\widetilde{\mathcal{S}}_{j}(t_{1})\mathcal{S}_{j^{\prime}}(t_{2})+\langle\widetilde{\mathcal{F}}_{j}(t_{1})\widetilde{\mathcal{F}}_{j^{\prime}}(t_{2})\rangle_{A}\widetilde{\mathcal{S}}_{j}(t_{1})\widetilde{\mathcal{S}}_{j^{\prime}}(t_{2})\Big{)}, (S12)

where A=TrA[ρ^A(0)]\langle\bullet\rangle_{A}=\operatorname{Tr}_{A}[\bullet\hat{\rho}_{A}(0)]. This expression can be further simplified by using the following property, TrA[j]=TrA[~j]\operatorname{Tr}_{A}[\mathcal{F}_{j}\bullet]=\operatorname{Tr}_{A}[\widetilde{\mathcal{F}}_{j}\bullet]. For the Hamiltonian coupling part, it is clear to see TrA[A^j]=TrA[A^j]\operatorname{Tr}_{A}[\hat{A}_{j}\bullet]=\operatorname{Tr}_{A}[\bullet\hat{A}_{j}]. For the Lindblad coupling part in Eq. S2, it can be seen from direct calculation,

TrA[iL^ji2(L^j+L^j)]=TrA[i2(L^jL^j)]=TrA[iL^j+i2(L^j+L^j)].\operatorname{Tr}_{A}[i\bullet\hat{L}^{\prime{\dagger}}_{j}-\frac{i}{2}(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j})\bullet]=\operatorname{Tr}_{A}[\frac{i}{2}(\hat{L}^{\prime{\dagger}}_{j}-\hat{L}^{\prime}_{j})\bullet]=\operatorname{Tr}_{A}[-i\hat{L}^{\prime}_{j}\bullet+\frac{i}{2}\bullet(\hat{L}^{\prime}_{j}+\hat{L}^{\prime{\dagger}}_{j})]. (S13)

Combining this property with the trace-preserving property, TrA[A]=0\operatorname{Tr}_{A}[\mathcal{L}_{A}\bullet]=0 or TrA[eAt]=TrA[]\operatorname{Tr}_{A}[e^{-\mathcal{L}_{A}t}\bullet]=\operatorname{Tr}_{A}[\bullet], we have,

j(t1)j(t2)A=~j(t1)j(t2)A,j(t1)~j(t2)A=~j(t1)~j(t2)A.\langle\mathcal{F}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A}=\langle\widetilde{\mathcal{F}}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A},\quad\langle\mathcal{F}_{j}(t_{1})\widetilde{\mathcal{F}}_{j^{\prime}}(t_{2})\rangle_{A}=\langle\widetilde{\mathcal{F}}_{j}(t_{1})\widetilde{\mathcal{F}}_{j^{\prime}}(t_{2})\rangle_{A}. (S14)

Furthermore, by using the fact that (j)=~j()(\mathcal{F}_{j}\bullet)^{\dagger}=\widetilde{\mathcal{F}}_{j}(\bullet)^{\dagger}, we have,

~j(t1)~j(t2)A=j(t1)j(t2)A.\langle\widetilde{\mathcal{F}}_{j}(t_{1})\widetilde{\mathcal{F}}_{j^{\prime}}(t_{2})\rangle_{A}=\langle\mathcal{F}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A}^{*}. (S15)

Therefore, by defining the BCF as,

Cjj(t1,t2)=j(t1)j(t2)A=TrA[jeA(t1t2)jeA(t2)ρ^A(0)],C_{jj^{\prime}}(t_{1},t_{2})=\langle\mathcal{F}_{j}(t_{1})\mathcal{F}_{j^{\prime}}(t_{2})\rangle_{A}=\operatorname{Tr}_{A}[\mathcal{F}_{j}e^{\mathcal{L}_{A}(t_{1}-t_{2})}\mathcal{F}_{j^{\prime}}e^{\mathcal{L}_{A}(t_{2})}\hat{\rho}_{A}(0)], (S16)

the superoperator 𝒲(t1,t2)\mathcal{W}(t_{1},t_{2}) is expressed as,

𝒲(t1,t2)=jj(𝒮j(t1)𝒮~j(t1))×(Cjj(t1,t2)𝒮j(t2)Cjj(t1,t2)𝒮~j(t2)).\begin{gathered}\mathcal{W}(t_{1},t_{2})=-\sum_{jj^{\prime}}\left(\mathcal{S}_{j}(t_{1})-\widetilde{\mathcal{S}}_{j}(t_{1})\right)\times\\ \left(C_{jj^{\prime}}(t_{1},t_{2})\mathcal{S}_{j^{\prime}}(t_{2})-C^{*}_{jj^{\prime}}(t_{1},t_{2})\widetilde{\mathcal{S}}_{j^{\prime}}(t_{2})\right).\end{gathered} (S17)

We see, therefore, that the influence of the bath on the system dynamics, as expressed through the influence superoperator, is entirely parametrized by the BCF. Thus, any two baths with the same BCF give rise to the same system dynamics, which is the equivalence condition.

SM-II BCF expression for the spin-boson model

In this section, we derive the BCF expression for the spin-boson model,

𝑪A(Δt)=(𝑽i𝑴)e𝒁Δt(𝑽+i𝑴).\bm{C}^{A}(\Delta t)=(\bm{V}-i\bm{M})e^{-\bm{Z}\Delta t}(\bm{V}+i\bm{M})^{\dagger}. (S18)

This can be shown from a direct calculation. First, eAtρ^A(0)=eAt|𝟎𝟎|=|𝟎𝟎|e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)=e^{\mathcal{L}_{A}t^{\prime}}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|} from A|𝟎𝟎|=0\mathcal{L}_{A}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}=0. We have,

j|𝟎𝟎|=k(VjkiMjk)d^k|𝟎𝟎|,\mathcal{F}_{j^{\prime}}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}=\sum_{k}(V_{j^{\prime}k}^{*}-iM_{j^{\prime}k}^{*})\hat{d}_{k}^{\dagger}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}, (S19)

from d^k|𝟎𝟎|=|𝟎𝟎|d^k=0\hat{d}_{k}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}\hat{d}_{k}^{\dagger}=0. Then, from A[d^k|𝟎𝟎|]=l(iHlkAΓlk)d^l|𝟎𝟎|\mathcal{L}_{A}[\hat{d}_{k}^{\dagger}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}]=\sum_{l}(-iH^{A}_{lk}-\Gamma_{lk})\hat{d}_{l}^{\dagger}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}, we get eAt[k(VjkiMjk)d^k|𝟎𝟎|]=lk(e𝒁t)lk(VjkiMjk)d^l|𝟎𝟎|e^{\mathcal{L}_{A}t}\left[\sum_{k}(V_{j^{\prime}k}^{*}-iM_{j^{\prime}k}^{*})\hat{d}_{k}^{\dagger}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}\right]=\sum_{lk}(e^{-\bm{Z}t})_{lk}(V_{j^{\prime}k}^{*}-iM_{j^{\prime}k}^{*})\hat{d}_{l}^{\dagger}\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|} where 𝒁=𝚪+i𝑯A\bm{Z}=\bm{\Gamma}+i\bm{H}^{A}. Using the property in Eq. S13,

TrA[j]=TrA[(k(VjkiMjk)d^k+(Vjk+iMjk)d^k)].\operatorname{Tr}_{A}\left[\mathcal{F}_{j}\bullet\right]=\operatorname{Tr}_{A}\left[(\sum_{k}(V_{jk}-iM_{jk})\hat{d}_{k}+(V^{*}_{jk}+iM^{*}_{jk})\hat{d}^{\dagger}_{k})\bullet\right]. (S20)

Therefore, this leads to,

CjjA(t+t,t)=TrA[jeAtjeAtρ^A(0)]=kl(VjliMjl)(e𝒁t)lk(VjkiMjk),C_{jj^{\prime}}^{A}(t+t^{\prime},t^{\prime})=\operatorname{Tr}_{A}\left[\mathcal{F}_{j}e^{\mathcal{L}_{A}t}\mathcal{F}_{j^{\prime}}e^{\mathcal{L}_{A}t^{\prime}}\hat{\rho}_{A}(0)\right]=\sum_{kl}(V_{jl}-iM_{jl})(e^{-\bm{Z}t})_{lk}(V_{j^{\prime}k}^{*}-iM_{j^{\prime}k}^{*}), (S21)

or in the matrix representation,

𝑪A(t)=(𝑽i𝑴)e𝒁t(𝑽+i𝑴).\bm{C}^{A}(t)=(\bm{V}-i\bm{M})e^{-\bm{Z}t}(\bm{V}+i\bm{M})^{\dagger}. (S22)

SM-III BCF expression for the fermionic impurity model

In this section, we discuss the quasi-Lindblad pseudomode formulation for the fermionic impurity model. The pseudomode model contains two different auxiliary baths, A1A_{1} and A2A_{2} with the initial state, ρ^SA(0)=ρ^S(0)ρ^A(0)\hat{\rho}_{SA}(0)=\hat{\rho}_{S}(0)\otimes\hat{\rho}_{A}(0) with ρ^A(0)=ρ^A1(0)ρ^A2(0)=|𝟎𝟎||𝟏𝟏|\hat{\rho}_{A}(0)=\hat{\rho}_{A_{1}}(0)\otimes\hat{\rho}_{A_{2}}(0)=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}\otimes\mathinner{|{\bm{1}}\rangle}\!\!\mathinner{\langle{\bm{1}}|}. We define index sets, 𝕀1\mathbb{I}_{1} and 𝕀2\mathbb{I}_{2}, to denote the auxiliary mode index k1𝕀1k_{1}\in\mathbb{I}_{1} (k2𝕀2k_{2}\in\mathbb{I}_{2}) for the fermions in the auxiliary bath A1A_{1} (A2A_{2}), respectively. The density operator evolves under the following Lindblad equation,

tρ^SA=i[H^aux,ρ^SA]+(𝒟A1+𝒟A2+𝒟SA1+𝒟SA2)ρ^SA.\partial_{t}\hat{\rho}_{SA}=-i[\hat{H}_{\text{aux}},\hat{\rho}_{SA}]+(\mathcal{D}_{A_{1}}+\mathcal{D}_{A_{2}}+\mathcal{D}_{SA_{1}}+\mathcal{D}_{SA_{2}})\hat{\rho}_{SA}. (S23)

The auxiliary Hamiltonian is given by,

H^aux=H^S+k𝕀1𝕀2Ekd^kd^k+ja^jA^j+A^ja^j,A^j=k𝕀1(𝑽1)jk1d^k1+k𝕀2(𝑽2)jk2d^k2.\hat{H}_{\text{aux}}=\hat{H}_{S}+\sum_{k\in\mathbb{I}_{1}\cup\mathbb{I}_{2}}E_{k}\hat{d}^{\dagger}_{k}\hat{d}_{k}+\sum_{j}\hat{a}^{\dagger}_{j}\hat{A}_{j}+\hat{A}_{j}^{\dagger}\hat{a}_{j},\quad\hat{A}_{j}=\sum_{k\in\mathbb{I}_{1}}(\bm{V}_{1})_{jk_{1}}\hat{d}_{k_{1}}+\sum_{k\in\mathbb{I}_{2}}(\bm{V}_{2})_{jk_{2}}\hat{d}_{k_{2}}. (S24)

The dissipators are given by,

𝒟A1=k1𝕀12γk1(d^k1d^k112{d^k1d^k1,}),𝒟A2=k2𝕀22γk2(d^k2d^k212{d^k2d^k2,}),𝒟SA1=ja^jL^1j+L^1ja^j12{L^1ja^j+a^jL^1j,},𝒟SA2=ja^jL^2j+L^2ja^j12{L^2ja^j+a^jL^2j,},\begin{gathered}\mathcal{D}_{A_{1}}\bullet=\sum_{k_{1}\in\mathbb{I}_{1}}2\gamma_{k_{1}}\left(\hat{d}_{k_{1}}\bullet\hat{d}_{k_{1}}^{\dagger}-\frac{1}{2}\{\hat{d}_{k_{1}}^{\dagger}\hat{d}_{k_{1}},\bullet\}\right),\\ \mathcal{D}_{A_{2}}\bullet=\sum_{k_{2}\in\mathbb{I}_{2}}2\gamma_{k_{2}}\left(\hat{d}^{\dagger}_{k_{2}}\bullet\hat{d}_{k_{2}}-\frac{1}{2}\{\hat{d}_{k_{2}}\hat{d}_{k_{2}}^{\dagger},\bullet\}\right),\\ \mathcal{D}_{SA_{1}}\bullet=\sum_{j}\hat{a}_{j}\bullet\hat{L}_{1j}^{\dagger}+\hat{L}_{1j}\bullet\hat{a}_{j}^{\dagger}-\frac{1}{2}\{\hat{L}^{\dagger}_{1j}\hat{a}_{j}+\hat{a}_{j}^{\dagger}\hat{L}_{1j},\bullet\},\\ \mathcal{D}_{SA_{2}}\bullet=\sum_{j}\hat{a}^{\dagger}_{j}\bullet\hat{L}_{2j}+\hat{L}^{\dagger}_{2j}\bullet\hat{a}_{j}-\frac{1}{2}\{\hat{L}_{2j}\hat{a}^{\dagger}_{j}+\hat{a}_{j}\hat{L}^{\dagger}_{2j},\bullet\},\end{gathered} (S25)

with L^1j=2k1𝕀1(𝑴1)jk1d^k1\hat{L}_{1j}=2\sum_{k_{1}\in\mathbb{I}_{1}}(\bm{M}_{1})_{jk_{1}}\hat{d}_{k_{1}} and L^2j=2k2𝕀2(𝑴2)jk2d^k2\hat{L}_{2j}=2\sum_{k_{2}\in\mathbb{I}_{2}}(\bm{M}_{2})_{jk_{2}}\hat{d}_{k_{2}}.

We adopt a super-fermion representation [78, 79, 30, 80] of Lindblad dynamics to formulate the superoperator formalism for the fermionic system. This representation maps a density operator to a state vector, ρ^|ρ\hat{\rho}\mapsto\mathinner{|{\rho}\rangle\!\rangle}, in the so-called ‘super-Fock’ space with twice the number of original orbitals. This is done with a left-vacuum vector,

|I=k(d^k+d~^k)|𝟎|𝟎,\mathinner{|{I}\rangle\!\rangle}=\prod_{k}(\hat{d}_{k}^{{\dagger}}+\hat{\tilde{d}}_{k}^{{\dagger}})\mathinner{|{\bm{0}}\rangle}\otimes\mathinner{|{\bm{0}}\rangle}, (S26)

where d~k\tilde{d}_{k}^{\dagger} is a fermionic creation operator in the ‘tilde’ space. The density operator is then represented as |ρ=(ρ^𝕀)|I\mathinner{|{\rho}\rangle\!\rangle}=(\hat{\rho}\otimes\mathbb{I})\mathinner{|{I}\rangle\!\rangle}. For example, the initial bath density operator, ρ^A(0)=|𝟎𝟎||𝟏𝟏|\hat{\rho}_{A}(0)=\mathinner{|{\bm{0}}\rangle}\!\!\mathinner{\langle{\bm{0}}|}\otimes\mathinner{|{\bm{1}}\rangle}\!\!\mathinner{\langle{\bm{1}}|} becomes,

ρ^A(0)|ρA(0)=|01𝕀1|10𝕀2.\hat{\rho}_{A}(0)\mapsto\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\mathinner{|{01}\rangle\!\rangle}^{\otimes\mathbb{I}_{1}}\otimes\mathinner{|{10}\rangle\!\rangle}^{\otimes\mathbb{I}_{2}}. (S27)

The trace operation can also be expressed with the left-vacuum vector,

Tr[O^ρ^]=I|O^𝕀|ρ.\operatorname{Tr}[\hat{O}\hat{\rho}]=\mathinner{\langle\!\langle{I}|}\hat{O}\otimes\mathbb{I}\mathinner{|{\rho}\rangle\!\rangle}. (S28)

The following conjugation rules for the left-vacuum vector will become useful,

d^k|I=d~^k|I,d^k=d~^k|I.\hat{d}_{k}\mathinner{|{I}\rangle\!\rangle}=\hat{\tilde{d}}_{k}\mathinner{|{I}\rangle\!\rangle},\quad\hat{d}_{k}^{\dagger}=-\hat{\tilde{d}}_{k}^{\dagger}\mathinner{|{I}\rangle\!\rangle}. (S29)

We will assume that the initial density operator is an even-parity density operator for simplicity. Any superoperator \mathcal{L} can be cast to the operator in the super-Fock space, ^\hat{\mathcal{L}}, or ρ^^|ρ\mathcal{L}\hat{\rho}\mapsto\hat{\mathcal{L}}\mathinner{|{\rho}\rangle\!\rangle}. For example, the following dissipator maps to,

𝒟=2γk(d^kd^k12{d^kd^k,})𝒟^=2γkd~^kd^kγkd^kd^k+γkd~^kd~^kγk.\begin{gathered}\mathcal{D}\bullet=2\gamma_{k}\left(\hat{d}_{k}\bullet\hat{d}_{k}^{\dagger}-\frac{1}{2}\{\hat{d}_{k}^{\dagger}\hat{d}_{k},\bullet\}\right)\mapsto\hat{\mathcal{D}}=2\gamma_{k}\hat{\tilde{d}}_{k}^{\dagger}\hat{d}_{k}-\gamma_{k}\hat{d}_{k}^{\dagger}\hat{d}_{k}+\gamma_{k}\hat{\tilde{d}}_{k}^{\dagger}\hat{\tilde{d}}_{k}-\gamma_{k}.\end{gathered} (S30)

This relation can be shown from the direct calculation of (𝒟ρ^)|I=𝒟^|ρ(\mathcal{D}\hat{\rho})\mathinner{|{I}\rangle\!\rangle}=\hat{\mathcal{D}}\mathinner{|{\rho}\rangle\!\rangle} with the conjugation rule in Eq. S29. The bath Liouvillian superoperator is written as,

^A\displaystyle\hat{\mathcal{L}}_{A} =k1𝕀1(iEk1γk1)d^k1d^k1+(iEk1+γk1)d~^k1d~^k1+2γk1d~^k1d^k1+iEk1γk1\displaystyle=\sum_{k_{1}\in\mathbb{I}_{1}}(-iE_{k_{1}}-\gamma_{k_{1}})\hat{d}^{\dagger}_{k_{1}}\hat{d}_{k_{1}}+(-iE_{k_{1}}+\gamma_{k_{1}})\hat{\tilde{d}}^{\dagger}_{k_{1}}\hat{\tilde{d}}_{k_{1}}+2\gamma_{k_{1}}\hat{\tilde{d}}_{k_{1}}^{\dagger}\hat{d}_{k_{1}}+iE_{k_{1}}-\gamma_{k_{1}}
+k2𝕀2(iEk2+γk2)d^k2d^k2+(iEk2γk2)d~^k2d~^k2+2γk2d^k2d~^k2+iEk2γk2.\displaystyle+\sum_{k_{2}\in\mathbb{I}_{2}}(-iE_{k_{2}}+\gamma_{k_{2}})\hat{d}^{\dagger}_{k_{2}}\hat{d}_{k_{2}}+(-iE_{k_{2}}-\gamma_{k_{2}})\hat{\tilde{d}}^{\dagger}_{k_{2}}\hat{\tilde{d}}_{k_{2}}+2\gamma_{k_{2}}\hat{{d}}_{k_{2}}^{\dagger}\hat{\tilde{d}}_{k_{2}}+iE_{k_{2}}-\gamma_{k_{2}}. (S31)

Note that ^A|ρA(0)=0\hat{\mathcal{L}}_{A}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=0. With this representation, we decompose the coupling superoperators in the following form,

^SA=ij(a^j^j+^j+a^j+a~^j~^j+~^j+a~^j).\hat{\mathcal{L}}_{SA}=-i\sum_{j}\left(\hat{a}_{j}^{\dagger}\hat{\mathcal{F}}^{-}_{j}+\hat{\mathcal{F}}^{+}_{j}\hat{a}_{j}+\hat{\tilde{a}}_{j}^{\dagger}\hat{\widetilde{\mathcal{F}}}^{-}_{j}+\hat{\widetilde{\mathcal{F}}}^{+}_{j}\hat{\tilde{a}}_{j}\right). (S32)

where

^j=A^ji2L^1j+i(L~^2j+12L^2j),^j+=A^j+i(L~^1j12L^1j)+i2L^2j,~^j=A~^j+i(L^1j+12L~^1j)i2L~^2j,~^j+=A~^j+i2L~^1j+i(L^2j12L~^2j),\begin{gathered}\hat{\mathcal{F}}^{-}_{j}=\hat{A}_{j}-\frac{i}{2}\hat{L}_{1j}+i(\hat{\widetilde{L}}_{2j}+\frac{1}{2}\hat{L}_{2j}),\\ \hat{\mathcal{F}}^{+}_{j}=\hat{A}^{\dagger}_{j}+i(\hat{\widetilde{L}}^{\dagger}_{1j}-\frac{1}{2}\hat{L}_{1j}^{\dagger})+\frac{i}{2}\hat{L}_{2j}^{\dagger},\\ \hat{\widetilde{\mathcal{F}}}^{-}_{j}=\hat{\widetilde{A}}_{j}+i(\hat{L}_{1j}+\frac{1}{2}\hat{\widetilde{L}}_{1j})-\frac{i}{2}\hat{\widetilde{L}}_{2j},\\ \hat{\widetilde{\mathcal{F}}}^{+}_{j}=\hat{\widetilde{A}}^{\dagger}_{j}+\frac{i}{2}\hat{\widetilde{L}}^{\dagger}_{1j}+i(\hat{L}_{2j}^{\dagger}-\frac{1}{2}\hat{\widetilde{L}}_{2j}^{\dagger}),\end{gathered} (S33)

with A~^j=k𝕀1𝕀2Vjkd~^k\hat{\widetilde{A}}_{j}=\sum_{k\in\mathbb{I}_{1}\cup\mathbb{I}_{2}}V_{jk}\hat{\tilde{d}}_{k}, L~^1j=2k1𝕀1(𝑴1)jk1d~^k1\hat{\tilde{L}}_{1j}=2\sum_{k_{1}\in\mathbb{I}_{1}}(\bm{M}_{1})_{jk_{1}}\hat{\tilde{d}}_{k_{1}} and L^2j=2k2𝕀2(𝑴2)jk2d~^k2\hat{L}_{2j}=2\sum_{k_{2}\in\mathbb{I}_{2}}(\bm{M}_{2})_{jk_{2}}\hat{\tilde{d}}_{k_{2}}. Note that the ^\hat{\mathcal{F}} operators with superscripts - (++) are composed of a linear combination of d^k\hat{d}_{k} and d~^k\hat{\tilde{d}}_{k} (d^k\hat{d}_{k}^{\dagger} and d~^k\hat{\tilde{d}}_{k}^{\dagger}), respectively. The associated superoperator BCFs are defined by,

Cjj>A(t1,t2)\displaystyle C^{>A}_{jj^{\prime}}(t_{1},t_{2}) =I|^je^A(t1t2)^j+e^At2|ρA(0),\displaystyle=\mathinner{\langle\!\langle{I}|}\hat{\mathcal{F}}^{-}_{j}e^{\hat{\mathcal{L}}_{A}(t_{1}-t_{2})}\hat{\mathcal{F}}^{+}_{j^{\prime}}e^{\hat{\mathcal{L}}_{A}t_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}, (S34)
Cjj<A(t1,t2)\displaystyle C^{<A}_{jj^{\prime}}(t_{1},t_{2}) =I|^j+e^A(t1t2)^je^At2|ρA(0).\displaystyle=\mathinner{\langle\!\langle{I}|}\hat{\mathcal{F}}^{+}_{j}e^{\hat{\mathcal{L}}_{A}(t_{1}-t_{2})}\hat{\mathcal{F}}^{-}_{j^{\prime}}e^{\hat{\mathcal{L}}_{A}t_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}. (S35)

Explicit calculation leads to the BCF expression in terms of a sum of complex exponential functions,

𝑪>A(t+t,t)=(𝑽1i𝑴1)e𝒁1t(𝑽1+i𝑴1),𝑪<A(t+t,t)=(𝑽2i𝑴2)e𝒁2t(𝑽2+i𝑴2)T,\begin{gathered}\bm{C}^{>A}(t+t^{\prime},t^{\prime})=(\bm{V}_{1}-i\bm{M}_{1})e^{-\bm{Z}_{1}t}(\bm{V}_{1}+i\bm{M}_{1})^{\dagger},\\ \bm{C}^{<A}(t+t^{\prime},t^{\prime})=(\bm{V}_{2}-i\bm{M}_{2})^{*}e^{-\bm{Z}_{2}t}(\bm{V}_{2}+i\bm{M}_{2})^{T},\end{gathered} (S36)

where 𝒁1\bm{Z}_{1} and 𝒁2\bm{Z}_{2} are diagonal matrices with elements zk1=γk1+iEk1z_{k_{1}}=\gamma_{k_{1}}+iE_{k_{1}} and zk2=γk2iEk2z_{k_{2}}=\gamma_{k_{2}}-iE_{k_{2}}.

From ^A|ρA(0)=0\hat{\mathcal{L}}_{A}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=0, e^At|ρA(0)=|ρA(0)e^{\hat{\mathcal{L}}_{A}t^{\prime}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}. We have,

^j+|ρA(0)=(A^ji2L^1j)|ρA(0)=k1𝕀1((𝑽1)jk1i(𝑴1)jk1)d^k1|ρA(0),^j|ρA(0)=(A^j+i2L^2j)|ρA(0)=k2𝕀2((𝑽2)jk2+i(𝑴2)jk2)d^k2|ρA(0).\begin{gathered}\hat{\mathcal{F}}^{+}_{j}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=(\hat{A}_{j}^{\dagger}-\frac{i}{2}\hat{L}_{1j}^{\dagger})\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\sum_{k_{1}\in\mathbb{I}_{1}}((\bm{V}_{1})^{*}_{jk_{1}}-i(\bm{M}_{1})^{*}_{jk_{1}})\hat{d}_{k_{1}}^{\dagger}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle},\\ \hat{\mathcal{F}}^{-}_{j}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=(\hat{A}_{j}+\frac{i}{2}\hat{L}_{2j})\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\sum_{k_{2}\in\mathbb{I}_{2}}((\bm{V}_{2})_{jk_{2}}+i(\bm{M}_{2})_{jk_{2}})\hat{d}_{k_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}.\end{gathered} (S37)

For k1𝕀1k_{1}\in\mathbb{I}_{1}, ^Ad^k1|ρA(0)=(iEk1γk1)d^k1|ρA(0)\hat{\mathcal{L}}_{A}\hat{d}_{k_{1}}^{\dagger}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=(-iE_{k_{1}}-\gamma_{k_{1}})\hat{d}_{k_{1}}^{\dagger}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}, and for k2𝕀2k_{2}\in\mathbb{I}_{2}, ^Ad^k2|ρA(0)=(iEk2γk2)d^k2|ρA(0)\hat{\mathcal{L}}_{A}\hat{d}_{k_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=(iE_{k_{2}}-\gamma_{k_{2}})\hat{d}_{k_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}. Hence,

e^At^j+|ρA(0)=k1𝕀1e(iEk1γk1)t((𝑽1)jk1i(𝑴1)jk1)d^k1|ρA(0),e^At^j|ρA(0)=k2𝕀2e(iEk2γk2)t((𝑽2)jk2+i(𝑴2)jk2)d^k2|ρA(0).\begin{gathered}e^{\hat{\mathcal{L}}_{A}t}\hat{\mathcal{F}}_{j}^{+}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\sum_{k_{1}\in\mathbb{I}_{1}}e^{(-iE_{k_{1}}-\gamma_{k_{1}})t}((\bm{V}_{1})^{*}_{jk_{1}}-i(\bm{M}_{1})^{*}_{jk_{1}})\hat{d}_{k_{1}}^{\dagger}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle},\\ e^{\hat{\mathcal{L}}_{A}t}\hat{\mathcal{F}}_{j}^{-}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}=\sum_{k_{2}\in\mathbb{I}_{2}}e^{(iE_{k_{2}}-\gamma_{k_{2}})t}((\bm{V}_{2})_{jk_{2}}+i(\bm{M}_{2})_{jk_{2}})\hat{d}_{k_{2}}\mathinner{|{\rho_{A}(0)}\rangle\!\rangle}.\end{gathered} (S38)

From the conjugation rule in Eq. S29, I|d^k=I|d~^k\mathinner{\langle\!\langle{I}|}\hat{d}_{k}^{\dagger}=\mathinner{\langle\!\langle{I}|}\hat{\tilde{d}}_{k}^{\dagger} and I|d^k=I|d~^k\mathinner{\langle\!\langle{I}|}\hat{d}_{k}=-\mathinner{\langle\!\langle{I}|}\hat{\tilde{d}}_{k}. Therefore,

I|^j=I|(A^ji2L^1ji2L^2j)=k𝕀1𝕀2I|(VjkiMjk)d^k,I|^j+=I|(A^j+i2L^1j+i2L^2j)=k𝕀1𝕀2I|(VjkiMjk)d^k.\begin{gathered}\mathinner{\langle\!\langle{I}|}\hat{\mathcal{F}}_{j}^{-}=\mathinner{\langle\!\langle{I}|}\left(\hat{A}_{j}-\frac{i}{2}\hat{L}_{1j}-\frac{i}{2}\hat{L}_{2j}\right)=\sum_{k\in\mathbb{I}_{1}\cup\mathbb{I}_{2}}\mathinner{\langle\!\langle{I}|}(V_{jk}-iM_{jk})\hat{d}_{k},\\ \mathinner{\langle\!\langle{I}|}\hat{\mathcal{F}}_{j}^{+}=\mathinner{\langle\!\langle{I}|}\left(\hat{A}_{j}^{\dagger}+\frac{i}{2}\hat{L}^{\dagger}_{1j}+\frac{i}{2}\hat{L}^{\dagger}_{2j}\right)=\sum_{k\in\mathbb{I}_{1}\cup\mathbb{I}_{2}}\mathinner{\langle\!\langle{I}|}(V_{jk}-iM_{jk})^{*}\hat{d}^{\dagger}_{k}.\end{gathered} (S39)

Combining Eq. S38 and Eq. S39, the BCFs are expressed as,

𝑪>A(t+t,t)=(𝑽1i𝑴1)e𝒁1t(𝑽1+i𝑴1),𝑪<A(t+t,t)=(𝑽2i𝑴2)e𝒁2t(𝑽2+i𝑴2)T,\begin{gathered}\bm{C}^{>A}(t+t^{\prime},t^{\prime})=(\bm{V}_{1}-i\bm{M}_{1})e^{-\bm{Z}_{1}t}(\bm{V}_{1}+i\bm{M}_{1})^{\dagger},\\ \bm{C}^{<A}(t+t^{\prime},t^{\prime})=(\bm{V}_{2}-i\bm{M}_{2})^{*}e^{-\bm{Z}_{2}t}(\bm{V}_{2}+i\bm{M}_{2})^{T},\end{gathered} (S40)

The equivalence condition with the BCF in the fermionic bath can be similarly constructed with the Dyson series expansion as in the bosonic bath. For a rigorous construction of the fermionic influence superoperator, we refer to Ref. [77].

SM-IV Optimal gauge choice minimizing violation of CP condition

In the single coupling limit with a diagonal 𝒁\bm{Z}, the relationship between the weight ww and the coupling parameters VV and MM is established as (ViM)(ViM)=w(V-iM)(V^{*}-iM^{*})=w. Here, w,V,Mw,V,M\in\mathbb{C}. Once ww is given, the coupling parameters VV and MM have a gauge choice parameterized as ViM=κw,V+iM=κ1wV-iM=\kappa\sqrt{w},V+iM=\kappa^{*-1}\sqrt{w^{*}} where κ\kappa\in\mathbb{C}. We show that the solution of VV and MM with minimum |M||M| is then given by ViM=wV-iM=\sqrt{w} with V,MV,M\in\mathbb{R}, (κ=1\kappa=1), or V=Re(w),M=Im(w)V=\operatorname{Re}(\sqrt{w}),\>M=-\operatorname{Im}(\sqrt{w}).

Proposition 1.

Given ww\in\mathbb{C}, for V,M\forall\ V,M\in\mathbb{C} satisfying the constraint, (ViM)(ViM)=w(V-iM)(V^{*}-iM^{*})=w, and its real solution, Vr,MrV_{r},M_{r}\in\mathbb{R}, Vr=Re(w),Mr=Im(w)V_{r}=\operatorname{Re}(\sqrt{w}),M_{r}=-\operatorname{Im}(\sqrt{w}), then |M||Mr||M|\geq|M_{r}|.

Proof.

With a parameterization, V=|V|eiθ1V=|V|e^{i\theta_{1}}, M=|M|eiθ2M=|M|e^{i\theta_{2}}, the constraint becomes, |V|2|M|22i|V||M|cos(θ1θ2)=w|V|^{2}-|M|^{2}-2i|V||M|\cos(\theta_{1}-\theta_{2})=w or |V|2|M|2=Re(w)|V|^{2}-|M|^{2}=\operatorname{Re}(w) and |V||M|cos(θ1θ2)=Im(w)/2|V||M|\cos(\theta_{1}-\theta_{2})=-\operatorname{Im}(w)/2. It leads to

|M|Re(w)+|M|2|cos(θ1θ2)|=|Im(w)|/2.|M|\sqrt{\operatorname{Re}(w)+|M|^{2}}|\cos(\theta_{1}-\theta_{2})|=|\operatorname{Im}(w)|/2.

Since |M|Re(w)+|M|2|M|\sqrt{\operatorname{Re}(w)+|M|^{2}} is an increasing function in |M||M|, we have minimum |M||M| when |cos(θ1θ2)|=1|\cos(\theta_{1}-\theta_{2})|=1, which corresponds to θ1=θ2\theta_{1}=\theta_{2} or θ1=θ2±π\theta_{1}=\theta_{2}\pm\pi. The real solution Vr,MrV_{r},M_{r} satisfies this condition on θ1\theta_{1} and θ2\theta_{2}, so |M||Mr||M|\geq|M_{r}|. ∎

SM-V Proofs of the stability condition in the noninteracting fermionic impurity model

Recall that for a noninteracting fermionic impurity model, its one-particle reduced density matrix (1-RDM) Ppq(t)=Tr[c^qc^pρ^SA(t)]P_{pq}(t)=\operatorname{Tr}[\hat{c}_{q}^{\dagger}\hat{c}_{p}\hat{\rho}_{SA}(t)] satisfies a continuous differential Lyapunov equation [22, 58]:

t𝑷=𝑿𝑷+𝑷𝑿+𝒀,\partial_{t}\bm{P}=\bm{X}\bm{P}+\bm{P}\bm{X}^{\dagger}+\bm{Y}, (S41)

where

𝑿=(i𝑯Si𝑽1𝑴1i𝑽2𝑴2i𝑽1𝑴1𝒁1𝟎i𝑽2𝑴2𝟎𝒁2),𝒀=(𝟎𝟎2𝑴2𝟎𝟎𝟎2𝑴2𝟎2𝚪2).\begin{gathered}\bm{X}=\left(\begin{array}[]{ccc}-i\boldsymbol{H}_{S}&-i\bm{V}_{1}-\bm{M}_{1}&-i\bm{V}_{2}-\bm{M}_{2}\\ -i\bm{V}_{1}^{\dagger}-\bm{M}_{1}^{\dagger}&-\bm{Z}_{1}&\bm{0}\\ -i\bm{V}_{2}^{\dagger}-\bm{M}_{2}^{\dagger}&\bm{0}&-\bm{Z}_{2}\end{array}\right),\\ \bm{Y}=\left(\begin{array}[]{ccc}\bm{0}&\bm{0}&2\bm{M}_{2}\\ \bm{0}&\bm{0}&\bm{0}\\ 2\bm{M}_{2}^{\dagger}&\bm{0}&2\bm{\Gamma}_{2}\end{array}\right).\end{gathered} (S42)
Proposition 2.

Eq. S41 is asymptotically stable if and only if the real part of all eigenvalues of 𝑿\bm{X} is strictly negative.

Proof.

Let Vec(𝑷)\text{Vec}(\bm{P}) denote the vectorization of 𝑷\bm{P}. Then Eq. S41 can be rewritten as:

tVec(𝑷)=(𝑿𝑰+𝑰𝑿)Vec(𝑷)+Vec(𝒀).\partial_{t}\text{Vec}(\bm{P})=\left(\bm{X}\otimes\bm{I}+\bm{I}\otimes\bm{X}^{*}\right)\text{Vec}(\bm{P})+\operatorname{Vec}(\bm{Y}). (S43)

Therefore, 𝑷(t)\bm{P}(t) is asymptotically stable if and only if all eigenvalues of 𝑿𝑰+𝑰𝑿\bm{X}\otimes\bm{I}+\bm{I}\otimes\bm{X}^{*} have strictly negative real parts. Note that (λ,𝒖)(\lambda,\bm{u}) is an eigen-pair of 𝑿𝑰\bm{X}\otimes\bm{I} if and only if 𝒖=u1u2\bm{u}=u_{1}\otimes u_{2}^{*} and 𝑿u1=λu1\bm{X}u_{1}=\lambda u_{1}, and similarly for 𝑰𝑿\bm{I}\otimes\bm{X}^{*}. Also, note that 𝑿𝑰\bm{X}\otimes\bm{I} and 𝑰𝑿\bm{I}\otimes\bm{X}^{*} commutes. Therefore λ~ij=λi+λj\widetilde{\lambda}_{ij}=\lambda_{i}+\lambda_{j}^{*} is the eigenvalue of 𝑿𝑰+𝑰𝑿\bm{X}\otimes\bm{I}+\bm{I}\otimes\bm{X}^{*} where λi,λj\lambda_{i},\lambda_{j} are the eigenvalues of 𝑿\bm{X}. Therefore, Eq. S41 is asymptotically stable if and only if all eigenvalues of 𝑿\bm{X} have strictly negative real parts. ∎

For simplicity, we now assume a single-site system, i.e., H^S=ha^a^\hat{H}_{S}=h\hat{a}^{\dagger}\hat{a} where hh\in\mathbb{R}, and 𝑿\bm{X} is

𝑿=(ihi𝑽𝑴i𝑽𝑴𝒁),\bm{X}=\left(\begin{array}[]{cc}-ih&-i\bm{V}-\bm{M}\\ -i\bm{V}^{\dagger}-\bm{M}^{\dagger}&-\bm{Z}\end{array}\right), (S44)

where we do not separate baths 1 and 2 since there is no difference in the expression for 𝑿\bm{X}. The matrices 𝑽\bm{V} and 𝑴\bm{M} are now row vectors of size 1×Nb1\times N_{b}, where NbN_{b} is the total number of bath modes. When needed, we will interchangeably refer to these matrices as vectors.

Proposition 3.

When 𝑴<12minγk\|\bm{M}\|<\frac{1}{2}\min\gamma_{k} for k\forall k, there exists m>0m>0, such that if 𝑽>m\|\bm{V}\|>m, the dynamics Eq. S41 with 𝑿\bm{X} being Eq. S44 is asymptotically stable.

Proof.

With Proposition 2, Eq. S41 being asymptotically stable is equivalent to tu=𝑿u\partial_{t}u=\bm{X}u being asymptotically stable for any initial vector u(0)u(0). To study the stability of tu=𝑿u\partial_{t}u=\bm{X}u, let us define a Lyapunov function 𝒱(u)=12u𝚯u\mathcal{V}(u)=\frac{1}{2}u^{\dagger}\bm{\Theta}u, for which the quadratic form 𝚯\bm{\Theta} will be specified later. Then we have

𝒱˙(u(t))=12u(𝚯𝑿+𝑿𝚯)u,\dot{\mathcal{V}}(u(t))=\frac{1}{2}u^{\dagger}(\bm{\Theta}\bm{X}+\bm{X}^{\dagger}\bm{\Theta})u,

Then, by the Lyapunov stability theorem, if 𝚯𝑿+𝑿𝚯\bm{\Theta}\bm{X}+\bm{X}^{\dagger}\bm{\Theta} is negative definite, then tu=𝑿u\partial_{t}u=\bm{X}u is asymptotically stable. Let us consider

𝚯=(1iϵ𝑽1+|𝑽|2iϵ𝑽1+|𝑽|2𝑰).\bm{\Theta}=\left(\begin{array}[]{cc}1&-i\frac{\epsilon\boldsymbol{V}}{1+|\boldsymbol{V}|^{2}}\\ i\frac{\epsilon\boldsymbol{V}^{\dagger}}{1+|\boldsymbol{V}|^{2}}&\boldsymbol{I}\end{array}\right). (S45)

Here ϵ\epsilon are parameters that will be determined later. We define Q(𝑽)=(𝚯𝑿+𝑿𝚯)Q(\boldsymbol{V})=-(\bm{\Theta}\bm{X}+\bm{X}^{\dagger}\bm{\Theta}). Let 𝑽=c𝑽~\boldsymbol{V}=c\widetilde{\boldsymbol{V}} where 𝑽~\widetilde{\boldsymbol{V}} is a unit vector, then we have

limcQ(c𝑽~)=limc(𝚯𝑿+𝑿𝚯)=(2ϵ2𝑴2𝑴2𝚪2ϵ𝑽~𝑽~):=Q.\lim_{c\rightarrow\infty}Q(c\widetilde{\boldsymbol{V}})=-\lim_{c\rightarrow\infty}(\bm{\Theta}\bm{X}+\bm{X}^{\dagger}\bm{\Theta})=\left(\begin{array}[]{cc}2\epsilon&2\boldsymbol{M}\\ 2\boldsymbol{M}^{\dagger}&2\boldsymbol{\Gamma}-2\epsilon\widetilde{\boldsymbol{V}}^{\dagger}\widetilde{\boldsymbol{V}}\end{array}\right):=Q_{\infty}.

If we let ϵ=12minγk\epsilon=\frac{1}{2}{\min\gamma_{k}}, 𝚪2ϵ𝑰\boldsymbol{\Gamma}\geq 2\epsilon\boldsymbol{I}, and Q(2ϵ2𝑴2𝑴4ϵ𝑰2ϵ𝑽~𝑽~)(2ϵ2𝑴2𝑴2ϵ𝑰)Q_{\infty}\geq\left(\begin{array}[]{cc}2\epsilon&2\boldsymbol{M}\\ 2\boldsymbol{M}^{\dagger}&4\epsilon\boldsymbol{I}-2\epsilon\widetilde{\boldsymbol{V}}^{\dagger}\widetilde{\boldsymbol{V}}\end{array}\right)\geq\left(\begin{array}[]{cc}2\epsilon&2\boldsymbol{M}\\ 2\boldsymbol{M}^{\dagger}&2\epsilon\boldsymbol{I}\end{array}\right). Since 𝑴<ϵ=12minγk\|\boldsymbol{M}\|<\epsilon=\frac{1}{2}\min\gamma_{k}, then Q>0Q_{\infty}>0. Therefore, for any unit vector 𝑽~0\widetilde{\boldsymbol{V}}_{0}, there exists a constant c0>0c_{0}>0 depending on 𝑽~0\widetilde{\boldsymbol{V}}_{0} such that if c>c0c>c_{0}, then Q=Q(c𝑽~0)>0Q=Q(c\widetilde{\boldsymbol{V}}_{0})>0. Since both QQ and c0c_{0} depend continuously on 𝑽~\widetilde{\boldsymbol{V}}, and the unit sphere is compact, there exists a positive constant mm such that when c=𝑽>mc=\|\boldsymbol{V}\|>m, we have Q>0Q>0 and the dynamics is asymptotically stable. ∎

Proposition 3 implies that the dynamics, though violating the positivity condition, is stable with sufficiently large system-bath Hamiltonian coupling H^SA\hat{H}_{SA} and small Lindblad coupling 𝒟SA\mathcal{D}_{SA}.

SM-VI Details of Fitting procedures

SM-VI.1 ESPRIT algorithm for complex exponential function fitting with complex weights

In this section, we briefly describe the Estimation of Signal Parameters via Rotational Invariant Techniques (ESPRIT) algorithm [37] used for fitting the BCF as a sum of complex exponential functions with complex weights, C(t)kwkezktC(t)\approx\sum_{k}w_{k}e^{-z_{k}t}, where wkw_{k} and zkz_{k} are complex-valued. The ESPRIT algorithm has been widely adopted in a signal processing community [81, 82, 83, 38, 84], and has recently been applied to quantum computing applications [85, 86]. In this work, we take the ESPRIT algorithm based on the following Hankel matrix [38]:

H=[C(t0)C(t1)C(t2)C(tn1)C(t1)C(t2)C(t3)C(tn)C(t2)C(t3)C(t4)C(tn+1)C(tn1)C(tn)C(tn+1)C(t2n2)]H=\left[\begin{array}[]{ccccc}C(t_{0})&C(t_{1})&C(t_{2})&\cdots&{C(t_{n-1})}\\ C(t_{1})&C(t_{2})&C(t_{3})&\cdots&C(t_{n})\\ C(t_{2})&C(t_{3})&C(t_{4})&\cdots&{C(t_{n+1})}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ C(t_{n-1})&C(t_{n})&C(t_{n+1})&\cdots&C(t_{2n-2})\end{array}\right] (S46)

or Hab=C(ta+b)H_{ab}=C(t_{a+b}), where ta=aδtt_{a}=a\delta t after discretizing time with a timestep δt\delta t. The ESPRIT algorithm is summarized in the following steps to obtain the complex exponent zkz_{k}.

  • Step 1: Perform the singular-value decomposition (SVD) of the matrix HH: H=UΣVH=U\Sigma V, where the singular values (diagonal entries of the diagonal matrix Σ\Sigma) are ordered non-increasingly.

  • Step 2: let U1=U[1:n1,1:Nexp]U_{1}=U[1:n-1,1:N_{\exp}], U2=U[2:n,1:Nexp]U_{2}=U[2:n,1:N_{\exp}]. Calculate W=U1+U2W=U_{1}^{+}U_{2}, where U1+=(U1U1)1U1U_{1}^{+}=(U_{1}^{\dagger}U_{1})^{-1}U_{1}^{\dagger} is the Moore-Penrose pseudo inverse of U1U_{1}.

  • Step 3: Calculate the eigenvalue λ1,,λNexp\lambda_{1},\cdots,\lambda_{N_{\exp}} of the matrix WW. Let zk=log(λk)/Δtz_{k}=-\log(\lambda_{k})/\Delta t, where we select the branch of log()\log(\cdot) such that Im(log(λk))[π,π)\operatorname{Im}(\log(\lambda_{k}))\in[-\pi,\pi).

After we get the complex exponent zkz_{k}, the complex weights wkw_{k} are obtained via the following least-squares fitting:

min{wk}a|C(ta)kwkezkta|2.\min_{\{w_{k}\}}\sum_{a}\left|C(t_{a})-\sum_{k}w_{k}e^{-z_{k}t_{a}}\right|^{2}. (S47)

In the multi-site case, the BCFs are given by a matrix-valued function, Cjj(t)k(𝑾k)jjezktC_{jj^{\prime}}(t)\approx\sum_{k}(\bm{W}_{k})_{jj^{\prime}}e^{-z_{k}t}. We construct a scalar quantity C(t)C(t) by summing over all the entries in Cjj(t)C_{jj^{\prime}}(t), i.e., C(t)=jjCjj(t)C(t)=\sum_{jj^{\prime}}C_{jj^{\prime}}(t), to estimate zkz_{k} as above. Then, we obtain the matrix-valued complex weights, (𝑾k)jj(\bm{W}_{k})_{jj^{\prime}}, with the elementwise least-squares fitting with Eq. S47.

SM-VI.2 Complex exponential function fitting with positive weights

As a comparison to the complex exponential function fitting with complex weights, we performed the complex exponential function fitting with positive weights (i.e., wk+w_{k}\in\mathbb{R}^{+}), which is a BCF decomposition for the Lorentzian pseudomode model, with numerical fitting using the L-BFGS-B algorithm. We minimized the following fitting error function,

fit=δta|C(ta)kwkezkta|2.\mathcal{E}_{\text{fit}}=\delta t\sqrt{\sum_{a}\left|C(t_{a})-\sum_{k}w_{k}e^{-z_{k}t_{a}}\right|^{2}}. (S48)

During the optimization steps, we only used the gradient with respect to zkz_{k} and determined wkw_{k} from the non-negative least-squares fitting following [87]. As an initial guess for the optimization problem, we followed the deterministic procedure from Ref. [22, 23, 24, 25], which determines the imaginary part of zkz_{k} from the direct discretization of the unitary bath and sets the real part of zkz_{k} to some finite values. We chose the imaginary part of zkz_{k} from a set of Gauss-Legendre quadrature nodes on the real axis, and the real part of zkz_{k} as Re[zk]=0.1×W/Nexp\text{Re}[z_{k}]=0.1\times W/N_{\exp} where WW is a width of the spectrum.

In the multi-site case, the fitting error function is modified as,

fit=δtjja|Cjj(ta)k(𝑾k)jjezkta|2,\mathcal{E}_{\text{fit}}=\delta t\sqrt{\sum_{jj^{\prime}}\sum_{a}\left|C_{jj^{\prime}}(t_{a})-\sum_{k}(\bm{W}_{k})_{jj^{\prime}}e^{-z_{k}t_{a}}\right|^{2}}, (S49)

with the constraint that 𝑾k0\bm{W}_{k}\geq 0.

SM-VII Details on DMRG calculations

We describe some details of the time-dependent density matrix renormalization group (tdDMRG) calculations, focusing on the fermionic impurity models. We first express the fermionic Lindblad master equation in the super-fermion representation. This allows us to represent the density operator as a state vector with Schrödinger equation-like propagation so that the two-site time-dependent variational principle, implemented in the Block2 [61, 62] package, can be directly used for the state vector propagation. Furthermore, the Liouville operator has a number-conserving form in the super-fermion representation, so we can take advantage of 𝕌(1)\mathbb{U}(1) symmetry adaptation in the tdDMRG calculations.

Ordering of orbitals is important in DMRG calculations. We employed two different ordering schemes illustrated in Fig. S1. The illustration is based on the spinless single-site impurity model with two bath orbitals in each bath A1A_{1} and A2A_{2}. We have ρ^SA(0)=ρ^S(0)ρ^A1(0)ρ^A2(0)\hat{\rho}_{SA}(0)=\hat{\rho}_{S}(0)\otimes\hat{\rho}_{A_{1}}(0)\otimes\hat{\rho}_{A_{2}}(0), where ρ^A1(0)=|0000|\hat{\rho}_{A_{1}}(0)=\mathinner{|{00}\rangle}\!\!\mathinner{\langle{00}|} and ρ^A2(0)=|1111|\hat{\rho}_{A_{2}}(0)=\mathinner{|{11}\rangle}\!\!\mathinner{\langle{11}|}. In the super-fermion representation, |ρSA(0)=|ρS(0)|ρA1(0)|ρA2(0)\mathinner{|{\rho_{SA}(0)}\rangle\!\rangle}=\mathinner{|{\rho_{S}(0)}\rangle\!\rangle}\otimes\mathinner{|{\rho_{A_{1}}(0)}\rangle\!\rangle}\otimes\mathinner{|{\rho_{A_{2}}(0)}\rangle\!\rangle} where |ρA1(0)=|0101\mathinner{|{\rho_{A_{1}}(0)}\rangle\!\rangle}=\mathinner{|{0101}\rangle\!\rangle} and |ρA2(0)=|1010\mathinner{|{\rho_{A_{2}}(0)}\rangle\!\rangle}=\mathinner{|{1010}\rangle\!\rangle}. The orbitals from the ‘tilde’ space are drawn with hatching lines in Fig. S1. The two orbital ordering schemes are distinguished by whether the bath A1A_{1} and A2A_{2} are on the same side (Fig. S1a) or different sides (Fig. S1b) of the impurity. We chose the first scheme for the spinful impurity model to put two different spins on different sides of the impurity [88]. We chose the second scheme for the spinless impurity model. The remaining ordering of the orbitals within each bath is determined from the imaginary part of zkz_{k}, or equivalently, the energy part, EkE_{k}, following the physical insights in [89, 90].

Refer to caption
Figure S1: Two different orbital ordering schemes for tdDMRG calculations.