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

Integral fluctuation theorems and trace-preserving map

Zhiqiang Huang (黄志强) [email protected] Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430071, China
Abstract

The detailed fluctuation theorem implies symmetry in the generating function of entropy production probability. The integral fluctuation theorem directly follows from this symmetry and the normalization of the probability. In this paper, we rewrite the generating function by integrating measurements and evolution into a constructed mapping. This mapping is completely positive, and the original integral FT is determined by the trace-preserving property of these constructed maps. We illustrate the convenience of this method by discussing the eigenstate fluctuation theorem and heat exchange between two baths. This set of methods is also applicable to the generating functions of quasi-probability, where we observe the Petz recovery map arising naturally from this approach.

I Introduction

As the generalized second law of thermodynamics from a microscopic perspective, the fluctuation theorem (FT) is of fundamental importance in non-equilibrium statistical mechanics. The FT focuses mainly on the irreversibility of entropy production. While a general FT can be given under unitary evolution [1], it is only useful when the entropy production can be expressed in terms of physical and measurable quantities, which usually requires the initial state of the bath to be the canonical distribution. However, when this condition is not met, the FT will deviate. By calculating the characteristic (generating) function of entropy production, [2, 3] proved that the integral FT approximation holds in both the long and short time regimes, even when the initial state of the bath is a single energy eigenstate of many-body systems. The proof of this result is complex and uses various forms of the eigenstate thermalization hypothesis (ETH), which prevents further development of the approaches used therein and makes it difficult to tighten the errors. The difficulty in estimating deviation stems from the lack of a unified framework to study the impact of initial state deviation, as the generating function also includes two-point measurements and evolution, each of which can have additional effects based on the initial state deviation. The forms of these operators are quite different in conventional quantum mechanics formalism, which makes deviation estimation even more difficult.

The integral FT is directly related to the normalization of the backward processes, which in turn depends on the trace preservation (TP) property of the backward map. In this sense, the generating function should be determined by the properties of some mapping. The modified propagator approach is one such method [1]. It incorporates measurements and Fourier transforms into the modified propagator. Here we propose a different approach. We start with the general FT under unitary evolution and incorporate measurements and Fourier transforms into quantum operations. These quantum operations, along with the evolution, form a complete positive (CP) map. If this constructed map is trace preserving when λ=i\lambda=i, then the generating function G(i)=1G(i)=1 and the integral FT holds. Since the generating function is just rewritten here, it is equivalent to the original FT. However, by unifying the measurements and evolution into the constructed mapping, considering the impact of initial state deviation becomes simpler and more unified. To illustrate this advantage, we first consider the general quantum lattice model in [2, 3]. Our aim is to provide a streamlined proof of the eigenstate FT, relying on fewer assumptions (specifically, the weak ETH assumption rather than several strong versions of ETH). After that, we consider another important class of FTs concerning the heat exchange between two baths [4]. We will investigate whether two baths, whose initial states consist of single-energy eigenstates, also exhibit FT.

The exploration of FTs through the construction of CP mappings has long been proposed by [5, 6]. These studies primarily delve into the properties of the evolution of quantum open systems. Their approaches are closely connected to the FTs for the quantum channel [7, 8], which has established another general framework of quantum FTs. Using the Petz recovery map, it provides the detailed FT for two-point measurement quasiprobabilities. Notably, the resulting FT does not incorporate measurements of the environment. This distinction highlights a fundamental difference from the entropy production considered in the FT under unitary evolution. However, since the quantum channel can emerge naturally from the open system formalism, there likely exists a relationship between the quantum channel FTs and the unitary evolution FTs. If the global unitary operation UU has a global fixed point, [9] demonstrates that both methods can yield the same entropy production. Correspondingly, by employing the approach outlined in this paper, we demonstrate that the Petz recovery map can be naturally derived by analyzing the generating function of quasi-probabilities. This offers an alternative perspective on the relationship between the two methods.

The paper is organized as follows: In section II, we first briefly introduce the operator-state formalism. Then, we rewrite the generating function with complete positive mappings and briefly discuss the properties of the constructed mappings. We delve into the core component 𝒩β(t)\mathcal{N}_{\beta}(t) of the constructed mapping in detail in appendix B. In section III and section IV, we utilize the constructed mapping to discuss the eigenstate FT. The integral FT is shown to hold within a small margin of error, and a detailed calculation of error estimation can be found in appendix C. In section V, we discuss the integral FT for the quantum channel and uncover the significance of the corresponding constructed mapping. Section VI concludes.

II The integral fluctuation theorem and trace-preserving map

II.1 Preliminaries

Since we are mainly concerned with CP mappings in this paper, we use the superoperator form for simplicity. Furthermore, the FT necessitates the consideration of inverse mappings and measurements. These components have simpler expressions within the operator-state formalism, which we also employ. It’s noteworthy that these formalisms maintain mathematical equivalence with the conventional framework of quantum mechanics.

Let us first introduce the operator state |O)|O) for an operator OO on the Hilbert space S\mathcal{H}_{S} of quantum states of a system SS, which belongs to a new Hilbert space 𝖧\mathsf{H} with an inner product defined by

(O1|O2):=Tr(O1O2).(O_{1}|O_{2}):=\text{Tr}(O_{1}^{\dagger}O_{2}). (1)

An orthonormal basis for 𝖧\mathsf{H} is |Πij)|\Pi_{ij}) where Πij=|ij|\Pi_{ij}=\ket{i}\bra{j}, as one can check that

(Πkl|Πij)=δikδjl.(\Pi_{kl}|\Pi_{ij})=\delta_{ik}\delta_{jl}. (2)

The completeness of this basis {|Πij)}\{|\Pi_{ij})\} is

=ij|Πij)(Πij|.\mathcal{I}=\sum_{ij}|\Pi_{ij})(\Pi_{ij}|. (3)

If only the diagonal elements are summed, the corresponding superoperator is actually a dephasing map

i|Πi)(Πi|ρ)=|i|ii|ρ|ii|).\sum_{i}|\Pi_{i})(\Pi_{i}|\rho)=|\sum_{i}\ket{i}\bra{i}\rho\ket{i}\bra{i}). (4)

The conjugate relation reads

(O1|O2):=Tr(O1O2)=Tr(O2O1)\displaystyle(O_{1}|O_{2})^{\dagger}:=\text{Tr}(O_{1}^{\dagger}O_{2})^{\dagger}=\text{Tr}(O_{2}^{\dagger}O_{1})
=(O2|O1)=(O1|O2).\displaystyle=(O_{2}|O_{1})=(O_{1}^{\dagger}|O_{2}^{\dagger}). (5)

The unitary superoperator acting on the operator state gives

𝒰|O):=|UOU).\mathcal{U}|O):=|UOU^{\dagger}). (6)

The conjugate relation shows that

(O1|𝒰|O2)=Tr(O1UO2U)=(O1|𝒰|O2)\displaystyle(O_{1}|\mathcal{U}|O_{2})^{\dagger}=\text{Tr}(O_{1}UO_{2}^{\dagger}U^{\dagger})=(O_{1}^{\dagger}|\mathcal{U}|O_{2}^{\dagger})
=Tr(O2UO1U)=(O2|𝒰|O1).\displaystyle=\text{Tr}(O^{\dagger}_{2}U^{\dagger}O_{1}U)=(O_{2}|\mathcal{U}^{\dagger}|O_{1}). (7)

The trace of an operator ρ\rho can be given by (I|ρ)(I|\rho). The superoperator 𝒩\mathcal{N} is trace-preserving if and only if it satisfies (I|𝒩=(I|(I|\mathcal{N}=(I|.

Now we briefly introduce the FT. The FT can be framed in a unifying language in terms of entropy production [9]. The entropy production Δσ\Delta\sigma is generally related to thermodynamic quantities such as Gibbs-von Neumann entropy, work, heat, and free energy. Its formulation depends on the underlying physical system. In detailed FT (Crooks FT), the forward entropy production is exponentially more likely than the reverse

PF(Δσ)PB(Δσ)=eΔσ,\frac{P_{F}(\Delta\sigma)}{P_{B}(-\Delta\sigma)}=e^{\Delta\sigma}, (8)

where PFP_{F} is the forward probability distribution of entropy production and PBP_{B} is the backward one. They are given by the probability density function of the difference between two measurements

PF(Δσ)=σt,σ0δ(Δσ(σtσ0))PF(σt,σ0).P_{F}(\Delta\sigma)=\sum_{\sigma_{t},\sigma_{0}}\delta(\Delta\sigma-(\sigma_{t}-\sigma_{0}))P_{F}(\sigma_{t},\sigma_{0}). (9)

Its generating function

GF(λ):=𝑑ΔσeiλΔσPF(Δσ)G_{F}(\lambda):=\int_{-\infty}^{\infty}d\Delta\sigma e^{i\lambda\Delta\sigma}P_{F}(\Delta\sigma) (10)

is often used. The parameter λ\lambda is a complex number, and people are generally concerned about the properties of the generating function when λ=i\lambda=i. The Crooks FT implies the fundamental symmetry on the generating function

GF(λ)=𝑑ΔσeiλΔσ+ΔσPB(Δσ)=GB(iλ).G_{F}(\lambda)=\int_{-\infty}^{\infty}d\Delta\sigma e^{i\lambda\Delta\sigma+\Delta\sigma}P_{B}(-\Delta\sigma)=G_{B}(i-\lambda). (11)

An integral FT immediately follows from the normalization of PBP_{B}

eΔσ=GF(i)=GB(0)=1.\braket{e^{-\Delta\sigma}}=G_{F}(i)=G_{B}(0)=1. (12)

Combining it with Jensen’s inequality eXeX\braket{e^{-X}}\geq e^{-\braket{X}}, we can get a generalization of the second law of thermodynamics

ΔσlneX=0.\braket{\Delta\sigma}\geq-\ln\braket{e^{-X}}=0. (13)

II.2 Constructs a completely positive map from the generating function

In this section, we explore how to construct a CP map from the generating function and connect the FT with the TP property of this constructed CP map.

Consider an isolated, possible driven, quantum system evolves according to the unitary evolution. The probability density function of the difference between two measurements can be expressed as

PF(Δσ)=σt,σ0δ(Δσ(σtσ0))(Πσt|𝒰|Πσ0)(Πσ0|ρ0),P_{F}(\Delta\sigma)=\sum_{\sigma_{t},\sigma_{0}}\delta(\Delta\sigma-(\sigma_{t}-\sigma_{0}))(\Pi_{\sigma_{t}}|\mathcal{U}|\Pi_{\sigma_{0}})(\Pi_{\sigma_{0}}|\rho_{0}), (14)

where ρ0\rho_{0} is the initial density matrices for the forward process. The observable σ^t\hat{\sigma}_{t} is a Hermitian operator. The eigenvalues (eigenbasis) of σ^t\hat{\sigma}_{t} are denoted by σt\sigma_{t} (Πσt=|σtσt|\Pi_{\sigma_{t}}=\ket{\sigma_{t}}\bra{\sigma_{t}}): σ^t=σtσtΠσt\hat{\sigma}_{t}=\sum_{\sigma_{t}}\sigma_{t}\Pi_{\sigma_{t}}. The generating function of probability (14) is

GF(λ):=σ0,σt(eiλ(σ^tσ^0)|ΠσtΠσ0)(Πσt|𝒰|Πσ0)(Πσ0|ρ0).G_{F}(\lambda):=\sum_{\sigma_{0},\sigma_{t}}(e^{i\lambda(\hat{\sigma}_{t}-\hat{\sigma}_{0})}|\Pi_{\sigma_{t}}\otimes\Pi_{\sigma_{0}})(\Pi_{\sigma_{t}}|\mathcal{U}|\Pi_{\sigma_{0}})(\Pi_{\sigma_{0}}|\rho_{0}). (15)

The factor eiλσ^e^{-i\lambda\hat{\sigma}} rescaling the measurement results indeed. It is related to the rescaling map 𝒥Oα():=Oα()Oα\mathcal{J}^{\alpha}_{O}(\cdot):=O^{\alpha}(\cdot)O^{\alpha\dagger}, where OO is arbitrary operator and α\alpha is the power index. Here we only consider the cases where the real part of λ\lambda is zero, and then we have

𝒥eσ^0iλ/2|Πσ0)=|eiλσ^0/2Πσ0eiλσ^0/2)\displaystyle\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}|\Pi_{\sigma_{0}})=|e^{-i\lambda\hat{\sigma}_{0}/2}\Pi_{\sigma_{0}}e^{-i\lambda\hat{\sigma}_{0}/2})
=eiλσ0|Πσ0)=(eiλσ^0|Πσ0)|Πσ0).\displaystyle=e^{-i\lambda\sigma_{0}}|\Pi_{\sigma_{0}})=(e^{-i\lambda\hat{\sigma}_{0}}|\Pi_{\sigma_{0}})|\Pi_{\sigma_{0}}). (16)

Using it, we can rewrite some items in the generating function as follows

σ0(eiλσ^0|Πσ0)|Πσ0)(Πσ0|=𝒥eσ^0iλ/2σ^0,\sum_{\sigma_{0}}(e^{-i\lambda\hat{\sigma}_{0}}|\Pi_{\sigma_{0}})|\Pi_{\sigma_{0}})(\Pi_{\sigma_{0}}|=\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}, (17)

where σ^0:=σ0|Πσ0)(Πσ0|\mathcal{M}_{\hat{\sigma}_{0}}:=\sum_{\sigma_{0}}|\Pi_{\sigma_{0}})(\Pi_{\sigma_{0}}| is a dephasing map. The rescaling map is CP, but generally not TP. It is worth pointing out that the rescaling map here is very similar to the re-weighting approach used in [10], where this approach is used to prepare the thermal state. With Eq. 17, we can rewrite Eq. 15 as

GF(λ)=(I|𝒥eσ^tiλ/2σ^t𝒰(t)𝒥eσ^0iλ/2σ^0|ρ0)G_{F}(\lambda)=(I|\mathcal{J}_{e^{-\hat{\sigma}_{t}}}^{-i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{t}}\circ\mathcal{U}(t)\circ\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}|\rho_{0}) (18)

or its conjugated form

GF(λ)=(I|𝒥ρ01/2σ^0𝒥eσ^0iλ/2𝒰(t)|eiλσ^t)G_{F}(\lambda)=(I|\mathcal{J}_{\rho_{0}}^{1/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}\circ\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}\circ\mathcal{U}^{\dagger}(t)|e^{i\lambda\hat{\sigma}_{t}})^{\dagger} (19)

Where σ^t\mathcal{M}_{\hat{\sigma}_{t}} is omitted because it shares the same basis as eiλσ^te^{i\lambda\hat{\sigma}_{t}}. Additionally, we omitted the {\dagger} on \mathcal{M} and 𝒥eσ^\mathcal{J}_{e^{-\hat{\sigma}}} because they are self-conjugate. If {Πσ0}\{\Pi_{\sigma_{0}}\} is the diagonal basis of ρ0\rho_{0}, then σ^0\mathcal{M}_{\hat{\sigma}_{0}} can also be omitted. In Eqs. 18 and 19, the generating function is completely rewritten with the operator state and completely positive maps. As will be shown later, the integral FT will be determined by the TP property of these maps.

Take σ^t=lnρ^t\hat{\sigma}_{t}=-\ln\hat{\rho}_{t}, then we have

eiλσ^0=(ρ^0)iλ.e^{-i\lambda\hat{\sigma}_{0}}=(\hat{\rho}_{0})^{i\lambda}. (20)

In such cases, the σ^0\mathcal{M}_{\hat{\sigma}_{0}} can be omitted

𝒥eσ^0iλ/2σ^0|ρ0)=𝒥eσ^0iλ/2|ρ0).\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}|\rho_{0})=\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{i\lambda/2}|\rho_{0}). (21)

Combining Eqs. 18 and 21, we get

GF(i)=(I|𝒥eσ^t1/2𝒰𝒥eσ^01/2|ρ0)=(ρt|𝒰(t)|I)=1.\displaystyle G_{F}(i)=(I|\mathcal{J}_{e^{-\hat{\sigma}_{t}}}^{1/2}\circ\mathcal{U}\circ\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{-1/2}|\rho_{0})=(\rho_{t}|\mathcal{U}(t)|I)=1. (22)

Therefore, for a closed system, the entropy production ln(ρ^0/ρ^t)\ln(\hat{\rho}_{0}/\hat{\rho}_{t}) satisfies integral FT. Since lnρ^t=S(ρt)\braket{-\ln\hat{\rho}_{t}}=S(\rho_{t}), the average of this entropy production is the same as the change of the Gibbs-von Neumann entropy of the system.

For all cases in this paper, we have Treσ^t=1\text{Tr}e^{-\hat{\sigma}_{t}}=1, which can also be achieved in other cases by adding a normalization factor. Under this condition, according to Eq. 19, the integral FT is satisfied if and only if the following constructed CP map is also a TP map

𝒥ρ01/2σ^0𝒥eσ^01/2𝒰(t).\mathcal{J}_{\rho_{0}}^{1/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}\circ\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{-1/2}\circ\mathcal{U}^{\dagger}(t). (23)

This map provides a new perspective for considering the integral FT. The mapping (23) is not necessarily a TP mapping, it depends on the initial state ρ0\rho_{0} as well as the initial measurements σ^0\hat{\sigma}_{0}. If observable are taken as Eq. 20, it is easy to prove that the constructed map (23) is TP

(I|𝒥ρ01/2σ^0𝒥eσ^01/2𝒰(t)=(I|,(I|\mathcal{J}_{\rho_{0}}^{1/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}\circ\mathcal{J}_{e^{-\hat{\sigma}_{0}}}^{-1/2}\circ\mathcal{U}^{\dagger}(t)=(I|, (24)

which is consistent with the previous conclusion that the integral FT holds.

II.3 The constructed map for the open quantum system

Now consider an open quantum system SS interacting with an environment EE; SS and EE form a closed system with unitary evolution. If there is no correlation between the initial system and the environment, we can set σ^t=lnρ^S(t)lnρ^E\hat{\sigma}_{t}=-\ln\hat{\rho}_{S}(t)-\ln\hat{\rho}_{E}, and then we have

eiλσ^0=ρ^Siλ(0)ρ^Eiλ.e^{-i\lambda\hat{\sigma}_{0}}=\hat{\rho}^{i\lambda}_{S}(0)\otimes\hat{\rho}^{i\lambda}_{E}. (25)

Similar to the case of closed systems, it is easy to prove that open systems also satisfy integral FT

GF(i)=(I|𝒰SE(t)|ρ^S(t)ρ^E)=1.G_{F}(i)=(I|\mathcal{U}^{\dagger}_{SE}(t)|\hat{\rho}_{S}(t)\otimes\hat{\rho}_{E})^{\dagger}=1. (26)

This integral FT is completely general, but only contains quantities like entropy, which require knowledge of all information about the state of the environment. In open quantum systems, the environment is usually considered to be relatively large and obtaining all information about it is difficult. Therefore, it is necessary to introduce some thermodynamic quantities such as work, free energy, and heat to simplify the measurement. For example, it is usually assumed that the initial state of the heat bath is the canonical ensemble ρ^E=eβH^E/ZE\hat{\rho}_{E}=e^{-\beta\hat{H}_{E}}/Z_{E}, so we can set

σ^t=β(H^EFE)lnρ^S(t).\hat{\sigma}_{t}=\beta(\hat{H}_{E}-F_{E})-\ln\hat{\rho}_{S}(t). (27)

If we make no assumptions about the initial state of the environment, but still set σ^t\hat{\sigma}_{t} as Eq. 27, then we have

GF(λ)=(ISE|𝒥ρS(t)ρEcaniλ/2𝒰SE(t)\displaystyle G_{F}(\lambda)=(I_{SE}|\mathcal{J}_{\rho_{S}(t)\otimes\rho^{\text{can}}_{E}}^{-i\lambda/2}\circ\mathcal{U}_{SE}(t)
𝒥ρS(0)ρEcaniλ/2σ^0E|ρS(0)ρE(0)).\displaystyle\circ\mathcal{J}_{\rho_{S}(0)\otimes\rho^{\text{can}}_{E}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}^{E}|\rho_{S}(0)\otimes\rho_{E}(0)). (28)

Since the environment ρE(0)\rho_{E}(0) can deviate from the canonical ensemble ρEcan\rho^{\text{can}}_{E}, the integral FT will have errors and GF(i)G_{F}(i) can deviate from 1.

Refer to caption
Figure 1: A quantum circuit diagram for the generating function Eq. 33. In this rewritten form, the terms related to the operator expansion are grouped as 𝒩β(t)\mathcal{N}^{\dagger}_{\beta}(t). According to the Lieb-Robinson bound, only part of the environment effectively participates in the expansion. Therefore, 𝒩β(t)\mathcal{N}^{\dagger}_{\beta}(t) can be replaced by 𝒩βT(t)\mathcal{N}^{T{\dagger}}_{\beta}(t) with only a small error. After the replacement, according to the ETH, the total environment energy eigenstate is also indistinguishable from the canonical ensemble, so ρE(0)\rho_{E}(0) can be replaced by ρScan\rho_{S}^{\text{can}}. As shown in section III.1, a TP mapping can be obtained after several replacements. Deviations from the integral FT can be derived by considering the errors caused by these replacements.

The deviation of the environment state is not a sufficient condition for the violation of the integral FT, which is also related to the system-environment interaction. For non-driven systems, the Hamiltonian of the whole system can generally be written as H=HS+HE+HIH=H_{S}+H_{E}+H_{I}. If there is no system-environment interaction HI=0H_{I}=0, then the overall generating function can be separated into a system part and an environment part GF(λ)=GFS(λ)GFE(λ)G_{F}(\lambda)=G^{S}_{F}(\lambda)G^{E}_{F}(\lambda). Since the system state does not deviate, we have GFS(i)=1G^{S}_{F}(i)=1. And the environment part gives

GFE(λ)=(IE|𝒥ρEcaniλ/2𝒰E(t)𝒥ρEcaniλ/2σ^0E|ρE(0)).G^{E}_{F}(\lambda)=(I_{E}|\mathcal{J}_{\rho^{\text{can}}_{E}}^{-i\lambda/2}\circ\mathcal{U}_{E}(t)\circ\mathcal{J}_{\rho^{\text{can}}_{E}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}^{E}|\rho_{E}(0)). (29)

Since [UE(t),ρEcan]=0[U_{E}(t),\rho^{\text{can}}_{E}]=0, there is always GFE(i)=1G^{E}_{F}(i)=1, no matter whether ρE(0)\rho_{E}(0) deviates from ρEcan\rho^{\text{can}}_{E} or not. This inspires us to decompose the rescaling map in section II.3 as follows

𝒥ρS(t)ρEcan=𝒥ρScanρEcan𝒥ρScan1𝒥ρS(t).\mathcal{J}_{\rho_{S}(t)\otimes\rho^{\text{can}}_{E}}=\mathcal{J}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}\circ\mathcal{J}^{-1}_{\rho^{\text{can}}_{S}}\circ\mathcal{J}_{\rho_{S}(t)}. (30)

𝒥ρScan1𝒥ρS(t)\mathcal{J}^{-1}_{\rho^{\text{can}}_{S}}\circ\mathcal{J}_{\rho_{S}(t)} is a local superoperator of the system. Using decomposition (30), we can rewrite the generating function (II.3) as

GF(λ)=(ISE|𝒥ρS(t)iλ/2𝒥ρScaniλ/2𝒩β(t)\displaystyle G_{F}(\lambda)=(I_{SE}|\mathcal{J}_{\rho_{S}(t)}^{-i\lambda/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{i\lambda/2}\circ\mathcal{N}_{\beta}(t)
𝒥ρScaniλ/2𝒥ρS(0)iλ/2|ρS(0)ρE(0)),\displaystyle\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-i\lambda/2}\circ\mathcal{J}_{\rho_{S}(0)}^{i\lambda/2}|\rho_{S}(0)\otimes\rho^{\prime}_{E}(0)), (31)

where

𝒩β(t)():=𝒥ρScanρEcaniλ/2𝒰SE(t)𝒥ρScanρEcaniλ/2(),\mathcal{N}_{\beta}(t)(\cdot):=\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}\circ\mathcal{U}_{SE}(t)\circ\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}(\cdot), (32)

and ρE(0)=σ^0E[ρE(0)]\rho^{\prime}_{E}(0)=\mathcal{M}_{\hat{\sigma}_{0}}^{E}[\rho_{E}(0)] is the environment density matrix resulting from the decoherence of the environment due to the measurement.

Similar to Eq. 23, the conjugated form of section II.3 gives 111In Eq. 33, we omit the IEI_{E} in the right parenthesis, which will change the meaning of the first 𝒥ρE1/2\mathcal{J}_{\rho_{E}}^{1/2} in 𝒩FT\mathcal{N}_{\text{FT}}: it becomes an assignment map ρSρSρE\rho_{S}\to\rho_{S}\otimes\rho_{E}, which is linear and CPTP. The similar approaches are also used in section IV.

GF(i)=(ISE|𝒩FT(t)|ρS(t)),G_{F}(i)=(I_{SE}|\mathcal{N}_{\text{FT}}(t)|\rho_{S}(t))^{\dagger}, (33)

and the integral FT holds if and only if the following constructed map is a TP map

𝒩FT(t):=𝒥ρE(0)1/2𝒥ρScan1/2𝒩β(t)𝒥ρScan1/2.\mathcal{N}_{\text{FT}}(t):=\mathcal{J}_{\rho^{\prime}_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}_{\beta}^{\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}. (34)

The key parts of the formulae (34) are the mapping 𝒩β(t)\mathcal{N}_{\beta}(t) related to the evolution and the mapping 𝒥ρE(0)1/2\mathcal{J}_{\rho^{\prime}_{E}(0)}^{1/2} related to the initial state and the initial measurement. If the initial state of the environment is the canonical ensemble, then we have 𝒩FT(t)=𝒰SE(t)𝒥ρEcan1/2\mathcal{N}_{\text{FT}}(t)=\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}. It acts on the operator state and gives 𝒩FT(t)|ρS(t))=𝒰SE(t)|ρS(t)ρEcan)\mathcal{N}_{\text{FT}}(t)|\rho_{S}(t))=\mathcal{U}^{\dagger}_{SE}(t)|\rho_{S}(t)\otimes\rho^{\text{can}}_{E}). Therefore, it is indeed a TP map in this case. When the environment ρE(0)\rho_{E}(0) deviates from the canonical ensemble ρEcan\rho^{\text{can}}_{E}, the constructed map 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) is generally not a TP map. The integral FT will have errors and GF(i)G_{F}(i) can deviate from 1. The inverse temperature in Eq. 34 remains arbitrary, allowing for optimization of the deviation estimate by selecting it based on the state of the environment. For instance, when considering an environment XX initialized in a single-energy eigenstate |EaX\ket{E^{X}_{a}}, the choice of inverse temperature is typically determined by aligning the total energy, often set as βa\beta_{a} such that

Ea=(eβaHX/Zβa|HX).E_{a}=(e^{-\beta_{a}H_{X}}/Z_{\beta_{a}}|H_{X}). (35)

III Eigenstate fluctuation theorem

In this section, we will discuss the properties of 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) and the FT when the initial state of the environment is the energy eigenstate. Since ρE(0)\rho_{E}(0) is the energy eigenstate of H^E\hat{H}_{E}, we have ρE(0)=ρE(0)\rho^{\prime}_{E}(0)=\rho_{E}(0). Our proof draws on the idea in article [3] about using Lieb-Robinson bound for short-time regime and time averaging for long-time regime. But the approach is very different: we mainly consider the deviation of the generating function through the properties of the mapping 𝒩FT(t)\mathcal{N}_{\text{FT}}(t). It is worth pointing out that if we replace the 𝒩β(t)\mathcal{N}_{\beta}(t) in Eq. 33 with 𝒰SE(t)\mathcal{U}_{SE}(t), then we will get the δGS(i)\delta G_{S}(i) in [3]. Therefore, 𝒩β(t)𝒰SE(t)\mathcal{N}_{\beta}(t)-\mathcal{U}_{SE}(t) is directly related to the interaction-induced error δGI\delta G_{I} in [3].

The underlying physical insight behind the proof is as follows: The error of the integral FT is determined by the interaction between the system and the environment. According to the Lieb-Robinson bound [11] or the operator growth hypothesis [12], the influence range of the interaction is limited under the (imaginary) time evolution. In the short-time regime, the expanded operator is still unable to distinguish the pure state of the many-body systems from the canonical ensemble, according to the ETH. In the long-time regime, the expanded energy width is limited, and the expanded operator is unable to distinguish these states according to the ETH.

III.1 Short-time regime

When the evolution time tends to zero, it is obvious that 𝒩FT(t)=𝒥ρE(0)1/2\mathcal{N}_{\text{FT}}(t)=\mathcal{J}_{\rho_{E}(0)}^{1/2}. It acts on the operator state and gives 𝒩FT(t)|ρS(t))=|ρS(t)ρE(0))\mathcal{N}_{\text{FT}}(t)|\rho_{S}(t))=|\rho_{S}(t)\otimes\rho_{E}(0)). Therefore, it is a TP map and the FT holds. If the evolution time is short, it is foreseeable that the error can be bounded, as we will discuss in detail below.

When the evolution time is short, the two measurements about the system can only be influenced by a small piece of environment due to the limitation of information flow speed. Considering the ETH, it is difficult to distinguish the overall pure state of the environment from the canonical bath in this small area. This makes the integral FT approximately valid. From the perspective of constructed map 𝒩FT\mathcal{N}_{\text{FT}}, we can rewrite it as

𝒩FT(t)=𝒥ρE(0)1/2𝒥ρScan1/2[𝒩β(t)𝒩βT(t)]𝒥ρScan1/2\displaystyle\mathcal{N}_{\text{FT}}(t)=\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ[\mathcal{N}_{\beta}^{\dagger}(t)-\mathcal{N}_{\beta}^{T\dagger}(t)]\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}
+[𝒥ρE(0)1/2𝒥ρEcan1/2]𝒥ρScan1/2𝒩βT(t)𝒥ρScan1/2\displaystyle+[\mathcal{J}_{\rho_{E}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}_{\beta}^{T\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}
+𝒥ρEcan1/2𝒥ρScan1/2[𝒩βT(t)𝒩β(t)]𝒥ρScan1/2\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ[\mathcal{N}_{\beta}^{T\dagger}(t)-\mathcal{N}_{\beta}^{\dagger}(t)]\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}
+𝒥ρEcan1/2𝒥ρScan1/2𝒩β(t)𝒥ρScan1/2,\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}_{\beta}^{\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}, (36)

where 𝒩βT(t)\mathcal{N}_{\beta}^{T}(t) is defined in section B.2. It is a truncated version of 𝒩β(t)\mathcal{N}_{\beta}(t) based on the speed limit of information flow. The last term of formula (III.1) is indeed a TP mapping. Therefore, the deviation of the integral FT can be estimated using the bounds of other terms. The terms in the first and third lines of section III.1 represent the error due to truncation, which can be bounded by the Lieb-Robinson bound. See section C.1 for specific calculations. The term in the second line of section III.1 represents the error caused by the deviation of the initial state. Since the mapping 𝒥ρScan1/2𝒩βT(t)𝒥ρScan1/2\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}_{\beta}^{T\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}} acts on state |ρS(t))|\rho_{S}(t)) and gives a local operator, this part of the error can be bounded by ETH. See section C.2 for details. Finally, in the large-environment limit, by combining the error terms in section III.1, we obtain

|δGF(i)||GF(i)GFT(i)|+|GFT(i)GFcan,T(i)|\displaystyle\lvert\delta G_{F}(i)\rvert\leq\lvert G_{F}(i)-G^{T}_{F}(i)\rvert+\lvert G^{T}_{F}(i)-G^{\text{can},T}_{F}(i)\rvert
+|GFcan,T(i)GFcan(i)|=o(1),\displaystyle+\lvert G^{\text{can},T}_{F}(i)-G^{\text{can}}_{F}(i)\rvert=o(1), (37)

where the difference in the generating function corresponds one-to-one with the error terms in section III.1. The superscript ”can” represents replacing the initial environment state in 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) with the canonical ensemble, while ”TT” represents replacing 𝒩β\mathcal{N}_{\beta} in 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) with 𝒩βT\mathcal{N}^{T}_{\beta}. We illustrate these replacements in fig. 1. From the figure, we can observe that when using the truncated mapping 𝒩βT(t)\mathcal{N}_{\beta}^{T}(t), only a portion of the environment will participate, which is why ETH can be applied.

III.2 Long-time regime

When the evolution time is long, the system can gather sufficient information about the environment, allowing it to distinguish the pure state environment from the thermal state environment. Consequently, under the long-term evolution, the integral FT may exhibit significant deviations. However, while large deviations are possible, they are typically transient. This transient nature stems from the system eventually reaching equilibrium as the evolution time lengthens. For most times, the total system tends to approach a fixed steady state [13]. Considering the ETH, distinguishing between these steady states with local observations becomes challenging. As a result, the integral FT remains approximately valid in the long-term average.

For any quantity O(t)O(t), its long-time average is defined as follows

O(t)¯:=limT1T0TO(t)𝑑t.\overline{O(t)}:=\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}O(t)dt. (38)

In the generating function, not only does 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) change with time, but the final state of the system also evolves over time

|ρS(t))=t|ρS(0))=(IE|𝒰SE(t)|ρS(0)ρE(0)).|\rho_{S}(t))=\mathcal{E}_{t}|\rho_{S}(0))=(I_{E}|\mathcal{U}_{SE}(t)|\rho_{S}(0)\otimes\rho_{E}(0)). (39)

Combining it with Eqs. 33 and III.1, we have

𝒩FT(t)t¯=[𝒩FT(t)𝒩FTT(t)]t¯\displaystyle\overline{\mathcal{N}_{\text{FT}}(t)\mathcal{E}_{t}}=\overline{[\mathcal{N}_{\text{FT}}(t)-\mathcal{N}^{T^{\prime}}_{\text{FT}}(t)]\circ\mathcal{E}_{t}}
+[𝒥ρE(0)1/2𝒥ρEcan1/2]𝒥ρScan1/2𝒩βT(t)𝒥ρScan1/2t¯\displaystyle+[\mathcal{J}_{\rho_{E}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\overline{\mathcal{N}_{\beta}^{T^{\prime}\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}\circ\mathcal{E}_{t}}
+𝒥ρEcan1/2𝒥ρScan1/2[𝒩βT(t)𝒩β(t)]𝒥ρScan1/2t¯\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\overline{[\mathcal{N}_{\beta}^{T^{\prime}\dagger}(t)-\mathcal{N}_{\beta}^{\dagger}(t)]\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}\circ\mathcal{E}_{t}}
+𝒥ρEcan1/2𝒥ρScan1/2𝒩β(t)𝒥ρScan1/2t¯,\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\overline{\mathcal{N}_{\beta}^{\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}\circ\mathcal{E}_{t}}, (40)

where 𝒩βT(t)\mathcal{N}_{\beta}^{T^{\prime}}(t) is defined in Eq. 87, representing another truncated version of 𝒩β(t)\mathcal{N}_{\beta}(t) based on the limit imposed by information flow speed under the imaginary time evolution. By substituting 𝒩β\mathcal{N}_{\beta} with 𝒩βT\mathcal{N}_{\beta}^{T^{\prime}} in 𝒩FT\mathcal{N}_{\text{FT}}, we obtain 𝒩FTT\mathcal{N}^{T^{\prime}}_{\text{FT}}. The last term of section III.2 is equivalent to 𝒰SE(t)𝒥ρEcan1/2t¯\overline{\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}\circ\mathcal{E}_{t}}, which is also a TP mapping. The terms in the first and third lines of section III.2 represent the error due to truncation, which can be constrained by the Lieb-Robinson bound. Detailed calculations are provided in section C.1. Under long-time averaging, a single unitary evolution yields the dephasing map [14]

𝒰SE(t)¯=a|Πa)(Πa|,\overline{\mathcal{U}^{\dagger}_{SE}(t)}=\sum_{a}|\Pi_{a})(\Pi_{a}|, (41)

where Πa=|EaEa|\Pi_{a}=\ket{E_{a}}\bra{E_{a}} is the energy eigenstate of the total Hamiltonian HH. A unitary evolution 𝒰SE(t)\mathcal{U}^{\dagger}_{SE}(t) together with a dynamical evolution t\mathcal{E}_{t} give the following map

𝒰SE(t)t¯=𝒰SE(t)¯t¯\displaystyle\overline{\mathcal{U}^{\dagger}_{SE}(t)\circ\dots\circ\mathcal{E}_{t}}=\overline{\mathcal{U}^{\dagger}_{SE}(t)}\circ\dots\circ\overline{\mathcal{E}_{t}}
+a,bab|Πab)(Πab|(IE|Πab)(Πab|ρE(0)),\displaystyle+\sum_{\begin{subarray}{c}a,b\\ a\neq b\end{subarray}}|\Pi_{ab})(\Pi_{ab}|\circ\dots\circ(I_{E}|\Pi_{ab})(\Pi_{ab}|\rho_{E}(0)), (42)

where Πab=|EbEa|\Pi_{ab}=\ket{E_{b}}\bra{E_{a}}. Correspondingly, the error caused by the second line in section III.2 can be divided into two parts. The first part is

(ISE|[𝒥ρE(0)1/2𝒥ρEcan1/2]𝒥ρScan1/2𝒩βT(t)¯𝒥ρScan1/2|ρS(t)¯).(I_{SE}|[\mathcal{J}_{\rho_{E}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\overline{\mathcal{N}_{\beta}^{T^{\prime}\dagger}(t)}\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}|\overline{\rho_{S}(t)})^{\dagger}. (43)

We discuss its bound in detail in section C.3. The second part is

a,bab(ρScan[ρE(0)ρEcan]|𝒩IT(iβ/2)|Πab)×\displaystyle\sum_{\begin{subarray}{c}a,b\\ a\neq b\end{subarray}}(\rho_{S}^{\text{can}}\otimes[\rho_{E}(0)-\rho^{\text{can}}_{E}]|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{ab})^{\dagger}\times
(Πab|𝒩IT(iβ/2)𝒥ρScan1/2|TrEΠabIE)\displaystyle(\Pi_{ab}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\text{Tr}_{E}\Pi_{ab}\otimes I_{E})^{\dagger}
×(Πab|ρS(0)ρE(0)).\displaystyle\times(\Pi_{ab}|\rho_{S}(0)\otimes\rho_{E}(0))^{\dagger}. (44)

We discuss its bound in detail in section C.4. Finally, in the large-environment limit, by combining sections III.2, 108 and 113, we have

|δGF¯(i)||GF¯(i)GFT¯(i)|+|GFT¯(i)GFcan,T¯(i)|\displaystyle\lvert\delta\overline{G_{F}}(i)\rvert\leq\lvert\overline{G_{F}}(i)-\overline{G^{T^{\prime}}_{F}}(i)\rvert+\lvert\overline{G^{T^{\prime}}_{F}}(i)-\overline{G^{\text{can},T^{\prime}}_{F}}(i)\rvert
+|GFcan,T¯(i)GFcan¯(i)|=o(1),\displaystyle+\lvert\overline{G^{\text{can},T^{\prime}}_{F}}(i)-\overline{G^{\text{can}}_{F}}(i)\rvert=o(1), (45)

where ”can” represents replacing the initial environment state in 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) with the canonical ensemble and ”TT^{\prime}” denotes the substitution of 𝒩β\mathcal{N}_{\beta} in 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) with 𝒩βT\mathcal{N}^{T^{\prime}}_{\beta}. In conclusion, the first and third terms of section III.2 depends on the locality of the imaginary time evolution component of the map 𝒩β(t)\mathcal{N}_{\beta}(t) and can be constrained by the Lieb-Robinson bound. The second term of section III.2 depends on the indistinguishability of the states within a typical energy shell and can be restricted by ETH.

IV Eigenstate fluctuation theorem for direct heat exchange between two baths

We now consider the interaction of two baths AA and BB. The whole system evolves according to the unitary evolution with the Hamiltonian H=HA+HB+HIH=H_{A}+H_{B}+H_{I}. If the initial state of two baths is the canonical ensemble with different temperatures ρ^Xcan=eβXH^X/ZX\hat{\rho}^{\text{can}}_{X}=e^{-\beta_{X}\hat{H}_{X}}/Z_{X} and X=A,BX=A,B, then we should choose

σ^t=βA(H^AFA)+βB(H^BFB).\hat{\sigma}_{t}=\beta_{A}(\hat{H}_{A}-F_{A})+\beta_{B}(\hat{H}_{B}-F_{B}). (46)

If we do not make any assumptions about the initial environment state, but still set σ^t\hat{\sigma}_{t} as Eq. 46, then we have

GF(λ)=(IAB|𝒥ρAcanρBcaniλ/2𝒰AB(t)\displaystyle G_{F}(\lambda)=(I_{AB}|\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{-i\lambda/2}\circ\mathcal{U}_{AB}(t)
𝒥ρAcanρBcaniλ/2σ^0|ρA(0)ρB(0)).\displaystyle\circ\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}_{0}}|\rho_{A}(0)\otimes\rho_{B}(0)). (47)

Since the two baths can deviate from the canonical ensemble ρcan\rho^{\text{can}}, the integral FT will exhibit deviations. By defining a map like 𝒩β\mathcal{N}_{\beta} but with different temperatures (see section B.4), we can rewrite the conjugated form of section IV as GF(i)=(IAB|𝒩FT(t)|1)G_{F}(i)=(I_{AB}|\mathcal{N}_{\text{FT}}(t)|1)^{\dagger}, where

𝒩FT(t):=𝒥ρA(0)ρB(0)1/2𝒩βA,βB(t).\mathcal{N}_{\text{FT}}(t):=\mathcal{J}_{\rho^{\prime}_{A}(0)\otimes\rho^{\prime}_{B}(0)}^{1/2}\circ\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t). (48)

When the baths ρX(0)\rho_{X}(0) deviate from the canonical ensemble ρXcan\rho^{\text{can}}_{X}, the constructed map 𝒩FT(t)\mathcal{N}_{\text{FT}}(t) is generally not a TP map.

Here we consider the situation where the initial states of AA and BB are energy eigenstates. In this case, we have ρX(0)=ρX(0)\rho^{\prime}_{X}(0)=\rho_{X}(0). If the evolution time tends to zero, it is obvious that 𝒩FT(t)=𝒥ρA(0)ρB(0)1/2\mathcal{N}_{\text{FT}}(t)=\mathcal{J}_{\rho^{\prime}_{A}(0)\otimes\rho^{\prime}_{B}(0)}^{1/2}. It acts on |1)|1) gives |ρA(0)ρB(0))|\rho^{\prime}_{A}(0)\otimes\rho^{\prime}_{B}(0)). Therefore, it is also a TP assignment map. For finite time evolution, similar to section III.1, we can rewrite the constructed map as

𝒩FT(t)=𝒥ρA(0)ρB(0)1/2[𝒩βA,βB(t)𝒩βA,βBT(t)]\displaystyle\mathcal{N}_{\text{FT}}(t)=\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ[\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t)-\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)]
+[𝒥ρA(0)ρB(0)1/2𝒥ρAcanρBcan1/2]𝒩βA,βBT(t)\displaystyle+[\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}]\circ\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)
+𝒥ρAcanρBcan1/2[𝒩βA,βBT(t)𝒩βA,βB(t)]\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}\circ[\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)-\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t)]
+𝒥ρAcanρBcan1/2𝒩βA,βB(t),\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}\circ\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t), (49)

where the truncated 𝒩βA,βBT(t)\mathcal{N}^{T}_{\beta_{A},\beta_{B}}(t) is defined in section B.4. It is a truncated version of 𝒩βA,βB(t)\mathcal{N}_{\beta_{A},\beta_{B}}(t) based on the speed limit of information flow. The last term of section IV is equal to 𝒰SE(t)𝒥ρAcanρBcan1/2\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}. It acts on |1)|1) gives 𝒰SE(t)|ρAcanρBcan)\mathcal{U}^{\dagger}_{SE}(t)|\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}). So it’s a TP mapping. The terms in the first and third lines of section IV represent the error due to truncation, which can also be bounded by the Lieb-Robinson bound. We discuss these bounds in detail in section C.1. The term in the second line of section IV represents the error caused by the deviation of the initial state. Since the mapping 𝒩βA,βBT(t)\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t) acts on |1)|1) also gives a local operator, this part of the error can be bounded by ETH. See section C.2 for details. Finally, in the large-environment limit, by combining the error terms in section IV, we can obtain a similar bound as section III.1. Therefore, the error of FT should vanish in the thermodynamic limit.

If both heat baths have the same inverse temperature as given by Eq. 35, we can still utilize the perturbation expansion in section B.3. Since the final state measurement here is independent of time, we only need to consider

𝒩FT(t)¯=𝒥ρA(0)ρB(0)1/2[𝒩β(t)𝒩βT(t)]¯\displaystyle\overline{\mathcal{N}_{\text{FT}}(t)}=\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ\overline{[\mathcal{N}^{\dagger}_{\beta}(t)-\mathcal{N}^{T^{\prime}\dagger}_{\beta}(t)]}
+[𝒥ρA(0)ρB(0)1/2𝒥ρAcanρBcan1/2]𝒩βT(t)¯\displaystyle+[\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}]\circ\overline{\mathcal{N}^{T^{\prime}\dagger}_{\beta}(t)}
+𝒥ρAcanρBcan1/2[𝒩βT(t)𝒩β(t)]¯\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}\circ\overline{[\mathcal{N}^{T^{\prime}\dagger}_{\beta}(t)-\mathcal{N}^{\dagger}_{\beta}(t)]}
+𝒥ρAcanρBcan1/2𝒩β(t)¯,\displaystyle+\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}\circ\overline{\mathcal{N}^{\dagger}_{\beta}(t)}, (50)

where 𝒩βT(t)\mathcal{N}_{\beta}^{T^{\prime}}(t) is defined by Eq. 87, except that SESE is replaced by ABAB and truncate both AA and BB. The last term of section IV is equal to 𝒰AB(t)¯𝒥ρAcanρBcan1/2\overline{\mathcal{U}^{\dagger}_{AB}(t)}\circ\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}, which is also a TP mapping. The terms in the first and third lines of section IV represent the error due to truncation, which can be bounded by the Lieb-Robinson bound. The detailed calculation can be found in section C.1. The term in the second line of section IV represents the error caused by the deviation of the initial state. This part of the error can be bounded with ETH. See section C.2 for details. Finally, in the large-environment limit, by combining the error terms in section IV, we can obtain a similar bound as section III.2. Therefore, the error of FT in the long-time average should vanish in the thermodynamic limit.

If the temperature of the two baths is different, the corresponding term ρAcanρBcan\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}differs significantly from the imaginary time evolution USE(it)U_{SE}(it), rendering the perturbation expansion in section B.3 inapplicable in this scenario. Additionally, considering the evolution perspective, since both AA and BB are large baths, their equilibration time may be considerably long, and they may not reach a steady state quickly. Consequently, it is unlikely that the integral fluctuation theorem holds for a long-time average in this case.

V Integral fluctuation theorem from quasi-measurements

From Eq. 27, we observe that, in general, the entropy production in the FTs of open systems is closely linked to the state of the environment. However, when the evolution map possesses a global fixed point, the entropy flux can be expressed solely in terms of system-related quantities [9]. Now, let us briefly consider these cases using the method described herein.

For systems with global fixed point

𝒰SE(ρSρE)=ρSρE,\mathcal{U}_{SE}(\rho_{S}^{*}\otimes\rho_{E})=\rho^{*}_{S}\otimes\rho_{E}, (51)

we have

𝒥ρSρE1/2𝒰SE(t)𝒥ρSρE1/2=𝒰SE(t).\mathcal{J}^{1/2}_{\rho^{*}_{S}\otimes\rho_{E}}\circ\mathcal{U}_{SE}(t)\circ\mathcal{J}^{-1/2}_{\rho^{*}_{S}\otimes\rho_{E}}=\mathcal{U}_{SE}(t). (52)

Supposing there is no correlation between the initial system and environment, if the environment is not measured, substituting Eq. 52 into Eq. 18, we can get

GF(λ)=(ISρE|𝒥eσ^tSiλ/2𝒥ρS1/2𝒰(t)\displaystyle G_{F}(\lambda)=(I_{S}\otimes\rho_{E}|\mathcal{J}_{e^{-\hat{\sigma}^{S}_{t}}}^{-i\lambda/2}\circ\mathcal{J}_{\rho^{*}_{S}}^{1/2}\circ\mathcal{U}(t)
𝒥ρS1/2𝒥eσ^0Siλ/2σ^0S|ρS(0)IE).\displaystyle\circ\mathcal{J}_{\rho^{*}_{S}}^{-1/2}\circ\mathcal{J}_{e^{-\hat{\sigma}^{S}_{0}}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}^{S}_{0}}|\rho_{S}(0)\otimes I_{E}). (53)

However, it is not possible to choose a suitable σ^S(t)\hat{\sigma}_{S}(t) such that GF(i)=1G_{F}(i)=1. The two-point measurement of the local system cannot satisfy the integral FT. Previous work [8] established a general Crooks FT for open quantum processes, where only local measurements of the system are required, but these measurements are no longer the conventional two-point measurements. To obtain the so-called quasiprobability, quasi-measurements need to be introduced [15]. This is a measure of the density matrix and can be reconstructed from positive-operator valued measurements. After incorporating quasi-measurements, the generating function can be expressed as

GF(λ):=σ0,σt,ij,kl(eiλ(σ^t+τ^σ^0τ^)|\displaystyle G_{F}(\lambda):=\sum_{\sigma_{0},\sigma_{t},ij,k^{\prime}l^{\prime}}(e^{i\lambda(\hat{\sigma}_{t}+\hat{\tau}^{\prime}-\hat{\sigma}_{0}-\hat{\tau}})|
ΠσtΠσ0ΠijΠkl)(O(σt,σ0,ij,kl)|𝒮),\displaystyle\Pi_{\sigma_{t}}\otimes\Pi_{\sigma_{0}}\otimes\Pi_{ij}\otimes\Pi_{k^{\prime}l^{\prime}})(O^{(\sigma_{t},\sigma_{0},ij,k^{\prime}l^{\prime})}|\mathcal{S}), (54)

where {|i}\{\ket{i}\} is the diagonal basis of τ^\hat{\tau} and {|k}\{\ket{k^{\prime}}\} is the diagonal basis of τ^\hat{\tau}^{\prime}. The process state 𝒮\mathcal{S} is the Choi-state form of the process tensors and O(σt,σ0,ij,kl)O^{(\sigma_{t},\sigma_{0},ij,k^{\prime}l^{\prime})} is the Choi-state of the measurement. For specific definitions, refer to [15]. Unlike Eq. 17, according to the completeness relation, we have

ij(eiλτ^|Πij)|Πij)(Πij|=𝒥eτ^iλ/2.\sum_{ij}(e^{-i\lambda\hat{\tau}}|\Pi_{ij})|\Pi_{ij})(\Pi_{ij}|=\mathcal{J}_{e^{-\hat{\tau}}}^{i\lambda/2}. (55)

Similar to the procedure used in Eq. 18, we can rewrite the generating function as

GF(λ)=(IS|𝒥eσ^tSiλ/2𝒥eτiλ/2𝒩(t)\displaystyle G_{F}(\lambda)=(I_{S}|\mathcal{J}_{e^{-\hat{\sigma}^{S}_{t}}}^{-i\lambda/2}\circ\mathcal{J}_{e^{-\tau^{\prime}}}^{-i\lambda/2}\circ\mathcal{N}(t)
𝒥eτiλ/2𝒥eσ^0Siλ/2σ^0S|ρS(0)),\displaystyle\circ\mathcal{J}_{e^{-\tau}}^{i\lambda/2}\circ\mathcal{J}_{e^{-\hat{\sigma}^{S}_{0}}}^{i\lambda/2}\circ\mathcal{M}_{\hat{\sigma}^{S}_{0}}|\rho_{S}(0)), (56)

where 𝒩(t)=(IE|𝒰(t)|ρE)\mathcal{N}(t)=(I^{E}|\mathcal{U}(t)|\rho_{E}). We can set σ^t=lnρ^S(t)\hat{\sigma}_{t}=-\ln\hat{\rho}_{S}(t) and τ=τ=lnρS\tau=\tau^{\prime}=\ln\rho^{*}_{S}, then the conjugated form of generating function gives

GF(iλ)=(IS|𝒥eσ^0Siλ/2𝒥eτiλ/2𝒥ρS1/2𝒩(t)\displaystyle G_{F}(i-\lambda)=(I_{S}|\mathcal{J}_{e^{-\hat{\sigma}^{S}_{0}}}^{-i\lambda/2}\circ\mathcal{J}_{e^{-\tau}}^{-i\lambda/2}\circ\mathcal{J}_{\rho^{*}_{S}}^{1/2}\circ\mathcal{N}^{\dagger}(t)
𝒥ρS1/2𝒥eτiλ/2𝒥eσ^tSiλ/2|ρS(t)).\displaystyle\circ\mathcal{J}_{\rho^{*}_{S}}^{-1/2}\circ\mathcal{J}_{e^{-\tau^{\prime}}}^{i\lambda/2}\circ\mathcal{J}_{e^{-\hat{\sigma}^{S}_{t}}}^{i\lambda/2}|\rho_{S}(t))^{\dagger}. (57)

From this formula, we can naturally obtain a backward process map

(t)=𝒥ρS1/2𝒩(t)𝒥𝒩(ρS)1/2,\mathcal{R}(t)=\mathcal{J}_{\rho^{*}_{S}}^{1/2}\circ\mathcal{N}^{\dagger}(t)\circ\mathcal{J}_{\mathcal{N}(\rho^{*}_{S})}^{-1/2}, (58)

which is just the Petz recovery map used in [5, 7, 8]. From section V, we know that the integral FT is guaranteed by the TP property of the Petz recovery map

GF(i)=(I|(t)|ρS(t))=1.G_{F}(i)=(I|\mathcal{R}(t)|\rho_{S}(t))^{\dagger}=1. (59)

Different from Eq. 25, the entropy production here depends only on the local state of the system and does not need to know the state of the environment.

VI Conclusion and discussion

In this paper, we have redefined the generating function of the general FT under unitary evolution and obtained a constructed map. The initial state, measurements, their Fourier transform, and evolution are all encompassed in this map. The constructed map is CP, and its TP property determines the integral FT. This unified form can help simplify the impact of initial state deviation on the FT. In particular, for the measurement of energy H^\hat{H}, the corresponding rescaling map is imaginary time evolution. The commutativity between imaginary time evolution and real-time evolution can effectively simplify calculations. With the help of this new formalism, we have proved the eigenstate FT in a simpler manner that requires fewer and weaker assumptions. The rationale behind these proofs is roughly as follows: The error of the FT can be divided into two components. One arises from the non-locality of the map 𝒩β(t)\mathcal{N}_{\beta}(t) and can be bounded by the Lieb-Robinson bound. The other originates from the distinguishability of environment states and can be restricted by ETH. Both errors vanish in the thermodynamic limit, ensuring the FT’s validity even when the initial state of the bath is a single energy eigenstate of a many-body systems.

We also explored the heat exchange between two baths. The initial state of both baths is a pure state. We observed that the integral FT remains approximately valid in the short-time regime. In the long-time regime, when the temperatures of the two baths are equal, the long-time average of the integral FT holds approximately. However, when the temperatures differ, the long-time average of the integral FT may not hold.

Referring to the FTs for the quantum channel, we have examined the generating function of the quasi-probability. The Petz recovery map can be naturally derived from this generating function. The integral FT is ensured by the TP property of the Petz recovery map.

An additional advantage of our method is its ease of generalization to multipoint measurements. The operator-state formalism employed here seamlessly integrates with process tensors [15]. Process tensors offer the convenience of studying multitime processes and multipoint measurements [16, 17]. FTs beyond two-point measurements have garnered extensive attention [19, 20, 21, 18, 22, 23]. The integral FT under multitime processes can aid in exploring the generalization of the fluctuation-dissipation theorem for n-point systems [24]. Since the memory effect can result in a negative entropy production rate [15], research in this area also contributes to a deeper understanding of the second law of thermodynamics: The second law implies complete irreversibility, while the Poincaré recurrence theorem corresponds to complete reversibility. FTs enable us to derive irreversibility from reversible evolution, while the memory effect can weaken irreversibility and yield negative entropy production rates. Investigating the influence of the ETH on FTs under multitime evolution allows us to further comprehend the impact and limitations of the memory effect. This, in turn, aids in understanding the role of the thermodynamic limit in the second law of thermodynamics and the Poincaré recurrence theorem.

Acknowledgements.
Z.H. is supported by the National Natural Science Foundation of China under Grants No. 12305035. We also thank X.-K. Guo for helpful comments.

Appendix A The crucial lemmas and hypothesis used from the literature

The lemma 3.1 in [25] tells us that when 0s<1gk0\leq s<\frac{1}{gk}, we have

esHAesHA×(1gks)R/gk,\lVert e^{sH}Ae^{-sH}\rVert\leq\lVert A\rVert\times(1-gks)^{-R/gk}, (60)

where HH is a local Hamiltonian, and AA is an arbitrary local operator. The parameters gg, kk, and RR are related to the lattice structure and interaction between particles and are independent of the system size. According to this lemma, as long as the system’s Hamiltonian meets the corresponding requirements, when τ\tau is sufficiently small, the term in Eq. 71 can be bounded as

HI(iτ)=Θ(1)×HI.\lVert H_{I}(i\tau)\rVert=\Theta(1)\times\lVert H_{I}\rVert. (61)

The lemma 5 in [11] gives the following Lieb-Robinson bound

eitHOXeitHeitHTOXeitHT|X|OX(2ζ0|t|)!,\lVert e^{itH}O_{X}e^{-itH}-e^{itH^{T}}O_{X}e^{-itH^{T}}\rVert\leq\lvert X\rvert\lVert O_{X}\rVert\frac{(2\zeta_{0}\lvert t\rvert)^{\ell}}{\ell!}, (62)

where HH is a local Hamiltonian, HTH^{T} is the truncated Hamiltonian, and OXO_{X} is any local operator. The parameter \ell represents the distance between OXO_{X} and the truncation boundary, and |X|\lvert X\rvert denotes the number of sites in the support XX. The constant ζ0=O(1)\zeta_{0}=O(1) depends on the lattice structure and interactions among lattice sites. According to this lemma, the error induced by truncation will be exponentially small

δHI(τ)=|HI|×HI×O(eμ)\lVert\delta H_{I}(\tau)\rVert=\lvert H_{I}\rvert\times\lVert H_{I}\rVert\times O(e^{-\mu\ell}) (63)

as long as the evolution time |τ|/vLR\lvert\tau\rvert\leq\ell/v_{LR}, where vLRv_{LR} is so-called Lieb-Robinson velocity.

According to the canonical typicality [26], weak ETH [2, 27] or subsystem ETH [28], we know that it is difficult for local operators to distinguish the global energy eigenstates from the canonical ensembles

|(Πa|O)(ρcan|O)|O×Nα,\displaystyle\lvert(\Pi_{a}|O)-(\rho^{\text{can}}|O)\rvert\leq\lVert O\rVert\times N^{-\alpha}, (64)
|(Πab|O)|O×Nα(ab),\displaystyle\lvert(\Pi_{ab}|O)\rvert\leq\lVert O\rVert\times N^{-\alpha}\quad(a\neq b), (65)

The parameter 0<α0<\alpha. NN denotes the size (the number of sites) of the entire system. The size of the local operator OO should be much smaller than NN. It is worth noting that in Eq. 64, the errors decay algebraically, which is weaker than the exponential decay usually expected in ETH. Therefore, it can be anticipated that this hypothesis can be satisfied in a broader range of systems.

Appendix B The properties of 𝒩β\mathcal{N}_{\beta}

B.1 Imaginary time evolution and strict energy conservation condition

The superoperator 𝒩β\mathcal{N}_{\beta} of Eq. 32 can also be written as

𝒩β(t)()=𝒰SE0(βλ/2)𝒰SE(t)𝒰SE0(βλ/2)()\displaystyle\mathcal{N}_{\beta}(t)(\cdot)=\mathcal{U}_{SE}^{0}(-\beta\lambda/2)\circ\mathcal{U}_{SE}(t)\circ\mathcal{U}_{SE}^{0}(\beta\lambda/2)(\cdot)
=λ(t)()λ(t),\displaystyle=\mathcal{M}_{\lambda}(t)(\cdot)\mathcal{M}_{\lambda}^{\dagger}(t), (66)

where

λ(t)=USE0(βλ/2)[USE(t)]USE0(βλ/2)\mathcal{M}_{\lambda}(t)=U^{0}_{SE}(-\beta\lambda/2)[U_{SE}(t)]U^{0}_{SE}(\beta\lambda/2) (67)

and

USE0(τ)=exp(i(HS+HE)τ).U^{0}_{SE}(\tau)=\exp(-i(H_{S}+H_{E})\tau). (68)

When λ\lambda is a pure imaginary number, the mapping 𝒰SE0(βλ/2)\mathcal{U}_{SE}^{0}(-\beta\lambda/2) is just the imaginary time evolution [29]. Unlike the real time evolution, imaginary time evolution is CP but not TP. From Eqs. 6 and II.1, we have

𝒰SE0(τ)=𝒰SE0(τ).\mathcal{U}^{0\dagger}_{SE}(\tau)=\mathcal{U}^{0}_{SE}(-\tau^{*}). (69)

Therefore, the imaginary time evolution is self-conjugate. Since USE(t)=ei(H0+HI)tU_{SE}(t)=e^{-i(H_{0}+H_{I})t} and [USE0,H0]=0[U^{0}_{SE},H_{0}]=0, we can rewrite Eq. 67 as

λ(t)=ei[H0+HI(βλ/2)]t,\mathcal{M}_{\lambda}(t)=e^{-i[H_{0}+H_{I}(\beta\lambda/2)]t}, (70)

where

HI(τ):=USE0(τ)HIUSE0(τ).H_{I}(\tau):=U^{0}_{SE}(-\tau)H_{I}U^{0}_{SE}(\tau). (71)

Notice that it is different from the Hermitian operator

𝒰SE0(τ)HI=USE0(τ)HIUSE0(τ).\mathcal{U}^{0}_{SE}(-\tau)H_{I}=U^{0}_{SE}(-\tau)H_{I}U^{0\dagger}_{SE}(-\tau). (72)

HI(βλ/2)H_{I}(\beta\lambda/2) is a pseudo-Hermitian operator [30]. It is easy to verify that with ρ=e2H0Im(τ)\rho=e^{2H_{0}\text{Im}(\tau)}, we have ρHI(τ)ρ1=HI(τ)\rho H_{I}(\tau)\rho^{-1}=H^{\dagger}_{I}(\tau). According to Eq. 70, if there is a strict energy conservation condition [H0,HI]=0[H_{0},H_{I}]=0 [31], or the temperature tends to infinity, there will be λ(t)=USE(t)\mathcal{M}_{\lambda}(t)=U_{SE}(t). And then we have 𝒩β(t)=𝒰SE(t)\mathcal{N}_{\beta}(t)=\mathcal{U}_{SE}(t).

B.2 Short-time evolution and truncated superoperator

The evolution map can be written in the integral form

𝒰SE(t)=𝒰SE(t)𝒰SE0(0)\displaystyle\mathcal{U}_{SE}(t)=\mathcal{U}_{SE}(t)\circ\mathcal{U}^{0}_{SE}(0)
=𝒰SE0(t)+0t𝑑ττ𝒰SE(τ)𝒰SE0(tτ)\displaystyle=\mathcal{U}^{0}_{SE}(t)+\int_{0}^{t}d\tau\frac{\partial}{\partial_{\tau}}\mathcal{U}_{SE}(\tau)\circ\mathcal{U}^{0}_{SE}(t-\tau)
=𝒰SE0(t)+0t𝑑τ𝒰SE(τ)I𝒰SE0(tτ)\displaystyle=\mathcal{U}^{0}_{SE}(t)+\int_{0}^{t}d\tau\mathcal{U}_{SE}(\tau)\circ\mathcal{L}_{I}\circ\mathcal{U}^{0}_{SE}(t-\tau) (73)

where I():=i[HI,]\mathcal{L}_{I}(\cdot):=-i[H_{I},\cdot]. Here we only take the first-order perturbation approximation

𝒰SE(t)𝒰SE0(t)[+0t𝑑τ𝒰SE0(τt)I𝒰SE0(tτ)].\mathcal{U}_{SE}(t)\simeq\mathcal{U}^{0}_{SE}(t)[\mathcal{I}+\int_{0}^{t}d\tau\mathcal{U}^{0}_{SE}(\tau-t)\circ\mathcal{L}_{I}\circ\mathcal{U}^{0}_{SE}(t-\tau)]. (74)

The error caused by taking the first-order approximation

δ𝒰SE(t)=0t𝑑τ[𝒰SE(τ)𝒰SE0(τ)]I𝒰SE0(tτ).\delta\mathcal{U}_{SE}(t)=\int_{0}^{t}d\tau[\mathcal{U}_{SE}(\tau)-\mathcal{U}^{0}_{SE}(\tau)]\circ\mathcal{L}_{I}\circ\mathcal{U}^{0}_{SE}(t-\tau). (75)

Using integral form section B.2, the error can be rewritten as

0t𝑑τ10τ1𝑑τ2𝒰SE(τ2)I𝒰SE0(τ1τ2)I𝒰SE0(tτ1).\int_{0}^{t}d\tau_{1}\int_{0}^{\tau_{1}}d\tau_{2}\mathcal{U}_{SE}(\tau_{2})\circ\mathcal{L}_{I}\circ\mathcal{U}^{0}_{SE}(\tau_{1}-\tau_{2})\circ\mathcal{L}_{I}\circ\mathcal{U}^{0}_{SE}(t-\tau_{1}). (76)

After it acts on arbitrary operator OO, the bound of error can be calculated with

δ𝒰SE(t)O0tdτ10τ1dτ2𝒰SE(τ2)𝒰SE0(τ2)\displaystyle\lVert\delta\mathcal{U}_{SE}(t)O\rVert\leq\int_{0}^{t}d\tau_{1}\int_{0}^{\tau_{1}}d\tau_{2}\lVert\mathcal{U}_{SE}(\tau_{2})\circ\mathcal{U}^{0}_{SE}(-\tau_{2})
I(τ2)I(τ1)𝒰SE0(t)O(2tHI)22O,\displaystyle\circ\mathcal{L}_{I}(-\tau_{2})\circ\mathcal{L}_{I}(-\tau_{1})\circ\mathcal{U}^{0}_{SE}(t)O\rVert\leq\frac{(2t\lVert H_{I}\rVert)^{2}}{2}\lVert O\rVert, (77)

where I(τ)O=i[HI(τ)OOHI(τ)]\mathcal{L}_{I}(\tau)O=-i\left[H_{I}(\tau)O-OH^{\dagger}_{I}(\tau)\right]. The error is small when the evolution time is short enough t<<1/(2HI)t<<1/(2\lVert H_{I}\rVert). Using the approximation Eq. 74, it is easy to show that

𝒩β(t)(OSIE)=𝒥ρScanρEcaniλ/2𝒰SE(t)𝒥ρScanρEcaniλ/2(OSIE)\displaystyle\mathcal{N}^{\dagger}_{\beta}(t)(O_{S}\otimes I_{E})=\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}\circ\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}(O_{S}\otimes I_{E})
𝒰SE0(t)(OSIE)+0t𝑑τ𝒰SE0(tτ)𝒥ρScanρEcaniλ/2I𝒥ρScanρEcaniλ/2𝒰SE0(τt)(𝒰SE0(t)OSIE)\displaystyle\simeq\mathcal{U}^{0\dagger}_{SE}(t)(O_{S}\otimes I_{E})+\int_{0}^{t}d\tau\mathcal{U}^{0\dagger}_{SE}(t-\tau)\circ\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}\circ\mathcal{L}^{\dagger}_{I}\circ\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{E}}\circ\mathcal{U}^{0\dagger}_{SE}(\tau-t)(\mathcal{U}^{0\dagger}_{SE}(t)O_{S}\otimes I_{E})
=𝒰S0(t)(OSIE)+0t𝑑τ𝒰SE0(tτ)𝒰SE0(βλ/2)I𝒰SE0(βλ/2)𝒰SE0(τt)(𝒰S0(t)OSIE)\displaystyle=\mathcal{U}^{0\dagger}_{S}(t)(O_{S}\otimes I_{E})+\int_{0}^{t}d\tau\mathcal{U}^{0\dagger}_{SE}(t-\tau)\circ\mathcal{U}_{SE}^{0\dagger}(-\beta\lambda/2)\circ\mathcal{L}^{\dagger}_{I}\circ\mathcal{U}_{SE}^{0\dagger}(\beta\lambda/2)\circ\mathcal{U}^{0\dagger}_{SE}(\tau-t)(\mathcal{U}^{0\dagger}_{S}(t)O_{S}\otimes I_{E})
=𝒰S0(t)(OSIE)+0t𝑑τI(tτβλ/2)(𝒰S0(t)OSIE)\displaystyle=\mathcal{U}^{0\dagger}_{S}(t)(O_{S}\otimes I_{E})+\int_{0}^{t}d\tau\mathcal{L}^{\dagger}_{I}(t-\tau-\beta\lambda/2)(\mathcal{U}^{0\dagger}_{S}(t)O_{S}\otimes I_{E}) (78)

where I(τ)O=i[HI(τ)OOHI(τ)]\mathcal{L}^{\dagger}_{I}(\tau)O=i\left[H_{I}(\tau)O-OH^{\dagger}_{I}(\tau)\right]. The first term on the right side of the last equation of (B.2) only contains the local evolution of the system, and its contribution to GF(λ)G_{F}(\lambda) is

Tr[𝒥ρS(t)iλ/2𝒥ρScaniλ/2𝒰S0(t)𝒥ρScaniλ/2𝒥ρS(0)iλ/2(ρS(0)ρE(0))]=Tr[𝒥ρS(t)iλ/2𝒰S0(t)𝒥ρS(0)iλ/2(ρS(0))]=λ=i1.\text{Tr}[\mathcal{J}_{\rho_{S}(t)}^{-i\lambda/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{i\lambda/2}\circ\mathcal{U}^{0}_{S}(t)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-i\lambda/2}\circ\mathcal{J}_{\rho_{S}(0)}^{i\lambda/2}(\rho_{S}(0)\otimes\rho^{\prime}_{E}(0))]=\text{Tr}[\mathcal{J}_{\rho_{S}(t)}^{-i\lambda/2}\circ\mathcal{U}^{0}_{S}(t)\circ\mathcal{J}_{\rho_{S}(0)}^{i\lambda/2}(\rho_{S}(0))]\overset{\lambda=i}{=}1. (79)

Now, let’s analyze the contribution of the second term. We divide the environment into two parts: B0B_{0}, which is the portion close to the system, and B0¯\overline{B_{0}}, which comprises the remaining parts. When the temperature is high and the evolution time is short, the scale of B0B_{0} can be larger than the propagation range of the Lieb-Robinson velocity vLR(|t+βλ/2|)v_{LR}(\lvert t+\beta\lambda/2\rvert), but much smaller than the overall environment scale LEL_{E}. The Hamiltonian of the environment can also be divided as HE=HB0+HB0¯+HB0B0¯H_{E}=H_{B_{0}}+H_{\overline{B_{0}}}+H_{B_{0}\overline{B_{0}}}. We define the truncated Hamiltonian H0T:=HS+HB0H^{T}_{0}:=H_{S}+H_{B_{0}}, the corresponding time evolution operator USET=ei(H0,T+HI)tU^{T}_{SE}=e^{-i(H_{0,T}+H_{I})t}, the truncated canonical ensemble ρB0can=eβHB0/ZB0\rho^{\text{can}}_{B_{0}}=e^{-\beta H_{B_{0}}}/Z_{B_{0}}, and the corresponding superoperator 𝒩βT(t):=𝒥ρScanρB0caniλ/2𝒰SET(t)𝒥ρScanρB0caniλ/2\mathcal{N}^{T}_{\beta}(t):=\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{B_{0}}}\circ\mathcal{U}^{T}_{SE}(t)\circ\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{S}\otimes\rho^{\text{can}}_{B_{0}}}. Following a similar procedure to that in sections B.2 and B.2, we have:

𝒩βT(t)(OSIE)𝒰S0(t)(OSIE)+0t𝑑τIT(tτβλ/2)(𝒰S0(t)OSIE),\mathcal{N}^{T\dagger}_{\beta}(t)(O_{S}\otimes I_{E})\simeq\mathcal{U}^{0\dagger}_{S}(t)(O_{S}\otimes I_{E})+\int_{0}^{t}d\tau\mathcal{L}^{T\dagger}_{I}(t-\tau-\beta\lambda/2)(\mathcal{U}^{0\dagger}_{S}(t)O_{S}\otimes I_{E}), (80)

where IT(τ)O=i[HIT(τ)OOHIT(τ)]\mathcal{L}^{T\dagger}_{I}(\tau)O=i\left[H^{T}_{I}(\tau)O-OH^{T{\dagger}}_{I}(\tau)\right]. HIT(τ):=eiH0TτHIeiH0TτH^{T}_{I}(\tau):=e^{iH^{T}_{0}\tau}H_{I}e^{-iH^{T}_{0}\tau} is also a pseudo-Hermitian operator.

B.3 Long-time evolution and truncated superoperator

When the evolution time is short, employing perturbation theory for the time evolution 𝒰SE(t)\mathcal{U}_{SE}(t) is suitable. However, this method may not be appropriate for long evolution times. An alternative approach, as demonstrated in [3], involves utilizing perturbation theory for the eigenstates of HH. Here, we explore a different strategy. Notably, when the Hamiltonian remains constant, the imaginary-time evolution and the real-time evolution commute. Therefore, the superoperator 𝒩β(t)\mathcal{N}_{\beta}(t) can be expressed as

𝒩β(t)=𝒩I(βλ/2)𝒰SE(t)𝒩I(βλ/2),\mathcal{N}_{\beta}(t)=\mathcal{N}_{I}(-\beta\lambda/2)\circ\mathcal{U}_{SE}(t)\circ\mathcal{N}^{\dagger}_{I}(\beta\lambda/2), (81)

where 𝒩I(βλ/2):=𝒰SE0(βλ/2)𝒰SE(βλ/2)\mathcal{N}_{I}(-\beta\lambda/2):=\mathcal{U}_{SE}^{0}(-\beta\lambda/2)\circ\mathcal{U}_{SE}(\beta\lambda/2). The imaginary time evolution can also be written in the integral form

𝒰SE(it)=𝒰SE0(0)𝒰SE(it)=𝒰SE0(it)+0t𝑑ττ𝒰SE0(itiτ)𝒰SE(iτ)=𝒰SE0(it)[+0t𝑑τ𝒰SE0(iτ)IA𝒰SE(iτ)],\mathcal{U}_{SE}(it)=\mathcal{U}^{0}_{SE}(0)\circ\mathcal{U}_{SE}(it)=\mathcal{U}^{0}_{SE}(it)+\int_{0}^{t}d\tau\frac{\partial}{\partial_{\tau}}\mathcal{U}^{0}_{SE}(it-i\tau)\circ\mathcal{U}_{SE}(i\tau)=\mathcal{U}^{0}_{SE}(it)\left[\mathcal{I}+\int_{0}^{t}d\tau\mathcal{U}^{0}_{SE}(-i\tau)\circ\mathcal{L}^{\text{A}}_{I}\circ\mathcal{U}_{SE}(i\tau)\right], (82)

where IA():={HI,}\mathcal{L}^{\text{A}}_{I}(\cdot):=\{H_{I},\cdot\}. Similar to Eq. 76, the error caused by taking the first-order approximation

δ𝒰SE(it)=0t𝑑τ10τ1𝑑τ2𝒰SE0(itiτ1)IA𝒰SE0(iτ1iτ2)IA𝒰SE(iτ2).\delta\mathcal{U}_{SE}(it)=\int_{0}^{t}d\tau_{1}\int_{0}^{\tau_{1}}d\tau_{2}\mathcal{U}^{0}_{SE}(it-i\tau_{1})\circ\mathcal{L}^{A}_{I}\circ\mathcal{U}^{0}_{SE}(i\tau_{1}-i\tau_{2})\circ\mathcal{L}^{A}_{I}\circ\mathcal{U}_{SE}(i\tau_{2}). (83)

After it acts an arbitrary operator OO, the bound of error can be calculated with

𝒰SE0(it)δ𝒰SE(it)O0t𝑑τ10τ1𝑑τ2IA(iτ1)IA(iτ2)𝒰SE0(iτ2)𝒰SE(iτ2)O\displaystyle\lVert\mathcal{U}^{0}_{SE}(-it)\circ\delta\mathcal{U}_{SE}(it)O\rVert\leq\int_{0}^{t}d\tau_{1}\int_{0}^{\tau_{1}}d\tau_{2}\lVert\mathcal{L}^{A}_{I}(i\tau_{1})\circ\mathcal{L}^{A}_{I}(i\tau_{2})\circ\mathcal{U}^{0}_{SE}(-i\tau_{2})\circ\mathcal{U}_{SE}(i\tau_{2})O\rVert
(2tHI(iτmax))22𝒩I(iτmax)O\displaystyle\leq\frac{(2t\lVert H_{I}(i\tau_{\max})\rVert)^{2}}{2}\lVert\mathcal{N}_{I}(-i\tau^{\prime}_{\max})O\rVert (84)

where IA(iτ)O=HI(iτ)O+OHI(iτ)\mathcal{L}^{\text{A}}_{I}(i\tau)O=H_{I}(i\tau)O+OH^{\dagger}_{I}(i\tau). The τmax[0,β/2]\tau_{\max}\in[0,\beta/2] makes HI(iτmax)\lVert H_{I}(i\tau_{\max})\rVert maximize and τmax[0,β/2]\tau^{\prime}_{\max}\in[0,\beta/2] makes 𝒩I(iτmax)O\lVert\mathcal{N}_{I}(-i\tau^{\prime}_{\max})O\rVert maximize. According to Eq. 61, the error is small when the temperature is high enough β<<1Θ(1)HI\beta<<\frac{1}{\Theta(1)\lVert H_{I}\rVert}. Applying the first-order perturbation approximation to the imaginary time evolution, we obtain

𝒩I(it)+0t𝑑τIA(iτ).\mathcal{N}_{I}(-it)\simeq\mathcal{I}+\int_{0}^{t}d\tau\mathcal{L}^{\text{A}}_{I}(i\tau). (85)

Similar to the procedure in section B.2, we can define the truncated superoperator 𝒰SE0,T\mathcal{U}_{SE}^{0,T} and 𝒩IT(βλ/2)\mathcal{N}^{T}_{I}(-\beta\lambda/2). Employing a similar approximation as in (85), we have

𝒩IT(it):=𝒰SE0,T(it)𝒰SET(it)+0t𝑑τIA,T(iτ).\mathcal{N}^{T}_{I}(-it):=\mathcal{U}_{SE}^{0,T}(-it)\circ\mathcal{U}^{T}_{SE}(it)\simeq\mathcal{I}+\int_{0}^{t}d\tau\mathcal{L}^{\text{A},T}_{I}(i\tau). (86)

Accordingly, we can define a truncated map

𝒩βT(t):=𝒩IT(βλ/2)𝒰SE(t)𝒩IT(βλ/2).\displaystyle\mathcal{N}^{T^{\prime}}_{\beta}(t):=\mathcal{N}^{T}_{I}(-\beta\lambda/2)\circ\mathcal{U}_{SE}(t)\circ\mathcal{N}^{T\dagger}_{I}(\beta\lambda/2). (87)

B.4 𝒩β\mathcal{N}_{\beta} with different temperature

In section B.1, we assume that both parts have the same temperature. However, when the temperatures of the two parts differ, the expression for 𝒥ρAcanρBcan\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}} in terms of 𝒰AB0\mathcal{U}_{AB}^{0} is no longer applicable. Instead, we can use the following mapping

𝒩βA,βB(t)():=𝒥ρAcanρBcaniλ/2𝒰AB(t)𝒥ρAcanρBcaniλ/2()=𝒰AB0(βA,βB,λ/2)𝒰AB(t)𝒰AB0(βA,βB,λ/2)\mathcal{N}_{\beta_{A},\beta_{B}}(t)(\cdot):=\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}\circ\mathcal{U}_{AB}(t)\circ\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}(\cdot)=\mathcal{U}_{AB}^{0}(\beta_{A},\beta_{B},-\lambda/2)\circ\mathcal{U}_{AB}(t)\circ\mathcal{U}_{AB}^{0}(\beta_{A},\beta_{B},\lambda/2) (88)

where UAB0(βA,βB,τ):=exp(i(HAβAτ+HEβBτ))U^{0}_{AB}(\beta_{A},\beta_{B},\tau):=\exp(-i(H_{A}\beta_{A}\tau+H_{E}\beta_{B}\tau)). Since the temperature of the two systems is different, the duration of imaginary time evolution time is also different. We can define

HI(τA,τB):=UA0(τA)UB0(τB)HIUA0(τA)UB0(τB).H_{I}(\tau_{A},\tau_{B}):=U^{0}_{A}(-\tau_{A})U^{0}_{B}(-\tau_{B})H_{I}U^{0}_{A}(\tau_{A})U^{0}_{B}(\tau_{B}). (89)

Similar to section B.2, it is easy to prove

𝒩βA,βB(t)+0t𝑑τI(tτβAλ/2,tτβBλ/2),\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t)\simeq\mathcal{I}+\int_{0}^{t}d\tau\mathcal{L}^{\dagger}_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2), (90)

where I(τA,τB)O=i[HI(τA,τB)OOHI(τA,τB)]\mathcal{L}^{\dagger}_{I}(\tau_{A},\tau_{B})O=i\left[H_{I}(\tau_{A},\tau_{B})O-OH^{\dagger}_{I}(\tau_{A},\tau_{B})\right]. We can partition XX (A and B) into two parts: X0X_{0}, the part close to the sites of HIH_{I}, and X1X_{1}, the part far away from HIH_{I}. When the temperature is high and the evolution time is short, the scale of X0X_{0} can exceed the propagation range of the Lieb-Robinson velocity vLR(|t+βXλ/2|)v_{LR}(\lvert t+\beta_{X}\lambda/2\rvert), but remains much smaller than the overall scale LXL_{X}. The Hamiltonian of the environment can also be divided as HX=HX0+HX1+HX0X1H_{X}=H_{X_{0}}+H_{X_{1}}+H_{X_{0}X_{1}}. We define the truncated Hamiltonian H0T:=HA0+HB0H^{T}_{0}:=H_{A_{0}}+H_{B_{0}}, the corresponding time evolution operator UABT=ei(H0,T+HI)tU^{T}_{AB}=e^{-i(H_{0,T}+H_{I})t}, and the corresponding superoperator 𝒩βT(t):=𝒥ρA0canρB0caniλ/2𝒰ABT(t)𝒥ρA0canρB0caniλ/2\mathcal{N}^{T}_{\beta}(t):=\mathcal{J}^{-i\lambda/2}_{\rho^{\text{can}}_{A_{0}}\otimes\rho^{\text{can}}_{B_{0}}}\circ\mathcal{U}^{T}_{AB}(t)\circ\mathcal{J}^{i\lambda/2}_{\rho^{\text{can}}_{A_{0}}\otimes\rho^{\text{can}}_{B_{0}}}. Similar to Eq. 80, we have:

𝒩βA,βBT(t)+0t𝑑τIT(tτβAλ/2,tτβBλ/2).\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)\simeq\mathcal{I}+\int_{0}^{t}d\tau\mathcal{L}^{T\dagger}_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2). (91)

Appendix C Error estimation

C.1 The error induced by truncation

According to Eqs. 33 and 80, the difference between GF(i)G_{F}(i) and the truncated one is

δGLR(i):=GF(i)GFT(i)=(ISE|𝒥ρE(0)1/2𝒥ρScan1/2[𝒩β(t)𝒩βT(t)]𝒥ρScan1/2|ρS(t))\displaystyle\delta G_{LR}(i):=G_{F}(i)-G^{T}_{F}(i)=(I_{SE}|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ[\mathcal{N}_{\beta}^{\dagger}(t)-\mathcal{N}_{\beta}^{T\dagger}(t)]\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}|\rho_{S}(t))^{\dagger}
0t𝑑τ(ISE|𝒥ρE(0)1/2𝒥ρScan1/2(I(tτiβ/2)IT(tτiβ/2))𝒰S0(t)𝒥ρScan1/2|ρS(t))\displaystyle\simeq\int_{0}^{t}d\tau(I_{SE}|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ(\mathcal{L}^{\dagger}_{I}(t-\tau-i\beta/2)-\mathcal{L}^{T\dagger}_{I}(t-\tau-i\beta/2))\circ\mathcal{U}^{0\dagger}_{S}(t)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\rho_{S}(t))^{\dagger}
=0t𝑑τ(ISE|𝒥ρE(0)1/2𝒰S0(tτiβ/2)[I(tτiβ/2)IT(tτiβ/2)]𝒰S0(τt+iβ/2)𝒰S0(τ)|ρS(t)).\displaystyle=\int_{0}^{t}d\tau(I_{SE}|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{U}^{0}_{S}(t-\tau-i\beta/2)\circ[\mathcal{L}^{\dagger}_{I}(t-\tau-i\beta/2)-\mathcal{L}^{T\dagger}_{I}(t-\tau-i\beta/2)]\circ\mathcal{U}^{0}_{S}(\tau-t+i\beta/2)\circ\mathcal{U}^{0\dagger}_{S}(\tau)|\rho_{S}(t))^{\dagger}. (92)

Using this expression, we can constrain the difference

|δGLR(i)||i0t𝑑τTr{ρE(0)eiHS(tτiβ/2)δHI(tτiβ/2)eiHS(tτiβ/2)[𝒰S0(τ)ρS(t)]}+h.c.|\displaystyle\lvert\delta G_{LR}(i)\rvert\simeq\lvert i\int_{0}^{t}d\tau\text{Tr}\left\{\rho_{E}(0)e^{-iH_{S}(t-\tau-i\beta/2)}\delta H_{I}(t-\tau-i\beta/2)e^{iH_{S}(t-\tau-i\beta/2)}[\mathcal{U}^{0\dagger}_{S}(\tau)\rho_{S}(t)]\right\}+\text{h.c.}\rvert
20t𝑑τeiHS(tτiβ/2)δHI(tτiβ/2)eiHS(tτiβ/2)×𝒰S0(τ)ρS(t)ρE(0)20t𝑑τδHI(tτiβ/2),\displaystyle\leq 2\int_{0}^{t}d\tau\lVert e^{-iH_{S}(t-\tau-i\beta/2)}\delta H_{I}(t-\tau-i\beta/2)e^{iH_{S}(t-\tau-i\beta/2)}\rVert\times\lVert\mathcal{U}^{0\dagger}_{S}(\tau)\rho_{S}(t)\otimes\rho_{E}(0)\rVert\leq 2\int_{0}^{t}d\tau\lVert\delta H^{\prime}_{I}(t-\tau-i\beta/2)\rVert, (93)

where δHI(τ)=eiHEτHIeiHEτeiHB0τHIeiHB0τ\delta H^{\prime}_{I}(\tau)=e^{iH_{E}\tau}H_{I}e^{-iH_{E}\tau}-e^{iH_{B_{0}}\tau}H_{I}e^{-iH_{B_{0}}\tau}. In the second line, we have utilized the Cauchy-Schwartz inequality. According to Eq. 63, we have |δGLR(i)|=o(1)\lvert\delta G_{LR}(i)\rvert=o(1) if |t+iβ/2|</vLR\lvert t+i\beta/2\rvert<\ell/v_{LR}. It’s important to note that we employ the first-order perturbation approximation in this calculation and throughout the remainder of this section.

According to Eqs. 33 and 86, the difference δGLR(i):=GF(i)GFT(i)\delta G^{\prime}_{LR}(i):=G_{F}(i)-G^{T^{\prime}}_{F}(i) is equal to

(I|𝒥ρE(0)1/2𝒥ρScan1/2[𝒩I(iβ/2)𝒰SE(t)𝒩I(iβ/2)𝒩IT(iβ/2)𝒰SE(t)𝒩IT(iβ/2)]𝒥ρScan1/2|ρS(t))\displaystyle(I|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ[\mathcal{N}_{I}(i\beta/2)\circ\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{N}^{\dagger}_{I}(-i\beta/2)-\mathcal{N}^{T}_{I}(i\beta/2)\circ\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{N}^{T\dagger}_{I}(-i\beta/2)]\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}|\rho_{S}(t))^{\dagger}
0β/2𝑑τ1(I|𝒥ρE(0)1/2𝒥ρScan1/2[IA(iτ1)IA,T(iτ1)]𝒰SE(t)[I+0β/2𝑑τ2IA(iτ2)]𝒥ρScan1/2|ρS(t))\displaystyle\simeq\int_{0}^{-\beta/2}d\tau_{1}(I|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\left[\mathcal{L}^{\text{A}}_{I}(i\tau_{1})-\mathcal{L}^{\text{A},T}_{I}(i\tau_{1})\right]\circ\mathcal{U}^{\dagger}_{SE}(t)\circ[I+\int_{0}^{\beta/2}d\tau_{2}\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2})]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\rho_{S}(t))^{\dagger}
+0β/2𝑑τ2(I|𝒥ρE(0)1/2𝒥ρScan1/2[I+0β/2𝑑τ1IA,T(iτ1)]𝒰SE(t)[IA(iτ2)IA,T(iτ2)]𝒥ρScan1/2|ρS(t)).\displaystyle+\int_{0}^{\beta/2}d\tau_{2}(I|\mathcal{J}_{\rho_{E}(0)}^{1/2}\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ[I+\int_{0}^{-\beta/2}d\tau_{1}\mathcal{L}^{\text{A},T}_{I}(i\tau_{1})]\circ\mathcal{U}^{\dagger}_{SE}(t)\circ\left[\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2})-\mathcal{L}^{\text{A},T\dagger}_{I}(i\tau_{2})\right]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\rho_{S}(t))^{\dagger}. (94)

Using this expression, we can constrain the difference

|δGLR(i)||0β/2𝑑τ1Tr{ρScanρE(0)δHI(iτ1)[𝒰SE(t)(I+0β/2𝑑τ2IA(iτ2))𝒥ρScan1/2ρS(t)]}+h.c.|\displaystyle\lvert\delta G^{\prime}_{LR}(i)\rvert\simeq\lvert\int_{0}^{-\beta/2}d\tau_{1}\text{Tr}\left\{\rho_{S}^{\text{can}}\otimes\rho_{E}(0)\delta H_{I}(i\tau_{1})[\mathcal{U}^{\dagger}_{SE}(t)\circ(I+\int_{0}^{\beta/2}d\tau_{2}\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2}))\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}\rho_{S}(t)]\right\}+\text{h.c.}\rvert
+|0β/2𝑑τ2Tr{[𝒰SE(t)(I+0β/2𝑑τ1IA,T(iτ1))ρScanρE(0)]δHI(iτ2)[𝒥ρScan1/2ρS(t)]}+h.c.|\displaystyle+\lvert\int_{0}^{\beta/2}d\tau_{2}\text{Tr}\left\{[\mathcal{U}_{SE}(t)\circ(I+\int_{0}^{-\beta/2}d\tau_{1}\mathcal{L}^{\text{A},T}_{I}(i\tau_{1}))\rho_{S}^{\text{can}}\otimes\rho_{E}(0)]\delta H_{I}(i\tau_{2})[\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}\rho_{S}(t)]\right\}+\text{h.c.}\rvert
20β/2𝑑τρScanρE(0)×δHI(iτ)×[2I+20β/2𝑑τ(HI(iτ)+HIT(iτ))]×𝒥ρScan1/2ρS(t).\displaystyle\leq 2\int_{0}^{\beta/2}d\tau\lVert\rho_{S}^{\text{can}}\otimes\rho_{E}(0)\rVert\times\lVert\delta H_{I}(i\tau)\rVert\times\left[2\lVert I\rVert+2\int_{0}^{\beta/2}d\tau^{\prime}(\lVert H_{I}(i\tau^{\prime})\rVert+\lVert H^{T}_{I}(i\tau^{\prime})\rVert)\right]\times\lVert\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}\rho_{S}(t)\rVert. (95)

The term 𝒥ρScan1/2ρS(t)\lVert\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}\rho_{S}(t)\rVert is related to the final state of the system. It can be greater than 11 but will not increase with the size of the environment. According to Eqs. 61 and 63, we have |δGLR(i)|=o(1)\lvert\delta G^{\prime}_{LR}(i)\rvert=o(1) if |iβ/2|</vLR\lvert i\beta/2\rvert<\ell/v_{LR}. Since |δGLR(i)¯||δGLR(i)|¯=o(1)\lvert\overline{\delta G^{\prime}_{LR}(i)}\rvert\leq\overline{\lvert\delta G^{\prime}_{LR}(i)\rvert}=o(1), the error is very small under long-time average.

According to Eqs. 48 and 91, the error induced by truncation in section IV can be estimated as

δGLR(i)=(IAB|𝒥ρA(0)ρB(0)1/2[𝒩βA,βB(t)𝒩βA,βBT(t)]|1)\displaystyle\delta G_{LR}(i)=(I_{AB}|\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ[\mathcal{N}^{\dagger}_{\beta_{A},\beta_{B}}(t)-\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)]|1)^{\dagger}
0t𝑑τ(IAB|𝒥ρA(0)ρB(0)1/2[I(tτβAλ/2,tτβBλ/2)IT(tτβAλ/2,tτβBλ/2)]|1)\displaystyle\simeq\int_{0}^{t}d\tau(I_{AB}|\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ[\mathcal{L}^{\dagger}_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2)-\mathcal{L}^{T\dagger}_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2)]|1)^{\dagger} (96)

Using this expression, we can constrain the difference as

|δGLR(i)||i0t𝑑τTr{ρA(0)ρB(0)δHI(tτβAλ/2,tτβBλ/2)}+h.c.|\displaystyle\lvert\delta G_{LR}(i)\rvert\simeq\lvert i\int_{0}^{t}d\tau\text{Tr}\left\{\rho_{A}(0)\otimes\rho_{B}(0)\delta H_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2)\right\}+\text{h.c.}\rvert
20t𝑑τδHI(tτβAλ/2,tτβBλ/2)×ρA(0)ρB(0).\displaystyle\leq 2\int_{0}^{t}d\tau\lVert\delta H_{I}(t-\tau-\beta_{A}\lambda/2,t-\tau-\beta_{B}\lambda/2)\rVert\times\lVert\rho_{A}(0)\otimes\rho_{B}(0)\rVert. (97)

In the second line, we have used the Cauchy-Schwartz inequality. According to Eq. 63, we have |δGLR(i)|=o(1)\lvert\delta G_{LR}(i)\rvert=o(1) if |t+iβX/2|<X/vLR\lvert t+i\beta_{X}/2\rvert<\ell_{X}/v_{LR} holds for both AA and BB.

According to Eq. 86, the error induced by truncation in section IV can be calculated as

(I|𝒥ρA(0)ρB(0)1/2[𝒩I(iβ/2)𝒰AB(t)𝒩I(iβ/2)𝒩IT(iβ/2)𝒰SE(t)𝒩IT(iβ/2)]|1)\displaystyle(I|\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ[\mathcal{N}_{I}(i\beta/2)\circ\mathcal{U}^{\dagger}_{AB}(t)\circ\mathcal{N}^{\dagger}_{I}(-i\beta/2)-\mathcal{N}^{T}_{I}(i\beta/2)\circ\mathcal{U}^{\dagger}_{SE}(t)\circ\mathcal{N}^{T\dagger}_{I}(-i\beta/2)]|1)^{\dagger}
0β/2𝑑τ1(I|𝒥ρA(0)ρB(0)1/2[IA(iτ1)IA,T(iτ1)]𝒰b(t)[I+0β/2𝑑τ2IA(iτ2)]|1)\displaystyle\simeq\int_{0}^{-\beta/2}d\tau_{1}(I|\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ\left[\mathcal{L}^{\text{A}}_{I}(i\tau_{1})-\mathcal{L}^{\text{A},T}_{I}(i\tau_{1})\right]\circ\mathcal{U}^{\dagger}_{b}(t)\circ[I+\int_{0}^{\beta/2}d\tau_{2}\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2})]|1)^{\dagger}
+0β/2𝑑τ2(I|𝒥ρA(0)ρB(0)1/2[I+0β/2𝑑τ1IA,T(iτ1)]𝒰AB(t)[IA(iτ2)IA,T(iτ2)]|1)\displaystyle+\int_{0}^{\beta/2}d\tau_{2}(I|\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}\circ[I+\int_{0}^{-\beta/2}d\tau_{1}\mathcal{L}^{\text{A},T}_{I}(i\tau_{1})]\circ\mathcal{U}^{\dagger}_{AB}(t)\circ\left[\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2})-\mathcal{L}^{\text{A},T\dagger}_{I}(i\tau_{2})\right]|1)^{\dagger} (98)

Using this expression, we can constrain the difference

|δGLR(i)||0β/2𝑑τ1Tr{ρA(0)ρB(0)δHI(iτ1)[𝒰AB(t)(I+0β/2𝑑τ2IA(iτ2))IAB]}+h.c.|\displaystyle\lvert\delta G^{\prime}_{LR}(i)\rvert\simeq\lvert\int_{0}^{-\beta/2}d\tau_{1}\text{Tr}\left\{\rho_{A}(0)\otimes\rho_{B}(0)\delta H_{I}(i\tau_{1})[\mathcal{U}^{\dagger}_{AB}(t)\circ(I+\int_{0}^{\beta/2}d\tau_{2}\mathcal{L}^{\text{A}\dagger}_{I}(i\tau_{2}))I_{AB}]\right\}+\text{h.c.}\rvert
+|0β/2𝑑τ2Tr{[𝒰AB(t)(I+0β/2𝑑τ1IA,T(iτ1))ρA(0)ρB(0)]δHI(iτ2)IAB}+h.c.|\displaystyle+\lvert\int_{0}^{\beta/2}d\tau_{2}\text{Tr}\left\{[\mathcal{U}_{AB}(t)\circ(I+\int_{0}^{-\beta/2}d\tau_{1}\mathcal{L}^{\text{A},T}_{I}(i\tau_{1}))\rho_{A}(0)\otimes\rho_{B}(0)]\delta H_{I}(i\tau_{2})I_{AB}\right\}+\text{h.c.}\rvert
20β/2𝑑τρA(0)ρB(0)×δHI(iτ)×[2I+20β/2𝑑τ(HI(iτ)+HIT(iτ))].\displaystyle\leq 2\int_{0}^{\beta/2}d\tau\lVert\rho_{A}(0)\otimes\rho_{B}(0)\rVert\times\lVert\delta H_{I}(i\tau)\rVert\times\left[2\lVert I\rVert+2\int_{0}^{\beta/2}d\tau^{\prime}(\lVert H_{I}(i\tau^{\prime})\rVert+\lVert H^{T}_{I}(i\tau^{\prime})\rVert)\right]. (99)

According to Eqs. 61 and 63, we have |δGLR(i)|=o(1)\lvert\delta G^{\prime}_{LR}(i)\rvert=o(1) if |iβ/2|<X/vLR\lvert i\beta/2\rvert<\ell_{X}/v_{LR} holds for both AA and BB. Since |δGLR(i)¯||δGLR(i)|¯=o(1)\lvert\overline{\delta G^{\prime}_{LR}(i)}\rvert\leq\overline{\lvert\delta G^{\prime}_{LR}(i)\rvert}=o(1), the error is very small under long-time average.

C.2 The error induced by deviation of the environment state in short-time regime

The error induced by deviation of the environment state in section III.1 equals to

(ISE|[𝒥ρE(0)1/2𝒥ρEcan1/2]𝒥ρScan1/2𝒩βT(t)𝒥ρScan1/2|ρS(t))=(OB0IB0¯|ρE(0)ρEcan)=(OB0|TrB0¯[ρE(0)ρEcan]),(I_{SE}|[\mathcal{J}_{\rho_{E}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}_{\beta}^{T\dagger}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}|\rho_{S}(t))^{\dagger}=(O_{B_{0}}\otimes I_{\overline{B_{0}}}|\rho_{E}(0)-\rho^{\text{can}}_{E})=(O_{B_{0}}|\text{Tr}_{\overline{B_{0}}}[\rho_{E}(0)-\rho^{\text{can}}_{E}]), (100)

where (OB0|=(ρS(t)ρB0can|𝒰SET(t)𝒥ρB0can1/2|IS)(O_{B_{0}}|=(\rho_{S}(t)\otimes\rho^{\text{can}}_{B_{0}}|\mathcal{U}^{T}_{SE}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{B_{0}}}|I_{S}) is a B0B_{0} local operator. For λ=i\lambda=i, using a similar approximation to Eq. 74 for 𝒰SET(t)\mathcal{U}^{T}_{SE}(t), we have

OB0IB0+0t𝑑τ(ρS(t)|𝒰S0(τ)𝒥ρB0can1/2I𝒥ρB0can1/2|IS)IB0+20t𝑑τHI(iβ/2)×ρS(t).\lVert O_{B_{0}}\rVert\simeq\lVert I_{B_{0}}+\int_{0}^{t}d\tau(\rho_{S}(t)|\mathcal{U}^{0}_{S}(\tau)\circ\mathcal{J}^{1/2}_{\rho^{\text{can}}_{B_{0}}}\circ\mathcal{L}_{I}\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{B_{0}}}|I_{S})\rVert\leq\lVert I_{B_{0}}\rVert+2\int_{0}^{t}d\tau\lVert H_{I}(i\beta/2)\rVert\times\lVert\rho_{S}(t)\rVert. (101)

According to Eq. 61, we proceed to bound the above expression, yielding OB0=Θ(1)\lVert O_{B_{0}}\rVert=\Theta(1). It’s worth mentioning that we employ the first-order perturbation approximation here, and higher-order terms can be bounded similarly. Furthermore, irrespective of the scenario, the operator OB0O_{B_{0}}’s norm should not escalate with the size of the entire environment. Consequently, in line with Eq. 64, we have:

(OB0|TrB0¯[ρE(0)ρEcan])=O(Nα).(O_{B_{0}}|\text{Tr}_{\overline{B_{0}}}[\rho_{E}(0)-\rho^{\text{can}}_{E}])=O(N^{-\alpha}). (102)

The error induced by deviation of the environment state in section IV is

(IAB|[𝒥ρA(0)ρB(0)1/2𝒥ρAcanρBcan1/2]𝒩βA,βBT(t)|1)=(OA0B0|TrA0¯B0¯[ρA(0)ρB(0)ρAcanρBcan]),(I_{AB}|[\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}]\circ\mathcal{N}^{T\dagger}_{\beta_{A},\beta_{B}}(t)|1)^{\dagger}=(O_{A_{0}B_{0}}|\text{Tr}_{\overline{A_{0}}\overline{B_{0}}}[\rho_{A}(0)\otimes\rho_{B}(0)-\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}]), (103)

where (OA0B0|=(ρA0canρB0can|𝒰ABT(t)𝒥ρA0canρB0can1/2(O_{A_{0}B_{0}}|=(\rho^{\text{can}}_{A_{0}}\otimes\rho^{\text{can}}_{B_{0}}|\mathcal{U}^{T}_{AB}(t)\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{A_{0}}\otimes\rho^{\text{can}}_{B_{0}}} is an A0B0A_{0}B_{0} local operator. Following the similar arguments as in Eq. 102, in accordance with Eq. 64, we should have

(OA0B0|TrA0¯B0¯[ρA(0)ρB(0)ρAcanρBcan])=O(NAα)+O(NBα),(O_{A_{0}B_{0}}|\text{Tr}_{\overline{A_{0}}\overline{B_{0}}}[\rho_{A}(0)\otimes\rho_{B}(0)-\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}])=O(N_{A}^{-\alpha})+O(N_{B}^{-\alpha}), (104)

where NXN_{X} is the size of the bath XX.

C.3 The error induced by deviation of the environment state in long-time regime

According to Eqs. 41 and 87, the term in Eq. 43 can be rewritten as

(ISE|[𝒥ρE(0)1/2𝒥ρEcan1/2]𝒥ρScan1/2𝒩βT(t)¯𝒥ρScan1/2|ρS(t)¯)\displaystyle(I_{SE}|[\mathcal{J}_{\rho_{E}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{E}}^{1/2}]\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\overline{\mathcal{N}_{\beta}^{T^{\prime}\dagger}(t)}\circ\mathcal{J}^{-1/2}_{\rho^{\text{can}}_{S}}|\overline{\rho_{S}(t)})^{\dagger}
=a(ρScan[ρE(0)ρEcan]|𝒩IT(iβ/2)|Πa)×(Πa|𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯).\displaystyle=\sum_{a}(\rho_{S}^{\text{can}}\otimes[\rho_{E}(0)-\rho^{\text{can}}_{E}]|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{a})^{\dagger}\times(\Pi_{a}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)})^{\dagger}. (105)

The energy width of ρScanρE(0)\rho_{S}^{\text{can}}\otimes\rho_{E}(0) and ρScanρEcan\rho_{S}^{\text{can}}\otimes\rho^{\text{can}}_{E} are both narrower than Θ(NS+B1/2)\Theta(N_{S+B}^{1/2}) [32]. According to the theorem 2.1 in [25], the superoperator 𝒩IT(iβ/2)\mathcal{N}^{T}_{I}(-i\beta/2) at most increase the energy width by Θ(NB0)\Theta(N_{B_{0}}). According to the theorems 2.2 and 2.3 in [25], the difference between the eigenstates of H0H_{0} and HH causes the energy width to increase by Θ(|HI|)\Theta(\lvert H_{I}\rvert) at most. For the above reasons, the a,ba,b in sections C.3 and III.2 can be limited to the energy shell [EcΔE/2,Ec+ΔE/2][E_{c}-\Delta E/2,E_{c}+\Delta E/2], where ΔE=Θ(NB0)+Θ(NS+B1/2)+Θ(|HI|)\Delta E=\Theta(N_{B_{0}})+\Theta(N_{S+B}^{1/2})+\Theta(\lvert H_{I}\rvert) and Ec=(ρScanρE(0)|H0)E_{c}=(\rho_{S}^{\text{can}}\otimes\rho_{E}(0)|H_{0}). The weights of eigenstates with energies above this range are suppressed exponentially. The energy width ΔE\Delta E is much smaller than the average energy Θ(NSB)\Theta(N_{SB}). Their inverse temperature range given by Eq. 35 should be narrow and negligible in the thermodynamic limit. This means that it is difficult to distinguish these states from the same canonical ensemble ρSEcan\rho_{SE}^{\text{can}} through local operator 𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯IE)\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)}\otimes I_{E}). Therefore, we can do the following approximate replacement

a|Πa)(Πa|a|Πa)(ρSEcan|=|ISE)(ρSEcan|,\sum_{a}|\Pi_{a})(\Pi_{a}|\sim\sum_{a}|\Pi_{a})(\rho_{SE}^{\text{can}}|=|I_{SE})(\rho_{SE}^{\text{can}}|, (106)

The error induced by replacement Eq. 106 can be bounded as

|a(ρScan[ρE(0)ρEcan]|𝒩IT(iβ/2)|Πa)×(ΠaρSEcan|𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯)|\displaystyle\lvert\sum_{a}(\rho_{S}^{\text{can}}\otimes[\rho_{E}(0)-\rho^{\text{can}}_{E}]|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{a})^{\dagger}\times(\Pi_{a}-\rho_{SE}^{\text{can}}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)})^{\dagger}\rvert
[|(ρScanρE(0)|𝒩IT(iβ/2)|I)|+|(ρScanρEcan|𝒩IT(iβ/2)|I)|]×|(ΠmaxρSEcan|𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯)|,\displaystyle\leq\left[\lvert(\rho_{S}^{\text{can}}\otimes\rho_{E}(0)|\mathcal{N}^{T}_{I}(i\beta/2)|I)^{\dagger}\rvert+\lvert(\rho_{S}^{\text{can}}\otimes\rho^{\text{can}}_{E}|\mathcal{N}^{T}_{I}(i\beta/2)|I)^{\dagger}\rvert\right]\times\lvert(\Pi_{\max}-\rho_{SE}^{\text{can}}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)})^{\dagger}\rvert, (107)

where Emax[EcΔE/2,Ec+ΔE/2]E_{\max}\in[E_{c}-\Delta E/2,E_{c}+\Delta E/2] makes |(ΠmaxρSEcan|𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯)|\lvert(\Pi_{\max}-\rho_{SE}^{\text{can}}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)})^{\dagger}\rvert maximize. In section C.3, we use the properties that the terms like (ρScanρE|𝒩IT(iβ/2)|Πa)(\rho_{S}^{\text{can}}\otimes\rho_{E}|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{a}) are positive. According to the previous analysis of the energy shell and Eq. 64, section C.3 can be bounded with O(Nα)O(N^{-\alpha}) and is small at the thermodynamic limit. The operator state (IS|𝒥ρScan1/2𝒩IT(iβ/2)|ISE)(I_{S}|\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}^{T}_{I}(i\beta/2)|I_{SE}) also gives a B0B_{0} local operator. After applying replacement (106) to section C.3 and combining Eq. 64, we have

(IS[ρE(0)ρEcan]|𝒥ρScan1/2𝒩IT(iβ/2)|ISE)×(ρSEcan|𝒩IT(iβ/2)𝒥ρScan1/2|ρS(t)¯)=O(Nα).(I_{S}\otimes[\rho_{E}(0)-\rho^{\text{can}}_{E}]|\mathcal{J}_{\rho_{S}^{\text{can}}}^{1/2}\circ\mathcal{N}^{T}_{I}(i\beta/2)|I_{SE})^{\dagger}\times(\rho_{SE}^{\text{can}}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\overline{\rho_{S}(t)})^{\dagger}=O(N^{-\alpha}). (108)

According to Eqs. 41 and 87, the error induced by the term in the second line of section IV can be calculated with

(IAB|[𝒥ρA(0)ρB(0)1/2𝒥ρAcanρBcan1/2]𝒩βT(t)¯|1)=a(ρA(0)ρB(0)ρAcanρBcan|𝒩IT(iβ/2)|Πa)×(Πa|𝒩IT(iβ/2)|IAB)(I_{AB}|[\mathcal{J}_{\rho_{A}(0)\otimes\rho_{B}(0)}^{1/2}-\mathcal{J}_{\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}}^{1/2}]\circ\overline{\mathcal{N}^{T^{\prime}\dagger}_{\beta}(t)}|1)^{\dagger}=\sum_{a}(\rho_{A}(0)\otimes\rho_{B}(0)-\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{a})^{\dagger}\times(\Pi_{a}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)|I_{AB})^{\dagger} (109)

Following the similar arguments in the preceding paragraph, the aa in Eq. 109 can be limited to the energy shell [EcΔE/2,Ec+ΔE/2][E_{c^{\prime}}-\Delta E/2,E_{c^{\prime}}+\Delta E/2], where ΔE=Θ(NB0+NA0)+Θ(NA+B1/2)+Θ(|HI|)\Delta E=\Theta(N_{B_{0}}+N_{A_{0}})+\Theta(N_{A+B}^{1/2})+\Theta(\lvert H_{I}\rvert) and Ec=(ρA(0)ρB(0)|H0)E_{c^{\prime}}=(\rho_{A}(0)\otimes\rho_{B}(0)|H_{0}). The weights of eigenstates with energies above this range are suppressed exponentially. Moreover, since 𝒩IT(iβ/2)|IAB)=:|OA0B0IA0B0¯)\mathcal{N}^{T\dagger}_{I}(-i\beta/2)|I_{AB})=:|O_{A_{0}B_{0}}\otimes I_{\overline{A_{0}B_{0}}}) gives an A0B0A_{0}B_{0} local operator and the energy width to be considered here is also much smaller than the average energy Θ(NAB)\Theta(N_{AB}), the ETH (64) allows us to replace (Πa|(\Pi_{a}| with (ρABcan|(\rho_{AB}^{\text{can}}| and introduce only a small error. Now the approximate replacement becomes a|Πa)(Πa|a|Πa)(ρABcan|=|IAB)(ρABcan|\sum_{a}|\Pi_{a})(\Pi_{a}|\sim\sum_{a}|\Pi_{a})(\rho_{AB}^{\text{can}}|=|I_{AB})(\rho_{AB}^{\text{can}}|. The error induced by this replacement can be bounded in similar ways as section C.3. It is also small at the thermodynamic limit. The operator state (ρA(0)|𝒩IT(iβ/2)|IAB)(\rho_{A}(0)|\mathcal{N}^{T}_{I}(i\beta/2)|I_{AB}) gives a B0B_{0} local operator and (ρBcan|𝒩IT(iβ/2)|IAB)(\rho^{\text{can}}_{B}|\mathcal{N}^{T}_{I}(i\beta/2)|I_{AB}) gives an A0A_{0} local operator. So according to Eq. 64, we have

(ρA(0)ρB(0)ρAcanρBcan|𝒩IT(iβ/2)|IAB)×(ρABcan|𝒩IT(iβ/2)|IAB)=O(NAα)+O(NBα).(\rho_{A}(0)\otimes\rho_{B}(0)-\rho^{\text{can}}_{A}\otimes\rho^{\text{can}}_{B}|\mathcal{N}^{T}_{I}(i\beta/2)|I_{AB})^{\dagger}\times(\rho_{AB}^{\text{can}}|\mathcal{N}^{T\dagger}_{I}(-i\beta/2)|I_{AB})^{\dagger}=O(N_{A}^{-\alpha})+O(N_{B}^{-\alpha}). (110)

C.4 The error caused by time-dependent final state measurements

Now let us estimate the bound of section III.2. The energy width of ρS(0)ρE(0)\rho_{S}(0)\otimes\rho_{E}(0) depend on the initial state, but it can be safely assumed to be narrower than Θ(NS)\Theta(N_{S}). After considering the difference between the eigenstates of H0H_{0} and HH, the a,ba,b in section III.2 should also be in the energy shell [EcΔE/2,Ec+ΔE/2][E^{\prime}_{c}-\Delta E^{\prime}/2,E^{\prime}_{c}+\Delta E^{\prime}/2], where ΔE=Θ(NS)+Θ(|HI|)\Delta E^{\prime}=\Theta(N_{S})+\Theta(\lvert H_{I}\rvert) and Ec=(H0|ρS(0)ρE(0))E^{\prime}_{c}=(H_{0}|\rho_{S}(0)\otimes\rho_{E}(0)). The a,ba,b should be in the intersection of [EcΔE/2,Ec+ΔE/2][E_{c}-\Delta E/2,E_{c}+\Delta E/2] and [EΔE/2,E+ΔE/2][E^{\prime}-\Delta E^{\prime}/2,E^{\prime}+\Delta E^{\prime}/2]. The weights of eigenstates with energies above this range are suppressed exponentially. Since 𝒩IT(iβ/2)𝒥ρScan1/2|TrEΠabIE)\mathcal{N}^{T\dagger}_{I}(-i\beta/2)\circ\mathcal{J}_{\rho_{S}^{\text{can}}}^{-1/2}|\text{Tr}_{E}\Pi_{ab}\otimes I_{E}) also gives an SB0SB_{0} local operator. The off-diagonal ETH (65) tells

|(Πab|OSB0IB0¯)|=O(Nα).\lvert(\Pi_{ab}|O_{SB_{0}}\otimes I_{\overline{B_{0}}})\rvert=O(N^{-\alpha}). (111)

The remaining part of section III.2 can be evaluated as follows

a,bab|(ρScanδρE|𝒩IT(iβ/2)|Πab)(Πab|ρS(0)ρE(0))|\displaystyle\sum_{\begin{subarray}{c}a,b\\ a\neq b\end{subarray}}\big{|}(\rho_{S}^{\text{can}}\otimes\delta\rho_{E}|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{ab})(\Pi_{ab}|\rho_{S}(0)\otimes\rho_{E}(0))\big{|}
a,b|(ρScanδρE|𝒩IT(iβ/2)|Πab)(Πa|ρS(0)ρE(0))(Πb|ρS(0)ρE(0))|\displaystyle\leq\sum_{\begin{subarray}{c}a,b\end{subarray}}\big{|}(\rho_{S}^{\text{can}}\otimes\delta\rho_{E}|\mathcal{N}^{T}_{I}(i\beta/2)|\Pi_{ab})\sqrt{(\Pi_{a}|\rho_{S}(0)\otimes\rho_{E}(0))}\sqrt{(\Pi_{b}|\rho_{S}(0)\otimes\rho_{E}(0))}\big{|}
|𝒩IT(iβ/2)(ρScanδρE)|P2×P2|𝒩IT(iβ/2)(ρScanδρE)|2(1+20β/2𝑑τ|HI(iτ)|)=Θ(1),\displaystyle\leq\lVert\lvert\mathcal{N}^{T{\dagger}}_{I}(i\beta/2)(\rho_{S}^{\text{can}}\otimes\delta\rho_{E})\rvert\sqrt{P}\rVert_{2}\times\lVert\sqrt{P}\rVert_{2}\leq\lVert\lvert\mathcal{N}^{T{\dagger}}_{I}(i\beta/2)(\rho_{S}^{\text{can}}\otimes\delta\rho_{E})\rvert\rVert\leq 2(1+2\int_{0}^{\beta/2}d\tau\lVert\lvert H_{I}(i\tau)\rvert\rVert)=\Theta(1), (112)

where the vector P=({(Πa|ρSE(0))})\sqrt{P}=(\{\sqrt{(\Pi_{a}|\rho_{SE}(0))}\}), 2\lVert\cdot\rVert_{2} is vector 2-norm and ||\lvert\cdot\rvert replaces the matrix elements with their absolute values. In the second line, we have inserted the nonnegative part of a=ba=b and used |ρab|2ρaρb\lvert\rho_{ab}\rvert^{2}\leq\rho_{a}\rho_{b}. In the third line have used Cauchy-Schwartz inequality. Here we assume that the norm \lVert\cdot\rVert is consistent with vector 2-norm. Additionally, we employ the approximation (86) and Eq. 61 in the third inequality sign on the third line. Combining it with Eq. 111, we can constrain section III.2 to the following range

Θ(1)×O(Nα).\Theta(1)\times O(N^{-\alpha}). (113)

References

  • [1] M. Esposito, U. Harbola, S. Mukamel, Nonequilibrium fluctuations, fluctuation theorems, and counting statistics in quantum systems, Reviews of Modern Physics, 81 (2009) 1665-1702.
  • [2] E. Iyoda, K. Kaneko, T. Sagawa, Fluctuation Theorem for Many-Body Pure Quantum States, Physical Review Letters, 119 (2017) 100601.
  • [3] E. Iyoda, K. Kaneko, T. Sagawa, Eigenstate fluctuation theorem in the short- and long-time regimes, Physical Review E, 105 (2022) 044106.
  • [4] C. Jarzynski, D.K. Wójcik, Classical and Quantum Fluctuation Theorems for Heat Exchange, Physical Review Letters, 92 (2004) 230602.
  • [5] G. Manzano, J.M. Horowitz, J.M.R. Parrondo, Nonequilibrium potential and fluctuation theorems for quantum maps, Physical Review E, 92 (2015) 032129.
  • [6] Á.M. Alhambra, L. Masanes, J. Oppenheim, C. Perry, Fluctuating Work: From Quantum Thermodynamical Identities to a Second Law Equality, Physical Review X, 6 (2016) 041017.
  • [7] G. Manzano, J.M. Horowitz, J.M.R. Parrondo, Quantum Fluctuation Theorems for Arbitrary Environments: Adiabatic and Nonadiabatic Entropy Production, Physical Review X, 8 (2018) 031037.
  • [8] H. Kwon, M.S. Kim, Fluctuation Theorems for a Quantum Channel, Physical Review X, 9 (2019) 031029.
  • [9] G.T. Landi, M. Paternostro, Irreversible entropy production: From classical to quantum, Reviews of Modern Physics, 93 (2021) 035008.
  • [10] Z. Holmes, G. Muraleedharan, R.D. Somma, Y. Subasi, B. Şahinoǧlu, Quantum algorithms from fluctuation theorems: Thermal-state preparation, Quantum, 6 (2022) 825.
  • [11] J. Haah, M.B. Hastings, R. Kothari, G.H. Low, Quantum algorithm for simulating real time evolution of lattice Hamiltonians, SIAM Journal on Computing, DOI (2021) FOCS18-250-FOCS218-284.
  • [12] D. E. Parker, X.-y. Cao, A. Avdoshkin, T. Scaffidi, and E. Altman, A universal operator growth hypothesis, Phys. Rev. X 9, 041017 (2019).
  • [13] L.P. García-Pintos, N. Linden, A.S.L. Malabarba, A.J. Short, A. Winter, Equilibration Time Scales of Physically Relevant Observables, Physical Review X, 7 (2017) 031027.
  • [14] P. Figueroa-Romero, K. Modi, F.A. Pollock, Equilibration on average in quantum processes with finite temporal resolution, Physical Review E, 102 (2020) 032144.
  • [15] Z. Huang, Multiple entropy production for multitime quantum processes, Physical Review A, 108 (2023) 032217.
  • [16] F. A. Pollock, C. Rodríguez-Rosario, T. Frauenheim, M. Paternostro, and K. Modi, Non-Markovian quantum processes: Complete framework and efficient characterization, Phys. Rev. A 97, 012127 (2018).
  • [17] S. Milz and K. Modi, Quantum stochastic processes and quantum non-Markovian phenomena, PRX Quantum 2, 030201 (2021).
  • [18] K. Micadei, G.T. Landi, E. Lutz, Quantum Fluctuation Theorems beyond Two-Point Measurements, Physical Review Letters, 124 (2020) 090602.
  • [19] J.M. Horowitz, S. Vaikuntanathan, Nonequilibrium detailed fluctuation theorem for repeated discrete feedback, Physical Review E, 82 (2010) 061120.
  • [20] S. Lahiri, S. Rana, A.M. Jayannavar, Fluctuation theorems in the presence of information gain and feedback, Journal of Physics A: Mathematical and Theoretical, 45 (2012) 065002.
  • [21] P.A. Camati, R.M. Serra, Verifying detailed fluctuation relations for discrete feedback-controlled quantum dynamics, Physical Review A, 97 (2018) 042127.
  • [22] V. Jakšić, C.A. Pillet, M. Westrich, Entropic Fluctuations of Quantum Dynamical Semigroups, Journal of Statistical Physics, 154 (2014) 153-187.
  • [23] J.-F. Bougron, L. Bruneau, Linear Response Theory and Entropic Fluctuations in Repeated Interaction Quantum Systems, Journal of Statistical Physics, 181 (2020) 1636-1677.
  • [24] E. Wang, U. Heinz, Generalized fluctuation-dissipation theorem for nonlinear response functions, Physical Review D, 66 (2002) 025008.
  • [25] I. Arad, T. Kuwahara, Z. Landau, Connecting global and local energy distributions in quantum spin models on a lattice, Journal of Statistical Mechanics: Theory and Experiment, 2016 (2016) 033301.
  • [26] A. Dymarsky, N. Lashkari, H. Liu, Subsystem eigenstate thermalization hypothesis, Physical Review E, 97 (2018) 012140.
  • [27] T. Kuwahara, K. Saito, Eigenstate Thermalization from the Clustering Property of Correlation, Physical Review Letters, 124 (2020) 200604.
  • [28] Z. Huang, X.-K. Guo, Subsystem eigenstate thermalization hypothesis for translation invariant systems, arXiv preprint arXiv:2312.00410 (2023).
  • [29] F. Verstraete, J.J. García-Ripoll, J.I. Cirac, Matrix Product Density Operators: Simulation of Finite-Temperature and Dissipative Systems, Physical Review Letters, 93 (2004) 207204.
  • [30] B. Bagchi, A. Fring, Minimal length in quantum mechanics and non-Hermitian Hamiltonian systems, Physics Letters A, 373 (2009) 4307-4310.
  • [31] R. Dann, N. Megier, R. Kosloff, Non-Markovian dynamics under time-translation symmetry, Physical Review Research, 4 (2022) 043075.
  • [32] H. Tasaki, On the Local Equivalence Between the Canonical and the Microcanonical Ensembles for Quantum Spin Systems, Journal of Statistical Physics, 172 (2018) 905-926.