Polaron-transformed quantum master equation (PQME) offers a unified framework to describe the dynamics of quantum systems in both limits of weak and strong couplings to environmental degrees of freedom. Thus, the PQME serves as an efficient method to describe charge and exciton transfer/transport dynamics for a broad range of parameters in condensed or complex environments. However, in some cases, the polaron transformation (PT) being employed in the formulation invokes an over-relaxation of slow modes and results in premature suppression of important coherence terms. A formal framework to address this issue is developed in the present work by employing a partial PT that has smaller weights for low frequency bath modes. It is shown here that a closed form expression of a second order time-local PQME including all the inhomogeneous terms can be derived for a general form of partial PT, although more complicated than that for the full PT. All the expressions needed for numerical calculation are derived in detail. Applications to a model of a two-level system coupled to a bath of harmonic oscillators, with test calculations focused on those due to homogeneous relaxation terms, demonstrate the feasibility and the utility of the present approach.

Polaron transformation (PT)1–31 has served as both an important conceptual framework and an efficient computational tool for describing charge and exciton transfer/transport dynamics in various condensed and complex media. PT creates a polaron picture where molecular vibrations and phonon modes of environments, collectively referred to as bath here, relax differently with respect to different site localized states. When such responses to localized states are fast, the states “dressed” with those responses, so-called polarons, serve as effective means to represent the contribution of the bath because they have already taken some contributions of the bath responses into consideration up to infinite order. Alternatively, PT can simply be viewed as a useful unitary transformation in the combined space of system and bath, which produces a new renormalized interaction Hamiltonian term that remains small even within the limit of strong couplings to the bath. As long as physical observables of interest remain unentangled with the bath through the PT, a quantum master equation (QME) derived by projecting out the bath in the polaron picture can be used for calculating such observables. Indeed, utilizing this fact has led to a general approach called polaron-transformed QME (PQME).14–22 

The merit of PQME is its efficiency in accounting for both weak and strong couplings to the bath, while offering reasonable interpolation between the two limits. Earlier versions of PQMEs14–17 employed a full PT where all the bath degrees of freedom, modeled as harmonic oscillators, are fully relaxed to site local system states. However, in some cases, this is not advantageous because it invokes an over-relaxation of some slow modes, which can cause premature suppression of important coherence terms, in particular, within the second order time-local approximation. Variational PQME18,19,21,22 that combines a variational ansatz6,9–11 with PQME and, more recently, frozen mode PQME32 approaches have been developed to address this issue. This work develops a formulation that can put these works18,21,32 into a broader context and can ameliorate the issue of over-relaxation by extending the PQME for partial PTs of a fairly general kind, where high frequency bath modes are transformed preferentially. It is shown here that a general closed form expression of a second order time-local PQME including all the inhomogeneous terms can be derived as in the case of the full PT,15 although the resulting expressions are more complicated. All the terms needed for the calculation of a second order partially polaron-transformed QME (p-PQME) are derived in detail. Then, numerical results are provided for a model with two system states, which demonstrate the feasibility and the utility of the p-PQME approach.

This paper is organized as follows: Sec. II provides the model Hamiltonian and presents the derivation of all terms involved in the second order time-local p-PQME. Section III considers a case of a two-level system and presents the results of model calculations. Section IV summarizes the main results of this paper and offers concluding remarks.

Let us consider a quantum system consisting of N coupled quantum states, |j⟩s, with each representing a localized electronic excitation or charge carrying state. The Hamiltonian of the quantum system consisting of these states, in general, can be expressed as

Hs=j=1NEj|jj|+j,k=1NJjk|jk|,
(1)

where Ej is the energy of the site local state |j⟩ and Jjk, for jk, is the electronic coupling between states |j⟩ and |k⟩, which is assumed to be a real number here. The double summation in the above equation also includes the case j = k, for which Jjk is assumed to be zero.

All other degrees of freedom constituting the total Hamiltonian are referred to here as bath, which include all the vibrational modes of molecules and the polarization response of environmental degrees of freedom. Some of these may exhibit significant anharmonic character in real molecular systems. However, for the sake of simplicity, it is assumed here that all of them can be modeled as coupled harmonic oscillators. In addition, all couplings of the bath to the system Hamiltonian are assumed to be diagonal with respect to the site local states |j⟩s and to be linear in the displacements of the bath oscillators. Thus, the total Hamiltonian is assumed to be of the following standard form:

H=Hs+Hb+Hsb,
(2)

where

Hb=nωnbnbn+12,
(3)
Hsb=j=1Nnωngn,j(bn+bn)|jj|.
(4)

In the above expressions, ωn and bn(bn) are the frequency and the lowering (raising) operator of each normal mode constituting the bath degrees of freedom, and gn,j represents the (dimensionless) coupling strength of each mode to the state |j⟩. The nature of this bath can be characterized collectively by introducing the following bath spectral density:

Jjj(ω)=πnδ(ωωn)ωn2gn,jgn,j.
(5)

The total Hamiltonian, Eq. (2), is a straightforward generalization of the spin-boson Hamiltonian33,34 and has been used widely for molecular excitons35–37 and charge transport dynamics.34,36,38 The full information on the combined system and bath degrees of freedom, in general, requires determining the total density operator ρ(t), which is governed by the following quantum Liouville equation:

ddtρ(t)=iLρ(t)i[H,ρ(t)].
(6)

Due to a large number of bath degrees of freedom typically involved, solving the above equation exactly is difficult even for simple forms of Hb and Hsb given by Eqs. (3) and (4). In the quantum master equation approach, only the reduced system degrees of freedom are solved explicitly, whereas the effects of the bath are treated only to certain extents that are necessary. Typically, these effects can be encoded entirely into appropriate bath spectral densities.

For the case of Ohmic or sub-Ohmic bath spectral density, namely, when Eq. (5) behaves linearly or sub-linearly in the low frequency limit, the sum of Huang–Rhys factors for the bath degrees of freedom diverges.34 This implies vanishing Debye–Waller factors, which result in premature suppression of some coherence terms when the full PT is applied first. This issue is significant, in particular, for charge transfer processes, where the bath spectral density is typically known to contain Ohmic or even sub-Ohmic low frequency components. A simple way to avoid such suppression is to limit the PT to only fast enough bath modes. To this end, let us introduce a weighting function Wh(ω) with the following limiting behavior:

Wh(ω)=O(ωα) for ω0,1 for ω.
(7)

The scaling behavior for ω → 0 in the above equation, for a sufficiently large value of α, can suppress the sluggish components of the bath spectral density and, thus, prevents the corresponding Debye–Waller factor from becoming zero. For the case of Ohmic bath, this means that α ≥ 1 at least. If the bath spectral density has sub-Ohmic components, the lower bound for α should increase accordingly.

Let us now define a generating function of a partial PT as follows:

G=j=1Nngn,jWh(ωn)(bnbn)|jj|.
(8)

The corresponding PT, when applied to the system Hamiltonian, Eq. (1), results in

H̃s=eGHseG=j=1NEj|jj|+j,k=1NJjkθjθk|jk|,
(9)

where

θj=engn,jWh(ωn)(bnbn).
(10)

On the other hand, it is straightforward to show that

eG(Hb+Hsb)eG=Hb+j=1Nnωngn,j(1Wh(ωn))(bn+bn)×|jj|j=1Nnωngn,j2Wh(ωn)×(2Wh(ωn))|jj|.
(11)

Combining the above expression with Eq. (9), one can obtain the following expression for the transformed total Hamiltonian:

H̃=eGHeG=H̃sp+H̃sb+Hb,
(12)

where

H̃sp=j=1NẼj|jj|.
(13)

In the above expression, Ẽj is a partially renormalized energy for site j given by

Ẽj=Ejλj,
(14)

with λj being the corresponding reorganization energy defined as

λj=nωngn,j2Wh(ωn)(2Wh(ωn))=1π0dωJjj(ω)ωWh(ω)(2Wh(ω)).
(15)

In the second equality of the above equation, the definition of the bath spectral density, Eq. (5), was used.

The second term in Eq. (12), H̃sb, is a partially renormalized system–bath interaction Hamiltonian given by

H̃sb=j,k=1NJjkθjθk|jk|+j=1Nnωngn,j(1Wh(ωn))(bn+bn)|jj|,
(16)

where θjθk can be expressed as

θjθk=enδgn,jkWh(ωn)(bnbn),
(17)

with δgn,jk = (gn,jgn,k).

A complete derivation of a QME defined in the partial polaron transformation (p-PT) space, as defined in Sec. II A, is provided below.

1. Quantum Liouville equation in the polaron and interaction picture

The total Hamiltonian in the partial polaron picture, Eq. (12), can be divided into new effective zeroth and first order terms as follows:

H̃=H̃0+H̃1,
(18)

where

H̃0=H̃sp+H̃sbb+HbH̃0,s+Hb,
(19)
H̃1=H̃sbH̃sbb.
(20)

In above expressions, H̃sbb=Trb{H̃sbρb} with ρb=eβHb/Trb{eβHb}. This term represents the average system–bath interaction in the partial polaron picture.

In Eq. (19), the system part of the new zeroth order Hamiltonian, H̃0,s=H̃sp+H̃scb, includes the average effect of system–bath interactions in the partial polaron picture and can be expressed as

H̃0,s=j=1NẼj|jj|+j,k=1NJ̃jk|jk|,
(21)

where Ẽj has been defined by Eq. (14) and J̃jk=wjkJjk with

wjk=θjθk=θkθj=encoth(βωn/2)δgn,jk2Wh(ωn)2/2.
(22)

Unlike the case of the full PT, the Debye–Waller factor wjk given above is non-zero even for the Ohmic bath spectral density given that the weighting function satisfies Eq. (7).

Similarly, the first order term H̃1, Eq. (20), can be expressed as

H̃1=j,k=1NB̃jk|jk|,
(23)

where

B̃jk=Jjk(θjθkwjk)+δjkDj.
(24)

In the above expressions δjk is the Kronecker-delta symbol and

Dj=nωngn,j(1Wh(ωn))(bn+bn).
(25)

Thus, the bath operator B̃jk given by Eq. (24) is a sum of the renormalized system–bath interaction term (relative to its average) due to partial PT (for jk) and of the remaining linear interaction term (for j = k) for portions of bath modes that have not been transformed.

Having defined H̃0, for which exact time evolution can be implemented numerically, let us now consider the dynamics in the interaction picture of H̃0. First, H̃1 in this interaction picture becomes

H̃1,I(t)eiH̃0t/H̃1eiH̃0t/=j,k=1NB̃jk(t)Tjk(t),
(26)

where

Tjk(t)=eiH̃0,st/|jk|eiH̃0,st/,
(27)
B̃jk(t)=eiHbt/B̃jkeiHbt/=Jjk(θj(t)θk(t)wjk)+δjkDj(t).
(28)

In the above expression,

θj(t)θk(t)=enδgn,jkWh(ωn)(bneiωntbneiωnt),
(29)
Dj(t)=nωngn,j(1Wh(ωn))(bneiωnt+bneiωnt).
(30)

In the interaction picture with respect to H̃0, the partially polaron-transformed total density operator becomes ρ̃I(t)=eiL̃0tρ̃(t), which evolves according to the following time dependent quantum Liouville equation:

ddtρ̃I(t)=iL̃1,I(t)ρ̃I(t)i[H̃1,I(t),ρ̃I(t)],
(31)

where the second equality serves as the definition of L̃1,I(t).

2. Quantum master equation for the reduced density operator

Taking trace of ρ̃I(t) over the bath degrees of freedom leads to the following interaction-picture reduced system density operator defined in the p-PT system–bath space:

σ̃I(t)Trbρ̃I(t).
(32)

While the above reduced density operator still retains full information on the system degrees of freedom in the p-PT space, it is important to note that the trace operation makes it impossible to retrieve the full information on the system prior to the application of p-PT. On the other hand, properties diagonal in the site basis, which are not affected by p-PT, remain intact.

A formally exact time evolution equation with time-convolution can be obtained for σ̃I(t) employing the standard projection operator technique37,39 for a well-known projection operator P()ρbTrb{()} as follows:

ddtσ̃I(t)=0tdτTrbL̃1,I(t)e(+)iτtdτQL̃1,I(τ)QL̃1,I(τ)ρb×σ̃I(τ)iTrbL̃1,I(t)e(+)i0tdτQL̃1,I(τ)Qρ̃(0),
(33)

where e(+) denotes the exponential operator with chronological time ordering, Q=1P and

Qρ̃(0)=eGρ(0)eGρbTrb{eGρ(0)eG}.
(34)

Alternatively, replacing ρ̃I(τ) with the backpropagation of ρ̃I(t) from t to τ within the projection operator formalism,37,40 one can obtain the following formally exact time-local equation:

ddtσ̃I(t)=0tdτTrbL̃1,I(t)(1+iΓ1,I(t))1e(+)iτtdτQL̃1,I(τ)×QL̃1,I(τ)Pe()iτtdτL̃1,I(τ)ρbσ̃I(t)iPL̃1,I(t)(1+iΓ1,I(t))1e(+)i0tdτQL̃1,I(τ)Qρ̃(0),
(35)

where e(−) denotes the exponential operator with anti-chronological time ordering and

Γ1,I(t)=0tdτe(+)iτtdτQL̃1,I(τ)QL̃1,I(τ)Pe()iτtdτL̃1,I(τ).
(36)

When approximated up to the second order, it is straightforward to show that Eq. (35) simplifies to

ddtσ̃I(t)=R(t)σ̃I(t)+I(t),
(37)

where

R(t)=0tdτTrb{L̃1,I(t)L̃1,I(τ)ρb},
(38)
I(t)=I(1)(t)+I(2)(t)=iTrb{L̃1,I(t)Qρ̃(0)}0tdτTrb{L̃1,I(t)L̃1,I(τ)Qρ̃(0)}.
(39)

In the above expression, I(1)(t) and I(2)(t) represent the first and second order inhomogeneous terms of the time evolution equation. Note that Eq. (37) can also be obtained from Eq. (33), by simply replacing σ̃I(τ) with σ̃I(t), which does not affect the accuracy at the second order level.

Equation (37) is the second order time local p-PQME expressed in Liouville space. In all previous studies and in the present paper, this time local form has been chosen due to its convenience. However, a time nonlocal second order expression can also be derived directly from Eq. (33), and its performance compared with the time-local form needs to be understood better through actual numerical studies. Many numerical tests so far seem to indicate that the performance of the time-local form is better than that of the time non-local form for exciton and charge transfer dynamics near room temperature. However, considering that the second order time-nonlocal PQME is equivalent to the non-interacting blip approximation,9,41–43 which accounts for a significant contribution of coherent dynamics even for the Ohmic bath, it is likely that the time non-local equation becomes more reliable as the bath becomes sluggish. In addition, a recent work44 also provides examples of the case where the performance of time-nonlocal QME is more satisfactory than the time-local form. Therefore, further tests and comparative calculations are necessary to make a more comprehensive assessment of the two approaches. With this point clear, the rest of this section provides detailed expressions for the relaxation superoperator, Eq. (38), and the inhomogeneous term, Eq. (39), in the Hilbert space.

3. Homogeneous terms of the second order time local p-PQME

The Hilbert space expression for R(t)σ̃I(t) in Eq. (37) can be obtained by employing Eqs. (26)(28) in Eq. (31) and by taking advantage of the cyclic invariance of the trace operation with respect to the bath degrees of freedom. The resulting expression is as follows:

R(t)σ̃I(t)=12j,k=1Nj,k=1N0tdτB̃jk(t)B̃jk(τ)×[Tjk(t),Tjk(τ)σ̃I(t)]+H.c.,
(40)

where H.c refers to Hermitian conjugates of all previous terms and B̃jk(t)B̃jk(τ) (with subscript b omitted) represents averaging over the equilibrium bath density operator, ρb.

 Appendix A describes calculation of all the terms constituting B̃jk(t)B̃jk(τ). When the resulting expressions, Eqs. (A2), (A15), (A21), and (A22), are used in Eq. (A1), it can be expressed as

B̃jk(t)B̃jk(τ)=J̃jkJ̃jkeKjk,jk(tτ)1+δjkJ̃jkMj,jk(tτ)+δjkJ̃jkMj,kj(tτ)+δjkδjkCjj(tτ),
(41)

where

Kjk,jk(t)=nδgn,jkδgn,jkWh(ωn)2×cothβωn2cos(ωnt)isin(ωnt),
(42)
Mj,jk(t)=nωngn,jδgn,jk(1Wh(ωn))Wh(ωn)×cos(ωnt)icothβωn2sin(ωnt),
(43)
Cjj(t)=n2ωn2gn,jgn,j(1Wh(ωn))2×cothβωn2cos(ωnt)isin(ωnt).
(44)

Similar expressions as above have also been derived in the context of variational PQME.21 Note that the three bath correlation functions defined above satisfy the following symmetry properties:

Kjk,jk(t)=Kkj,kj(t)=Kjk,kj(t)=Kkj,jk(t),
(45)
Mj,jk(t)=Mj,kj(t),
(46)
Cjj(t)=Cjj(t).
(47)

The three correlation functions, Eqs. (42)(44), can all be expressed in terms of the bath spectral density, Eq. (5). For more compact expressions of these, let us introduce the following auxiliary bath spectral densities,

Jj,jk(1)(ω)=πnδ(ωωn)ωn2gn,jδgn,jk=Jjj(ω)Jjk(ω),
(48)
Jjk,jk(2)(ω)=πnδ(ωωn)ωn2δgn,jkδgn,jk=Jjj(ω)+Jkk(ω)Jjk(ω)Jkj(ω).
(49)

Then, it is straightforward to show that Eqs. (42)(44) can be expressed as

Kjk,jk(t)=1π0dωJjk,jk(2)(ω)ω2Wh(ω)2×cothβω2cos(ωt)isin(ωt),
(50)
Mj,jk(t)=1π0dωJj,jk(1)(ω)ω(1Wh(ω))Wh(ω)×cos(ωt)icothβω2sin(ωt),
(51)
Cjj(t)=π0dωJjj(ω)(1Wh(ω))2×cothβω2cos(ωt)isin(ωt).
(52)

Note also that wjk defined by Eq. (22) can be expressed as follows:

wjk=eKjk,jk(0)/2=exp12π0dωJjk,jk(2)(ω)ω2Wh(ω)2cothβω2.
(53)

If Wh(ω) = 1, Mj,jk(t) and Cjj(t) become zero and the expressions for Kjk,jk(t) and wjk reduce to those for the original second order PQME based on the full PT. On the other hand, for Wh(ω) = 0, Kjk,jk(t) and Cjj(t) become zero and the expression for Cjj(t) reduce to that for a conventional second order time-local QME (without PT). In this sense, the above expressions can be viewed as general ones incorporating the two limiting cases. As the next simplest case, let us consider the case where Wh(ω) is a step function. For this case, Wh(ω)(1 − Wh(ω)) = 0 and, thus, Mj,jk(t) becomes zero. As a result, the relaxation superoperator becomes a simple sum of those due to the p-PT part and untransformed linear coupling terms.

4. Inhomogeneous terms of p-PQME

For the calculation of inhomogeneous terms, let us assume that the initial untransformed total density operator is given by

ρ(0)=σ(0)ρb,
(54)

where

σ(0)=j,k=1Nσjk(0)|jk|.
(55)

Then,

Qρ̃(0)=j,k=1Nσjk(0)Tjk(0)δρ̃b,jk,
(56)

where Eq. (27) with t = 0 has been used and

δρ̃b,jk=θjρbθkwjkρb,
(57)

with the convention that wjk = 1 for j = k. Therefore, the first order inhomogeneous term in Eq. (39) can be expressed as

I(1)(t)=iTrb{L̃1,I(t)Qρ̃(0)}=ij,k=1Nj,k=1NTrbB̃jk(t)δρ̃b,jkσjk(0)[Tjk(t),Tjk(0)].
(58)

In the above expression, the trace over the bath can be calculated explicitly employing the definitions for B̃jk(t) and δρ̃b,jk, Eqs. (28) and (57), respectively. Details of this calculation are provided in  Appendix B, and the resulting expression can be summed up as

Trb{B̃jk(t)δρ̃b,jk}=wjkJ̃jkeKjk,jk(t)fjk,k(t)1+δjkMj,jk(t)+hj,k(t),
(59)

where

fjk,k(t)=exp2ingn,kδgn,jkWh(ωn)2sin(ωnt)=exp2iπ0dωJk,jk(1)(ω)ω2Wh(ω)2sin(ωt),
(60)
hj,k(t)=2nωngn,jgn,k(1Wh(ωn))Wh(ωn)cos(ωnt)=2π0dωJjk(ω)ω(1Wh(ω))Wh(ω)cos(ωt).
(61)

Similarly, the second order inhomogeneous term in Eq. (39) can be expressed as

I(2)(t)=0tdτTrbL̃1,I(t)L̃1,I(τ)Qρ̃(0)=12j,k=1Nj,k=1Nj,k=1N×0tdτTrbB̃jk(t)B̃jk(τ)δρ̃b,jk×σjk(0)[Tjk(t),Tjk(τ)Tjk(0)]+H.c.,
(62)

where H.c. refers to Hermitian conjugates of all previous terms. Detailed expressions for the trace of bath operators in the above expression are derived in  Appendix C.

As in the case of the relaxation superoperator, the expressions for the inhomogeneous terms shown above reduce to those for the second order PQME for Wh(ω) = 1 and those for the regular second order QME for Wh(ω) = 0. For the case where Wh(ω)(1 − Wh(ω)) = 0, they become independent sums of those due to PT and due to untransformed linear system–bath couplings.

For both better conceptual understanding and efficient numerical calculation, it is convenient to consider the dynamics in the basis of eigenstates of H̃0,s. Let us denote the pth eigenstate and eigenvalue of H̃0,s as |φp⟩ and Ep. Then,

H̃0,s=p=1NEp|φpφp|.
(63)

The transformation matrix U between the site localized states |j⟩s and the eigenstates |φp⟩s can be defined such that Ujp = ⟨j|φp⟩. Then,

|j=p=1NUjp*|φp.
(64)

This transformation can be used to express Tjk(t) defined by Eq. (27) in the basis of |φp⟩s as follows:

Tjk(t)=p,q=1NUjp*UkqeiδEpqt/|φpφq|,
(65)

where δEpq=EpEq. Let us also introduce Spq(t)s such that

σ̃I(t)=p,q=1NSpq(t)|φpφq|.
(66)

Then, with some arrangement of dummy summation indices, it is straightforward to show that

[Tjk(t),Tjk(τ)σ̃I(t)]=p,q=1N|φpφq|p,q=1Nδqqr=1NUjp*UkrUjrUkpeiδEppt/eiδEpr(tτ)/Ujq*UkqUjp*Ukpei(δEppδEqq)t/eiδEpp(tτ)/Spq(t).
(67)

The above expression, in combination with Eq. (41), can be used to express R(t)σ̃I(t) in the basis of |φp⟩s. For a more compact expression, let us introduce

Wjk,jkpq(t)=J̃jkJ̃jk0tdτeiδEpq(tτ)/eKjk,jk(tτ)1,
(68)
Yj,jkpq(t)=J̃jk0tdτeiδEpq(tτ)/Mj,jk(tτ),
(69)
Xjjpq(t)=0tdτeiδEpq(tτ)/Cjj(tτ).
(70)

Then,

R(t)σ̃I(t)=p,q=1Np,q=1N|φpφq|Rpqpq(t)Spq(t),
(71)

where

Rpqpq(t)=12j,k=1Nj,k=1Nδqqr=1NUjp*UkrUjr*UkpeiδEppt/Wjk,jkpr(t)+δjkYj,jkpr(t)+δjkYj,kjpr(t)+δjkδjkXjjpr(t)Ujq*UkqUjp*Ukpei(δEppδEqq)t/Wjk,jkpp(t)+δjkYj,jkpp(t)+δjkYj,kjpp(t)+δjkδjkXjjpp(t)+[c.c.,pq,pq].
(72)

In the above expression, the last line represents complex conjugates along with the interchange of indices, pq and p′ ↔ q′, of all previous terms.

For the calculation of the first order inhomogeneous term, the commutator of system operators in Eq. (58) can be calculated in a manner similar to Eq. (67). The resulting expression is

[Tjk(t),Tjk(0)]=p,q=1N|φqφq|rUjp*UkrUjr*UkqeiδEprt/Ujr*UkqUjp*UkreiδErqt/,
(73)

which can also be obtained from Eq. (67) by replacing Spq(t) with δpq and assuming τ = 0. Combining the above expression with Eq. (59), one can express the first order inhomogeneous term, Eq. (58), as follows:

I(1)(t)=ip,q=1N|φpφq|r=1Nj,k=1Nj,k=1Nσjk(0)wjk×Ujp*UkrUjr*UkqeiδEprt/Ujr*UkqUjp*UkreiδErqt/×J̃jk(eKjk,jk(t)fjk,k(t)1)+δjkMj,jk(t)+hj,k(t).
(74)

Similar expressions for the second order inhomogeneous term, Eq. (62), can be obtained by replacing Spq(t) in Eq. (67) with Ujp*Ukq. The resulting expression is as follows:

[Tjk(t),Tjk(τ)Tjk(0)]=p,q=1N|φpφq|p,q=1Nδqqr=1NUjp*UkrUjrUkpeiδEppt/eiδEpr(tτ)/Ujq*UkqUjp*Ukpei(δEppδEqq)t/eiδEpp(tτ)/Ujp*Ukq.
(75)

Combining the above expression with bath correlation functions introduced in  Appendix C, one can express the second order inhomogeneous term as follows:

I(2)(t)=j,k=1Nj,k=1Nj,k=1Nσjk(0)p,q=1N|φpφq|p,q=1Nr=1NUjp*UkqeiδEppt/δqqUjp*UkrUjrUkpδrpUjq*UkqUjp*UkpeiδEqqt/×JjkJjkF̃jk,jkjk(t;δEpr)+δjkJjkH̃j,jk(1),jk(t;δEpr)+JjkδjkH̃jk,j(2),jk(t;δEpr)+δjkδjkL̃j,jjk(t;δEpr)+H.c.,
(76)

where the definitions of Eqs. (C28)(C31) in  Appendix C have been used.

Let us consider the simplest case where there are two site local system states (N = 2), with only one electronic coupling J = J12 = J21 and only one Debye–Waller factor, w = w12 = w21. Then, introducing θ=θ1θ2 for the present case, the renormalized zeroth order system Hamiltonian and the first order term of the Hamiltonian defined by Eq. (23) reduce to

H̃0,s=Ẽ1|11|+Ẽ2|22|+Jw(|12|+|21|),
(77)
H̃1=D1|11|+D2|22|+J(θw)|12|+J(θw)|21|,
(78)

where w = ⟨θ⟩ = ⟨θ⟩ and D1 and D2 have been defined by Eq. (25). Note also that, for the present case,

J̃12=J̃21=wJ,
(79)

and J̃11=J̃22=0 by definition. The eigenvalues of H̃0,s are given by

E1,2=Ẽ1+Ẽ22±Ẽ1Ẽ22sec(2ξ),
(80)

with subscripts 1 and 2 on the left hand side denoting and denoting the + sign, respectively − signs on the right-hand side, and ξ=tan1(2Jw/(Ẽ1Ẽ2))/2. The corresponding eigenstates are given by

|φ1=cosξ|1+sinξ|2,
(81)
|φ2=sinξ|1+cosξ|2.
(82)

Thus, for the present case, U11 = U22 = cos ξ and U21 = −U12 = sin ξ.

Let us also assume that the spectral densities for sites 1 and 2 are the same, which we denote as J(ω). Namely, J11(ω)=J22(ω)=J(ω). In addition, let us also assume that J12(ω)=J21(ω)=0. Then, J1,12(1)(ω)=J2,21(1)(ω)=J(ω) and J12,12(2)(ω)=J21,21(2)(ω)=2J(ω). For all other indices, Jj,jk(1)(ω)=0 and Jjk,jk(2)(ω)=0.

Then, all the bath correlation functions that enter the p-PQME can be represented by the following three functions:

K(t)K12,12(t)=2π0dωJ(ω)ω2Wh(ω)2×cothβω2cos(ωt)isin(ωt)=K21,21(t)=K12,21(t)=K21,12(t),
(83)
M(t)M1,12(t)=1π0dωJ(ω)ωWh(ω)(1Wh(ω))×cos(ωt)icothβω2sin(ωt)=M2,21(t)=M1,21(t)=M2,12(t),
(84)
C(t)C11(t)=π0dωJ(ω)(1Wh(ω))2×cothβω2cos(ωt)isin(ωt)=C22(t).
(85)

For all other components, Kjk,jk(ω)=0, Mj,jk(ω)=0, and Cjk(ω)=0.

In the site basis, Eq. (40) for the present case can be expressed as follows:

R(t)σ̃I(t)=0tdτC(tτ)[T11(t),T11(τ)σ̃I(t)]+[T22(t),T22(τ)σ̃I(t)]+JwM(tτ)[ΔTd(t),ΔTc(τ)σ̃I(t)][ΔTc(t),ΔTd(τ)σ̃I(t)]+J2w2eK(tτ)1[T12(t),T12(τ)σ̃I(t)]+[T21(t),T21(τ)σ̃I(t)]+J2w2eK(tτ)1[T12(t),T21(τ)σ̃I(t)]+[T21(t),T12(τ)σ̃I(t)]+H.c.,
(86)

where

ΔTd(t)=T22(t)T11(t),
(87)
ΔTc(t)=T21(t)T12(t).
(88)

Equation (86) above provides clear insights into the effects of p-PT and is useful for understanding the steady state behavior of the population dynamics. However, for general numerical calculation, it is convenient to use expressions defined in the basis of eigenstates of H̃0,s, as detailed in Sec. II C. For this, Rpqpq(t) given by Eq. (72) needs to be calculated, which in turn requires calculation of Wjk,jkpq(t), Yj,jkpq(t), and Xjjpq(t) defined by Eqs. (68)(70). For the present model of two state systems coupled to independent baths, most of these are zero except for a few functions, as defined below.

First, all of nonzero Wjk,jkpq(t)’s defined by Eq. (68) for the present case reduce to one of the following two functions:

Wpq(t)W12,12pq(t)=W21,21pq(t)=J2w20tdτeiδEpq(tτ)/eK(tτ)1,
(89)
W+pq(t)W12,21pq(t)=W21,12pq(t)=J2w20tdτeiδEpq(tτ)/eK(tτ)1.
(90)

All other terms of Wjk,jkpq(t) are zero by definition. Similarly, all of nonzero Yj,jkpq(t)’s and Xjjpq(t)’s can be specified by

Ypq(t)Y1,12pq(t)=Y2,21pq(t)=Jw0tdτeiδEpq(tτ)/M(tτ)=Y1,21pq(t)=Y2,12pq(t),
(91)
Xpq(t)X11pq(t)=X22pq(t)=0tdτeiδEpq(tτ)/C(tτ).
(92)

All other terms with indices different from the above are zero. Thus, with W±pq(t), Ypq(t), and Xpq(t) as defined above, all of Rpqpq(t) constituting Eq. (72) for the present case can be calculated.

The first order inhomogeneous term, Eq. (74), involves additional functions fjk,k(t) and hj,k(t), which can also be simplified for the present model. Let us define

f(t)exp2iπ0dωJ(ω)ω2Wh(ω)2sin(ωt)=f12,1(t).
(93)

Then, it is easy to show that f21,2(t) = f(t), f21,1(t) = f12,2(t) = f*(t). On the other hand, f11,1(t) = f22,1(t) = f11,2(t) = f22,2(t) = 1. For hj,k(t), only specification of the following function is necessary:

h(t)2π0dωJ(ω)ω(1Wh(ω))Wh(ω)cos(ωt)=h1,1(t)=h2,2(t).
(94)

For other cases, h1,2(t) = h2,1(t) = 0.

The second order inhomogeneous term, Eq. (76), consists of much more terms that involve time integration of four different types of time correlation functions as described in  Appendix C. Although the calculation of all these expressions is straightforward, actual numerical implementation is nontrivial and will be the subject of future work.

Numerical calculations neglecting the contribution of inhomogeneous terms are provided here. The objective of these calculations is to demonstrate the feasibility of using the p-PQME expressions derived in previous subsections and to investigate the effects of the weighting function Wh(ω). Thus, the results to be presented do not yet have the quantitative accuracy for all time due to the neglect of inhomogeneous terms. However, they still provide important qualitative information and reliable steady state limits for the cases where the inhomogeneous terms vanish.

First, it is useful to further simplify the expressions for Rpqpq(t)’s. Summing up all possible 16 cases of j, k, j′, k′ = 1, 2 in Eq. (72), it is straightforward to show that it can be simplified as follows:

Rpqpq(t)=12δqqeiδEppt/r=12AprrpXpr(t)+(Bpr(1)Brp(2)Bpr(2)Brp(1))Ypr(t)+Cprrp(1)Wpr(t)+Cprrp(2)W+pr(t)ei(δEppδEqq)t/×AqqppXpp(t)+(Bqq(1)Bpp(2)Bqq(2)Bpp(1))Ypp(t)+Cqqpp(1)Wpp(t)+Cqqpp(2)W+pp(t)+[C.C.,pq,pq],
(95)

where

Apqrs=U1pU1qU1rU1s+U2pU2qU2rU2s,
(96)
Bpq(1)=U1pU1qU2pU2q,
(97)
Bpq(2)=U1pU2qU2pU1q,
(98)
Cpqrs(1)=U1pU2qU1rU2s+U2pU1qU2rU1s,
(99)
Cpqrs(2)=U1pU2qU2rU1s+U2pU1qU1rU2s.
(100)

The bath spectral density being considered is the Ohmic form with the exponential cutoff as follows:

J(ω)=πηωeω/ωc.
(101)

It is assumed that the two parameters of the above spectral density are set to η = 1 and ℏωc = 200 cm−1. In addition, the temperature and electronic coupling are set to T = 300 K and J = 300 cm−1. For this choice of parameters, the bath dynamics, system electronic coupling, and thermal energy are all comparable. Thus, the dynamics are expected to be the border-line case between incoherent and coherent quantum dynamics. Two cases of relative site energies, E1 = E2 and E1E2 = 200 cm−1 will be considered.

There are infinite number of possible choices available for Wh(ω) that satisfies the requirement of Eq. (7). One simple but flexible choice is the following function known as (cumulative) Weibull distribution45 or the complement of a stretched exponential function,

Wh(ω)=1e(ω/ωh)α.
(102)

For small ω, Wh(ω)(ω/ωh)α. For large α, Wh(ω) approaches the step function at ω = ωh.

While choosing a value α in Eq. (102) such that α ≥ 1 is sufficient for ensuring that the Debye–Waller factor does not vanish, it is not clear whether it also guarantees a well-behaving time evolution equation. In order to check this, it is important to examine how the three time correlation functions defined by Eqs. (83)(85) behave for different choices of α. Figure 1 shows the real and imaginary parts of K(t) defined by Eq. (83) for two different values of α = 1 and 2 with the choice of ωh = ωc. In both cases, the real and imaginary parts decay to zero quickly enough to make both Wpq(t) and W+pq(t), defined, respectively, by Eqs. (89) and (90), converge to finite values.

FIG. 1.

Real (upper panel) and imaginary (Lower panel) parts of K(t) for ωh/ωc = 1 and for two cases of α = 1 and 2.

FIG. 1.

Real (upper panel) and imaginary (Lower panel) parts of K(t) for ωh/ωc = 1 and for two cases of α = 1 and 2.

Close modal

On the other hand, for the case of M(t), it turns out that the choice of α = 1 does not result in a stable time evolution equation, in general. Figure 2 shows the real and imaginary parts of M(t), also for α = 1 and 2 with the choice of ωh = ωc. For α = 1, while the real part decays to zero quickly, the imaginary part is seen to decay very slowly. In fact, due to the slow decay, the Ypq(t) defined by Eq. (91) diverges, in general, in this case. Test calculations for other values of α show that such divergence persists up to α = 3/2. On the other hand, for the case of α = 2 shown in Fig. 2, it is clear that the imaginary part decays to zero quickly, ensuring for Ypq(t) to converge to a finite value in the steady state limit.

FIG. 2.

Real (upper panel) and imaginary (lower panel) parts of M(t)/(ωc) for ωh/ωc = 1 and for two cases of α = 1 and 2.

FIG. 2.

Real (upper panel) and imaginary (lower panel) parts of M(t)/(ωc) for ωh/ωc = 1 and for two cases of α = 1 and 2.

Close modal

For the case of C(t), as shown in Fig. 3, both real and imaginary parts decay to zero quickly already for α = 1, resulting in well behaving Xpq(t). Summing up the results shown in Figs. 13, while the choice of α = 1 is acceptable as far as the Deby–Waller factor, K(t), and C(t) are concerned, it is not appropriate due to resulting slow decay of M(t). On the other hand, the choice of α = 2 results in all well-behaving and convergent functions that constitute the relaxation operator. This is also true for all the terms involved in the inhomogeneous terms as well.

FIG. 3.

Real (upper panel) and imaginary (lower panel) parts of C(t)/(2ωc2) for ωh/ωc = 1 and for two cases of α = 1 and 2.

FIG. 3.

Real (upper panel) and imaginary (lower panel) parts of C(t)/(2ωc2) for ωh/ωc = 1 and for two cases of α = 1 and 2.

Close modal

Figure 4 shows the results for E1 = E2 for different values of ωh with the choice of α = 2, where Wh(ω) becomes a complement of a Gaussian function. The result for the smallest value of ωh among those shown (ωh/ωc = 0.5) is close to the limit of full PQME, without coherence, whereas that for the largest value of ωh among those shown (ωh/ωc = 4) is close to the full second order time-local QME, which has maximum coherence. For ωh/ωc = 1, the time dependent population exhibits an intermediate character between the two limits.

FIG. 4.

Populations of the donor (site 1) for different values of ωh for E1 = E2 and α = 2 in Wh(ω), Eq. (102). Other parameters of the model are as follows: η = 1, ωc = 200 cm−1, T = 300 K, and J = 300 cm−1.

FIG. 4.

Populations of the donor (site 1) for different values of ωh for E1 = E2 and α = 2 in Wh(ω), Eq. (102). Other parameters of the model are as follows: η = 1, ωc = 200 cm−1, T = 300 K, and J = 300 cm−1.

Close modal

Figure 5 shows the results for an asymmetric case, where E1E2 = 200 cm−1. All other parameters, including α = 2, remain the same as those for Fig. 4. In this case, the steady state limits of the population as well as the coherence pattern vary with ωh. The variation of the steady state limit with ωh reflects a different extent of system–bath entanglement depending on the extent of polaron transformation. The smaller the value of ωh, the closer the steady states are to the original localized states 1 and 2, for which the energy gap becomes the maximum.

FIG. 5.

Populations of the donor (site 1) for E1E2 = 200 cm−1 and α = 2 in Wh(ω), Eq. (102). All other parameters are the same as in Fig. 4.

FIG. 5.

Populations of the donor (site 1) for E1E2 = 200 cm−1 and α = 2 in Wh(ω), Eq. (102). All other parameters are the same as in Fig. 4.

Close modal

This work has provided a general framework to overcome a known issue of the original second order PQME, namely, premature over-relaxation of the sluggish bath, by deriving full expressions for the second order time-local p-PQME. The main results of this work, represented by Eqs. (72), (74), and (76), can be applied for any kind of bath spectral densities and initial system states but will be particularly useful for the cases where the bath spectral densities are Ohmic or sub-Ohmic. It is important to note that the expressions provided here are applicable even to the case where the same bath mode is partly transformed, with the remaining part untransformed, and is, thus, more general than the case where the bath is divided into two disjoint transformed and untransformed groups.

Numerical tests for a simple two level system coupled to an Ohmic bath demonstrate that appropriate specification of the weighting function Wh(ω) can tune the extent of coherence and the extent of system–bath entanglement in the steady state limit. This adds a new dimension of flexibility in incorporating the PT approach into a QME calculation. The flexibility in choosing the weighting function Wh(ω) in all the expressions derived for the p-PQME presented here leaves open various possibilities of adapting or improving the methodology. For example, the variational theorem can be used for its optimization. Alternatively, benchmarking against numerically exact computational results followed by a machine learning based optimization can potentially lead to an optimized second order p-PQME that can best approximate exact dynamics. To this end, full calculations including all the inhomogeneous terms and benchmarking against a broad range of exact numerical results, the subject of forthcoming work, will be necessary.

The formulation developed here also will be useful for further extension of PT based QME approaches. For example, extension to the cases with time dependent Hamiltonian for driven quantum systems is straightforward. The formulations and theoretical identities employed in this work will also be useful for the development of new PT based approaches for general anharmonic bath and for the formulation of time dependent PT approach.

This work was mainly supported by the National Science Foundation (Grant No. CHE-1900170). The author also acknowledges partial support from the U.S. Department of Energy, Office of Sciences, Office of Basic Energy Sciences (Grant No. DE-SC0021413), and support from the Korea Institute for Advanced Study (KIAS) through its KIAS Scholar program.

The author has no conflicts to disclose.

Seogjoo J. Jang: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Project administration (lead); Resources (lead); Software (lead); Supervision (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead).

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

The trace of B̃jk(t)B̃jk(τ) with ρb in Eq. (40), namely, its thermal average is expressed as follows:

B̃jk(t)B̃jk(τ)=θj(t)θk(t)θj(τ)θk(τ)wjkθj(τ)θk(τ)wjkθj(t)θk(t)+wjkwjk+δjkDj(t)θj(τ)θk(τ)Dj(t)wjk+δjkθj(t)θk(t)Dj(τ)wjkDj(τ)+δjkδjkDj(t)Dj(τ)=θj(t)θk(t)θj(τ)θk(τ)wjkwjk+δjkDj(t)θj(τ)θk(τ)+δjkθj(t)θk(t)Dj(τ)+δjkδjkDj(t)Dj(τ).
(A1)

In obtaining the second equality of the above equation, the identities θj(τ)θk(τ)=wjk, θj(t)θk(t)=wjk, and ⟨Dj(t)⟩ = ⟨Dj(τ)⟩ = 0 have been used.

The first term in the second equality of Eq. (A1) has the same form as that for the full PT15 except for the additional factor Wh(ωn) multiplied with δgn,jk and δgn,jk, respectively. Thus, following the same procedure as in previous work,15 it can be shown to be

θj(t)θk(t)θj(τ)θk(τ)=wjkwjkeKjk,jk(tτ),
(A2)

where eKjk,jk(tτ) in the exponent is defined by Eq. (42).

The third term in the second equality of Eq. (A1) can be expressed as

Dj(t)θj(τ)θk(τ)=nωngn,j(1Wh(ωn))(bneiωnt+bneiωnt)×neδgn,jkWh(ωn)(bneiωnτbneiωnτ).
(A3)

In the above expression, the product of bath modes (indexed by n′) can be averaged independently for n′ ≠ n, resulting in

nneδgn,jkWh(ωn)(bneiωnτbneiωnτ)=nnecoth(βωn/2)δgn,jk2Wh(ωn)2/2.
(A4)

On the other hand, the term for n′ = n in Eq. (A3) needs to be calculated together with the linear term as follows:

(bneiωnt+bneiωnt)eδgn,jkWh(ωn)(bneiωnτbneiωnτ)=eiωntbneγ*bnγbn+eiωntbneγ*bnγbn,
(A5)

where γ is defined by

γ=δgn,jkWh(ωn)eiωnτ,
(A6)

and γ* is the complex conjugate of γ. Using the fact that eγ*bnγbn=eγbneγ*bne|γ|2/2 and the following identity,

bneγbneγ*bn=γeγbneγ*bn=γ*1eβωneγγ*/(1eβωn),
(A7)

it is straightforward to show that

bneγ*bnγbn=γ*1eβωnecoth(βωn/2)|γ|2/2.
(A8)

Similarly, the bath average in the second term of Eq. (A5) can be expressed as

bneγ*bnγbn=bneγbneγ*bne|γ|2/2.
(A9)

Using the following identity,

bneγbn=eγbnbn+γeγbn,
(A10)

the average over the bath in Eq. (A9) can be expressed as

bneγbneγ*bn=eγbnbneγ*bn+γeγbneγ*bn.
(A11)

Now, employing the following identity:

eγbnbneγ*bn=γ*eγbneγ*bn=γ1eβωneγγ*/(1eβωn),
(A12)

in Eq. (A11) and combining the resulting expression with Eq. (A9), we obtain the following expression:

bneγ*bnγbn=γeβωn1eβωnecoth(βωn/2)|γ|2/2.
(A13)

Combining Eqs. (A8) and (A13) with the definition of Eq. (A6) leads to

(bneiωnt+bneiωnt)eδgn,jkWh(ωn)(bneiωnτbneiωnτ)=δgn,jkWh(ωn)ecoth(βωn/2)δgn,jk2Wh(ωn)2/2×cos(ωn(tτ))icothβωn2sin(ωn(tτ)).
(A14)

Employing this identity in Eq. (A3), one can find that

Dj(t)θj(τ)θk(τ)=wjknωngn,j(1Wh(ωn))δgn,jkWh(ωn)×cos(ωn(tτ))icothβωn2sin(ωn(tτ)).
(A15)

The fourth term in the second equality of Eq. (A1) can be expressed as

θj(t)θk(t)Dj(τ)=nωngn,j(1Wh(ωn))×neδgn,jkWh(ωn)(bneiωntbneiωnt)×(bneiωnτ+bneiωnτ).
(A16)

The bath average term in the above expression can be calculated in a manner similar to the third term of Eq. (A1), which has been described above, but with a different definition of γ=δgn,jkWh(ωn)eiωnt. For this, eγ*bnγbnbn and eγ*bnγbnbn need to be calculated. For the first of these terms, the following identity can be used:

eγ*bnbn=(bnγ*)eγ*bn.
(A17)

Thus,

eγ*bnγbnbn=eγbneγ*bnbne|γ|2/2=eγbnbneγ*bnγ*eγbne|γ|2/2=bneγbneγ*bnγ*eγbneγbne|γ|2/2.
(A18)

Employing Eq. (A7), one can simplify the above expression as follows:

eγ*bnγbnbn=γ*eβωn1eβωne|γ|2coth(βωn/2).
(A19)

Combining the above identity with Eq. (A12) followed by further calculation leads to the following expression:

eδgn,jkWh(ωn)(bneiωntbneiωnt)(bneiωnτ+bneiωnτ)=δgn,kjWh(ωn)ecoth(βωn/2)δgn,jk2Wh(ωn)2/2×cos(ωn(tτ))icothβωn2sin(ωn(tτ)),
(A20)

where note the use of the fact that δgn,kj = −δgn,jk on the right-hand side of the above equation. As a result, Eq. (A16) can be expressed as

θj(t)θk(t)Dj(τ)=wjknωngn,j(1Wh(ωn))δgn,kjWh(ωn)×cos(ωn(tτ))icothβωn2sin(ωn(tτ)).
(A21)

The calculation of the last term in the second equality of Eq. (A1) is straightforward. The resulting expression is as follows:

Dj(t)Dj(τ)=n2ωn2gn,jgn,j(1Wh(ωn))2×(bnbneiωn(tτ)+bnbneiωn(tτ))=n2ωn2gn,jgn,j(1Wh(ωn))2×cothβωn2cos(ωn(tτ))isin(ωn(tτ)).
(A22)

In the second equality of the above equation, the fact that bnbn=bnbn1=eβωn/(1eβωn) has been used.

The bath portion of the first order inhomogeneous term, Eq. (58), can be expressed as follows:

Trb{B̃jk(t)δρ̃b,jk}=JjkTrbθkθj(t)θk(t)θjρbJjkwjkwjk+δjkTrbθkDj(t)θjρb.
(B1)

In the first term on the right-hand side of the above expression, the product of the first three operators within the trace operation can be expressed as

θkθj(t)θk(t)=θkθj(t)θk(t)θkθk=enδgn,jkWh(ωn)(θkbnθkeiωntθkbnθkeiωnt)θk.
(B2)

In the exponent on the right-hand side of the above expression,

θkbnθk=bnWh(ωn)gn,k[bnbn,bn]=bn+Wh(ωn)gn,k,
(B3)
θkbnθk=bnWh(ωn)gn,k[bnbn,bn]=bn+Wh(ωn)gn,k.
(B4)

Therefore,

nδgn,jkWh(ωn)(θkbnθkeiωntθkbnθkeiωnt)=nδgn,jkWh(ωn)(bneiωntbneiωnt)+2inδgn,jkgn,kWh(ωn)2sin(ωnt).
(B5)

Employing the above identity and using the definition of Eq. (60), one can show that Eq. (B2) can be expressed as

θkθj(t)θk(t)=θkθj(t)θk(t)θkθk=enδgn,jkWh(ωn)(θkbnθkeiωntθkbnθkeiωnt)θk=fjk,k(t)θj(t)θk(t)θk.
(B6)

The above identity implies that the first term of Eq. (B1) (without Jjk) can be simplified to

Trbθkθj(t)θk(t)θjρb=fjk,k(t)θj(t)θk(t)θkθj=fjk,k(t)wjkwjkeKjk,jk(t),
(B7)

where Eq. (A2) has been used.

On the other hand, in the last term of Eq. (B1), the product of the first two operators within the trace operation can be expressed as

θkDj(t)=θkDj(t)θkθk=nωngn,j(1Wh(ωn))×(θkbnθkeiωnt+θkbnθkeiωnt)θk=Dj(t)θk+2θknωngn,jgn,k×(1Wh(ωn))Wh(ωn)cos(ωnt),
(B8)

where Eqs. (B3) and (B4) have been used. The above identity leads to the following expression for the trace of the bath operators in the last term of Eq. (B1):

TrbθkDj(t)θjρb=Mj,jk(t)+hj,k(t)wjk,
(B9)

where the identity of Eq. (A15) for τ = 0, the definition of Eq. (43), and the definition of Eq. (61) have been used. Inserting Eqs. (B7) and (B9) into Eq. (B1), one can then obtain the following expression:

Trb{B̃jk(t)δρ̃b,jk}=Jjkwjk(fjk,k(t)eKjk,jk(t)1)wjk+δjk(Mj,jk(t)+hj,k(t))wjk.
(B10)

It is easy to confirm that this expression is equivalent to Eq. (59).

The trace over the bath in the second order inhomogeneous term, Eq. (62), can be expressed as follows:

TrbB̃jk(t)B̃jk(τ)δρ̃b,jk=TrbJjk(θj(t)θk(t)wjk)+δjkDj(t)Jjk(θj(τ)θk(τ)wjk)+δjkDj(τ)θjρbθkwjkρb=JjkJjkFjk,jkjk(t,τ)+δjkJjkHj,jk(1),jk(t,τ)+JjkδjkHjk,j(2),jk(t,τ)+δjkδjkLj,jjk(t,τ),
(C1)

where

Fjk,jkjk(t,τ)=Trb(θj(t)θk(t)wjk)(θj(τ)θk(τ)wjk)(θjρbθkwjkρb),
(C2)
Hj,jk(1),jk(t,τ)=TrbDj(t)(θj(τ)θk(τ)wjk)(θjρbθkwjkρb),
(C3)
Hjk,j(2),jk(t,τ)=Trb(θj(t)θk(t)wjk)Dj(τ)(θjρbθkwjkρb),
(C4)
Lj,jjk(t,τ)=TrbDj(t)Dj(τ)(θjρbθkwjkρb).
(C5)

Further calculation of each of the terms above is straightforward as described below.

The first bath term in Eq. (C1), Fjk,jkjk(t,τ) defined by Eq. (C2), can be expanded further and expressed as follows:

Fjk,jkjk(t,τ)=θkθj(t)θk(t)θj(τ)θk(τ)θjwjkθkθj(τ)θk(τ)θjwjkθkθj(t)θk(t)θj+wjkwjkθkθjwjk(θj(t)θk(t)wjk)×(θj(τ)θk(τ)wjk).
(C6)

The first term in the above expression can be calculated as follows:

θkθj(t)θk(t)θj(τ)θk(τ)θj=fjk,k(t)fjk,k(τ)θj(t)θk(t)θj(τ)θk(τ)θkθj=fjk,k(t)fjk,k(τ)wjkwjkwjk×eKjk,jk(t)Kjk,jk(τ)Kjk,jk(tτ).
(C7)

For the second and third terms of Eq. (C6), Eq. (B7) can be used. The last term of Eq. (C6) corresponds to the first term that appears in the evaluation of R(t). Combining all of these, one can show that

Fjk,jkjk(t,τ)=wjkwjkwjkeKjk,jk(tτ)×fjk,k(t)fjk,k(τ)eKjk,jk(t)Kjk,jk(τ)1fjk,k(t)eKjk,jk(t)fjk,k(τ)eKjk,jk(τ)+2.
(C8)

The bath term contributing to the second term of Eq. (C1), i.e., Hjk,jk(1),jk(t,τ) defined by Eq. (C3), is expressed as follows:

Hj,jk(1),jk(t,τ)=θkDj(t)θj(τ)θk(τ)θiwjkθkDj(t)θjwjkDj(t)(θj(τ)θk(τ)wjk).
(C9)

The first term in the above expression can be shown to be

θkDj(t)θj(τ)θk(τ)θj=fjk,k(τ)Dj(t)θj(τ)θk(τ)θkθj+fjk,k(τ)hj,k(t)θj(τ)θk(τ)θkθj,
(C10)

where the following identities have been used,

θkDj(t)θj(τ)θk(τ)=θkDj(t)θkθkθj(τ)θk(τ)θkθk,
(C11)
θkDj(t)θk=Dj(t)+hj,k(t),
(C12)
θkθj(τ)θk(τ)θk=fjk,k(τ)θj(τ)θk(τ),
(C13)

along with the definitions of Eqs. (60) and (61). In Eq. (C10), the first bath average on the right-hand side can be calculated employing identities similar to those leading to Eq. (A21). The resulting expression is as follows:

Dj(t)θj(τ)θk(τ)θkθj=(Mj,jk(tτ)+Mj,jk(t))×wjkwjkeKjk,jk(τ).
(C14)

Combining this with the identity given by Eq. (A2) (with the replacement of tτ and τ → 0), one can show that Eq. (C10) can be expressed as follows:

θkDj(t)θj(τ)θk(τ)θj=fjk,k(τ)Mj,jk(tτ)+Mj,jk(t)+hj,k(t)wjk×wjkeKjk,jk(τ).
(C15)

The second bath average term on the right-hand side of Eq. (C9) can be shown to be

θkDj(t)θj=(Mj,jk(t)+hj,k(t))wjk,
(C16)

where Eqs. (A21) and (C12) along with the definition of Eq. (43) have been used.

For the last term on the right-hand side of Eqs. (C9), Eq. (A15) can be employed. As a result, Eq. (C9) can be expressed as

Hj,jk(1),jk(t,τ)=fjk,k(τ)eKjk,jk(τ)1×wjkwjkMj,jk(tτ)+Mj,jk(t)+hj,k(t).
(C17)

Hjk,j(2),jk(t,τ) defined by Eq. (C4) can be calculated in a similar manner but can, in fact, be calculated using its relation to Hj,kj(1),kj(τ,t) as follows:

Hjk,j(2),jk(t,τ)=Hj,kj(1),kj(τ,t)*=fkj,j(t)eKkj,kj(t)1*wjkwjk×Mj,kj(τt)+Mj,kj(τ)+hj,j(τ)*=fjk,k(t)eKjk,jk(t)1wjkwjk×Mj,kj(tτ)+Mj,kj(τ)+hj,j(τ).
(C18)

The first equality of the above equation can be confirmed from the definitions, Eqs. (C3) and (C4), and the third equality results from the following identities: fkj,j(t)eKkj,kj(t)*=fjk,k(t)eKjk,jk(t); Mj′,kj(τt)* = Mj′,kj(tτ); Mj,kj(τ)*=Mj,kj(τ); and hj,j(τ)*=hj,j(τ).

Finally, Eq. (C5) can be expressed as

Lj,jjk(t,τ)=θkDj(t)Dj(τ)θjwjkDj(t)Dj(τ).
(C19)

In the above expression, the first term on the right-hand side can be expressed as

θkDj(t)Dj(τ)θj=Dj(t)Dj(τ)θkθj+hj,k(t)Dj(τ)×θkθj+hj,k(τ)Dj(t)θkθj+hj,k(t)hj,k(τ)wjk,
(C20)

where the first term can be calculated as follows:

Dj(t)Dj(τ)θkθjρb=nnωngn,j(1Wh(ωn))ωngn,j×(1Wh(ωn))ei(ωnt+ωnτ)×bnbnθkθj+ei(ωntωnτ)×bnbnθkθj+ei(ωntωnτ)×bnbnθkθj+ei(ωnt+ωnτ)×bnbnθkθj.
(C21)

In the above expression, the four averages over the bath can be calculated explicitly and can be expressed as

bnbnθkθj=wjkδgn,jkWh(ωn)(1eβωn)δgn,jkWh(ωn)(1eβωn),
(C22)
bnbnθkθj=wjk11eβωnδnnδgn,jkWh(ωn)(1eβωn)δgn,jkWh(ωn)(1eβωn)eβωn,
(C23)
bnbnθkθj=wjkeβωn1eβωnδnnδgn,jkWh(ωn)(1eβωn)×eβωnδgn,jkWh(ωn)(1eβωn),
(C24)
bnbnθkθj=wjkδgn,jkWh(ωn)(1eβωn)eβωn×δgn,jkWh(ωn)(1eβωn)eβωn.
(C25)

It is worth noting that the additional factor for n = n′ in Eq. (C23) comes from the first term of the following identity,

bnbneγ*bnγbn=1(1eβωn)|γ|2eβωn(1eβωn)2×ecoth(βωn/2)|γ|2/2.
(C26)

For the case of (C24), the fact that bnbn=bnbn1 can be combined with the above identity. When Eqs. (C22)(C25) are inserted into Eq. (C21), the term for n = n′ cancel wjkDj(t)Dj(τ) in Eq. (C19) and the double summation for n and n′ can be factored as follows:

Lj,jjk(t,τ)=wjk(Mj,jk(t)+hj,k(t))Mj,jk(τ)+hj,k(τ).
(C27)

For all the bath correlation functions calculated above, let us define the following time integrals:

F̃jk,jkjk(t;E)=0tdτeiE(tτ)/Fjk,jkjk(t,τ),
(C28)
H̃j,jk(1),jk(t;E)=0tdτeiE(tτ)/Hj,jk(1),jk(t,τ),
(C29)
H̃jk,j(2),jk(t;E)=0tdτeiE(tτ)/Hjk,j(2),jk(t,τ),
(C30)
L̃j,jjk(t;E)=0tdτeiE(tτ)/Lj,jjk(t,τ).
(C31)

The above time correlation functions are useful for expressing the second order inhomogeneous terms in the basis of eigenstates of H̃s.

1.
L. D.
Landau
and
S. I.
Pekar
, “
Effective mass of a polaron
,”
Ukr. J. Phys.
53
,
71
74
(
2008
).
2.
H.
Fröhlich
, “
Electrons in lattice fields
,”
Adv. Phys.
3
,
325
(
1954
).
3.
T.
Holstein
, “
Studies of polaron motion: Part 1. The molecular-crystal model
,”
Ann. Phys.
8
,
325
(
1959
).
4.
T.
Holstein
, “
Studies of polaron motion: Part 2. The ‘small’ polaron
,”
Ann. Phys.
8
,
343
(
1959
).
5.
T.
Holstein
, “
Quantal occurrence-probability treatment of small-polaron hopping
,”
Philos. Mag. B
37
,
49
(
1978
).
6.
V. J.
Emery
and
A.
Luther
, “
Low-temperature properties of the Kondo Hamiltonian
,”
Phys. Rev. B
9
,
215
225
(
1974
).
7.
S.
Rackovsky
and
R.
Silbey
, “
Electronic energy transfer in impure solids I. Two molecules embedded in a lattice
,”
Mol. Phys.
25
,
61
(
1973
).
8.
B.
Jackson
and
R.
Silbey
, “
On the calculation of transfer rate between impurity states in solids
,”
J. Chem. Phys.
78
,
4193
(
1983
).
9.
R.
Silbey
and
R. A.
Harris
, “
Variational calculation of the dynamics of a two level system interacting with a bath
,”
J. Chem. Phys.
80
,
2615
(
1984
).
10.
R. A.
Harris
and
R.
Silbey
, “
Variational calculation of the tunneling system interacting with a heat bath. II. Dynamics of an asymmetric tunneling system
,”
J. Chem. Phys.
83
,
1069
(
1985
).
11.
R.
Silbey
and
R. A.
Harris
, “
Tunnelingof molecules in low-temperature media: An elementary description
,”
J. Phys. Chem.
93
,
7062
(
1989
).
12.
A.
Nitzan
,
Chemical Dynamics in Condensed Phases
(
Oxford University Press
,
Oxford
,
2006
).
13.
Y.-C.
Cheng
and
R. J.
Silbey
, “
A unified theory for charge-carrier transport in organic crystals
,”
J. Chem. Phys.
128
,
114713
(
2008
).
14.
S.
Jang
,
Y.-C.
Cheng
,
D. R.
Reichman
, and
J. D.
Eaves
, “
Theory of coherent resonance energy transfer
,”
J. Chem. Phys.
129
,
101104
(
2008
).
15.
S.
Jang
, “
Theory of coherent resonance energy transfer for coherent initial condition
,”
J. Chem. Phys.
131
,
164101
(
2009
).
16.
S.
Jang
, “
Theory of multichromophoric coherent resonance energy transfer: A polaronic quantum master equation approach
,”
J. Chem. Phys.
135
,
034105
(
2011
).
17.
A.
Nazir
, “
Correlation-dependent coherent to incoherent transitions in resonant energy transfer dynamics
,”
Phys. Rev. Lett.
103
,
146404
(
2009
).
18.
D. P. S.
McCutcheon
and
A.
Nazir
, “
Consistent treatment of coherent and incoherent energy transfer dynamics using a variational master equation
,”
J. Chem. Phys.
135
,
114501
(
2011
).
19.
E. N.
Zimanyi
and
R. J.
Silbey
, “
Theoretical description of quantum effects in multi-chromophoric aggregates
,”
Philos. Trans. R. Soc., A
370
,
3620
(
2012
).
20.
L.
Yang
,
M.
Devi
, and
S.
Jang
, “
Polaronic quantum master equation theory of inelastic and coherent resonance energy transfer for soft systems
,”
J. Chem. Phys.
137
,
024101
(
2012
).
21.
F. A.
Pollock
,
D. P. S.
McCutcheon
,
B. W.
Lovett
,
E. M.
Gauger
, and
A.
Nazir
, “
A multi-site variational master equation approach to dissipative energy transfer
,”
New J. Phys.
15
,
075018
(
2013
).
22.
A.
Nazir
and
D. P. S.
McCutcheon
, “
Modelling exciton-phonon interactions in optically driven quantum dots
,”
J. Phys.: Condens. Matter
28
,
103002
(
2016
).
23.
V.
Pouthier
, “
The reduced dynamics of an exciton coupled to a phonon bath: A new approach combining the Lang-Firsov transformation and the perturbation theory
,”
J. Chem. Phys.
138
,
044108
(
2013
).
24.
D.
Chen
,
J.
Ye
,
H.
Zhang
, and
Y.
Zhao
, “
On the Munn-Silbey approach to polaron transport with off-diagonal coupling and temperature-dependent canonical transformations
,”
J. Phys. Chem. B
115
,
5312
5321
(
2011
).
25.
Y.
Zhao
,
B.
Luo
,
Y.
Zhang
, and
J.
Ye
, “
Dynamics of a Holstein polaron with off-diagonal coupling
,”
J. Chem. Phys.
137
,
084113
(
2012
).
26.
V.
Chorosajev
,
A.
Gelzinis
,
L.
Valkunas
, and
D.
Abramavicius
, “
Dynamics of exciton-polaron transition in molecular assemblies: The variational approach
,”
J. Chem. Phys.
140
,
244108
(
2014
).
27.
P.
Hamm
and
G. P.
Tsironis
, “
Barrier crossing to the small Holstein polaron regime
,”
Phys. Rev. B
78
,
092301
(
2008
).
28.
C. K.
Lee
,
J.
Moix
, and
J.
Cao
, “
Coherent quantum transport in disordered systems: A unified polaron treatment of hopping and band-like transport
,”
J. Chem. Phys.
142
,
164103
(
2015
).
29.
D.
Xu
and
J.
Cao
, “
Non-canonical distribution and non-equilibrium transport beyond weak system-bath coupling regime: A polaron transformation approach
,”
Front. Phys.
11
,
110308
(
2016
).
30.
Y.-C.
Wang
and
Y.
Zhao
, “
Variational polaron transformation approach toward the calculation of thermopower in organic crystals
,”
Phys. Rev. B
101
,
075205
(
2020
).
31.
D.
Balzer
,
T. J. A. M.
Smolders
,
D.
Blyth
,
S. N.
Hood
, and
I.
Kassal
, “
Delocalized kinetic Monte Carlo for simulating delocalization-enhanced charge and exciton transport in disordered materials
,”
Chem. Sci.
12
,
2276
2285
(
2021
).
32.
H.-H.
Teh
,
B.-Y.
Jin
, and
Y.-C.
Cheng
, “
Frozen-mode small polaron quantum master equation with variational bound for excitation energy transfer in molecular aggregates
,”
J. Chem. Phys.
150
,
224110
(
2019
).
33.
A. O.
Caldeira
and
A. J.
Leggett
, “
Quantum tunnelling in a dissipative system
,”
Ann. Phys.
149
,
374
(
1983
).
34.
U.
Weiss
,
Quantum Dissipative Systems
, Series in Modern Condensed Matter Physics Vol. 2 (
World Scientific
,
Singapore
,
1993
).
35.
V. M.
Kenkre
and
P.
Reineker
,
Exciton Dynamics in Molecular Crystals and Aggregates
(
Springer
,
Berlin
,
1982
).
36.
V.
May
and
O.
Kühn
,
Charge and Energy Transfer Dynamics in Molecular Systems
(
Wiley VCH
,
Weinheim
,
2011
).
37.
S. J.
Jang
,
Dynamics of Molecular Excitons
, Nanophotonics Series (
Elsevier
,
Amsterdam
,
2020
).
38.
V.
Coropceanu
,
J.
Cornil
,
D. A.
da Silva Filho
,
Y.
Olivier
,
R.
Silbey
, and
J.-L.
Brédas
, “
Charge transport in organic semiconductors
,”
Chem. Rev.
107
,
926
952
(
2007
).
39.
S.
Jang
,
J.
Cao
, and
R. J.
Silbey
, “
Fourth order quantum master equation and its Markovian bath limit
,”
J. Chem. Phys.
116
,
2705
(
2002
).
40.
F.
Shibata
and
T.
Arimitsu
, “
Expansion formulas in nonequilibrium statistical mechanics
,”
J. Phys. Soc. Jpn.
49
,
891
(
1980
).
41.
A. J.
Leggett
,
S.
Chakravarty
,
A. T.
Dorsey
,
M. P. A.
Fisher
,
A.
Garg
, and
W.
Zwerger
, “
Dynamics of the dissipative two-state system
,”
Rev. Mod. Phys.
59
,
1
85
(
1987
).
42.
C.
Aslangul
,
N.
Pottier
, and
D.
Saint-James
, “
Quantum ohmic dissipation: Transition from coherent to incoherent dynamics
,”
Phys. Lett. A
110
,
249
(
1985
).
43.
H.
Dekker
, “
Noninteracting-blip approximation for a two-level system coupled to heat bath
,”
Phys. Rev. A
35
,
1436
1437
(
1987
).
44.
Y.
Lai
and
E.
Geva
, “
On simulating the dynamics of electronic populations and coherences via quantum master equations based on treating off-diagonal electronic coupling terms as a small perturbation
,”
J. Chem. Phys.
155
,
204101
(
2021
).
45.
W.
Weibull
, “
A statistical distribution function of wide applicability
,”
ASME J. Appl. Mech.
18
,
293
297
(
1951
).