Multitime quantum correlation functions are central objects in physical science, offering a direct link between the experimental observables and the dynamics of an underlying model. While experiments such as 2D spectroscopy and quantum control can now measure such quantities, the accurate simulation of such responses remains computationally expensive and sometimes impossible, depending on the system’s complexity. A natural tool to employ is the generalized quantum master equation (GQME), which can offer computational savings by extending reference dynamics at a comparatively trivial cost. However, dynamical methods that can tackle chemical systems with atomistic resolution, such as those in the semiclassical hierarchy, often suffer from poor accuracy, limiting the credence one might lend to their results. By combining work on the accuracy-boosting formulation of semiclassical memory kernels with recent work on the multitime GQME, here we show for the first time that one can exploit a multitime semiclassical GQME to dramatically improve both the accuracy of coarse mean-field Ehrenfest dynamics and obtain orders of magnitude efficiency gains.

Multitime correlation functions, C(t1,t2,,tj), are fundamental dynamical objects that characterize chemical and physical systems. When a system is subject to external stimuli, its response can be formulated as a multitime correlation that relates directly to experimental observables, as in multidimensional spectroscopy1 and quantum control protocols.2–4 Through this formalism, the energetic structure and dynamics of all manner of physical systems become available, from nanocrystals to extended solids and from gas phase molecules to solvated chromophores. Beyond materials science, multitime correlation functions also reveal the extent of quantum information scrambling in model systems,5 chemical reactions,6,7 black holes,8 and many-body localized systems9 via out-of-time-ordered correlators. Hence, developing accurate and efficient theoretical and simulation frameworks to predict these objects is of paramount importance in chemistry, materials science, condensed matter physics, and astrophysics. Mathematically, while numerically exact solvers can predict the dynamics of few-state systems in contact with Gaussian heat baths,10–17 going beyond effective harmonic environments and accounting for non-Gaussian, strong coupling to anharmonic motions remains a fundamental challenge. Such non-Gaussian interactions abound, emerging in charge transfer within structured environments,18–21 optical spectroscopy of liquids,22,23 interacting spin systems,24,25 shot noise in measurements of current,26,27 and radiation pressure.28 Hence, a dynamical method should ideally capture faithful representations of fermion- and spin-nuclear couplings at the ab initio level. Path integrals and semiclassical methods29,30 can, in principle, achieve this. However, in practice, these methods must balance accuracy and efficiency.31 Indeed, some techniques are accurate but prohibitively expensive, while others are largely inaccurate but (somewhat) affordable. It is thus crucial to develop reliable tools to reduce the computational cost and improve the accuracy of such schemes.

Over the past decade, the generalized quantum master equation (GQME) framework32–34 has proven transformative in this regard for 1-time correlation functions: GQMEs can both drastically reduce the cost of path integral, semiclassical, and classical simulations—sometimes by multiple orders of magnitude—and also boost the accuracy of even starkly incorrect semiclassical results to within graphical accuracy of benchmark calculations. Such improvements have been shown for the 1-time nonequilibrium and equilibrium dynamics of Gaussian,35–44 fermionic,45–48 and atomisitic49 models of charge and energy transfer in the condensed phase, and even for many-level photosynthetic complexes.50,51 The GQME can even be constructed using mean-field theory (MFT) simulations,38,40,42,43,50,52 arguably the least accurate semiclassical method.53 

The GQME exactly rewrites a high-dimensional dynamics problem as a low-dimensional equation of motion of C(t), which encodes the correlation of the variables of interest. At the heart of the GQME is the choice of the projection operator,54,55 which accounts for the initial preparation of the chemical system (i.e., equilibrium vs nonequilibrium) and selects a few variables to track dynamically. The projection operator also defines the memory kernel, K(t), which encapsulates the effect of all dynamical variables not being explicitly tracked and places an upper limit on how much simulation time is required to predict the dynamics of C(t). The efficiency gains arise from adequately choosing a projection operator that minimizes the memory kernel lifetime. What leads to the accuracy gains is less clear. While early work posited that the accuracy improvements arise from the short-time accuracy combined with short-lived memory kernels, later work40 showed one could obtain efficiency gains even in memory kernels that are as long-lived as C(t). In fact, Ref. 40 proposed that the improved accuracy in 1-time nonequilibrium averages and correlation functions arises from the exact semiclassical sampling of bath operators at t = 0, which corrects the memory kernel through its self-consistent extraction from bath-augmented auxiliary kernels.

We have recently shown that one can formulate GQMEs that drastically reduce the cost of multitime simulations,56 but is it also possible to use these to improve semiclassical multitime simulations? After all, recent advances in semiclassical theory have enabled the prediction of multitime correlation functions,57–65 and the GQME could offer the ability to leverage this technology at a lower cost and greater accuracy. Yet, this is not a foregone conclusion. After all, our presumed source of improved accuracy, that is, the exact semiclassical sampling of bath correlations, is only strictly true at t = 0. This raises important and revealing questions:

  1. Does a bath-sampling GQME based on approximate semiclassical dynamics allow us to improve accuracy across external temporal interactions?

  2. How does additional bath sampling at later times impact the accuracy of the GQME predictions?

To address these questions, we first describe how one must formulate 1-time semiclassical GQMEs to improve efficiency and accuracy. We then introduce our extension of this method to 2-time simulations. By treating the simplest multitime case, we can proceed using fully continuous, rather than discretized, equations. For problems with additional intermediate-time interactions, the theory is trivially extendable, but by remaining at a single measurement, we preempt challenges of time and storage cost that add an additional layer of complexity, as in Ref. 56. We employ a spin-boson model with weak system–bath coupling and a fast decorrelating bath to compare our MFT and GQME results with an exact HEOM calculation. We refer the reader to  Appendix A for computational details.

What makes the semiclassical 1-time GQME work? In the Mori–Nakajima–Zwanzig GQME formalism,32–34 one can obtain a low-dimensional (N × N) equation of motion for the correlation function,
(1)
where A is a vector of N operators whose dynamics we aim to track, and the inner product can be defined as needed for the physical problem of interest. Using the projection operator, P=AA, yields the following time-nonlocal equation of motion for C,
(2)
Here, we use the Argyres–Kelley projector, such that A = {|0⟩⟨0|, |0⟩⟨1|, |1⟩⟨0|, |1⟩⟨1|}, and the inner product is the full trace over spectroscopically separable initial conditions, Cij(t)=trAjeiLtAiρB, where ρB is the canonical density of the nuclear bath (see  Appendix A).

The GQME shifts attention from the correlation function C(t) to the memory kernel K(t). That is, as long as one knows K(t), it is easy to integrate Eq. (2) to obtain the dynamics of C(t). In general, this memory kernel has a shorter lifetime, τK, than the correlation function’s equilibration time, teq, setting an upper limit on the amount of information required to characterize the evolution of C(t). That is, τK<teq, where K(t>τK)=0.

The 1-time memory kernel takes the form
(3)
where Q=1P is the operator that projects objects in the full space onto the complementary space. However, the presence of the projected propagator, eiQLt, makes solving for the memory kernel challenging. While much work in quantum dynamics has focused on obtaining perturbative expansions of K(t), we adopt its exact, self-consistent expansion,35,40,41
(4)
into auxiliary kernels,
(5)
and
(6)
that require evolution with the full propagator, eiLt, instead. This expansion enables researchers to choose a convenient dynamical method to simulate the auxiliary kernels, compute K(t), and ultimately predict C(t). The superscript notation K(3b) is chosen to be consistent with previous work in the Heisenberg picture.40 
When one is interested in employing the GQME framework exclusively to improve computational efficiency, one can re-express the auxiliary kernels in terms of C(t) and its time derivatives.40,41 To obtain such expressions, one substitutes the definition of Q in the auxiliary kernels and replaces all Liouvillians with the numerical time derivatives. For example,
(7)

Alternatively, one can use the tensor transfer method,66 which is formally equivalent. Thus, all relevant quantities can be extracted from reference dynamics for C(t).

For the GQME framework to improve the accuracy of 1-time semiclassical and quantum-classical dynamics, one must avoid fully decomposing the auxiliary kernels into C(t) and its time derivatives. Exceptions arise when using methods whose approximate propagator, eappiLt, commutes with the Liouvillian, L, or that preserves the canonical distribution when computing equilibrium correlation functions. In such cases, the GQME framework cannot lead to any improvement in accuracy,41 but can still be used to improve computational efficiency. That is, the quality of the dynamics obtained is the same as that of the original semiclassical method used to parameterize it. However, since these exceptions do not arise for most methods in the semiclassical hierarchy67–77 or that arise from the quantum-classical Liouville equation,78–88 one may expect a suitable formulation of the semiclassical GQME to lead to improved dynamics.

In anticipation of the multitime derivation, and to contextualize the extent of GQME-based improvement that one may expect in the multitime response, we first illustrate how the semiclassical GQME can improve the accuracy of the MFT electronic dynamics in a parameter regime corresponding to a biased, underdamped spin connected to a fast decorrelating bath. Figure 1 compares the performance of the MFT against benchmark HEOM calculations and illustrates the improvement one obtains from the GQME sampling bath correlations during the MFT simulation. The GQME brings the approximate MFT dynamics to within the graphical accuracy of the exact HEOM. We have chosen this parameter regime as it places particular strain on MFT: similar to many other semiclassical methods, MFT violates detailed balance, a flaw that becomes most evident in biased systems, and the classical approximation for the bath deteriorates the most for fast, high-frequency baths. While a lower temperature would have rendered the classical treatment of the bath in MFT more questionable and heightened the disagreement between MFT and HEOM, anticipating a nearly prohibitive cost of converging the multitime HEOM benchmark, we have adopted an intermediate temperature that allows faster convergence with the number of Matsubara frequencies.

These data demonstrate why the argument claiming that accuracy improvement arises from the short-time accuracy of MFT (leading to memory kernels that are accurate over their short lifetimes) is incomplete. Indeed, the GQME kernel is more accurate than the MFT equivalent even before the first minimum in the dynamics at very early times.

FIG. 1.

HEOM, MFT, and GQME results for a spin-boson model with Δ = ϵ, ωc = 7.5Δ, λ = 0.075Δ, β = 1/Δ, and an Ohmic spectral density. The gray line marks the cutoff, 0.7 Δ−1, used to generate the GQME dynamics. Left: difference between site populations after initializing on the lower-energy site. The GQME (red dotted) recovers the HEOM reference (black), in contrast to the direct MFT result (dashed yellow). Right: representative element of the memory kernel for all methods over the first 2 Δ−1. Inset: the GQME error given the truncation of the kernel (dotted green) at time τ. Errors increase after a minimum due to the poorer convergence of the MFT trajectories at later times.

FIG. 1.

HEOM, MFT, and GQME results for a spin-boson model with Δ = ϵ, ωc = 7.5Δ, λ = 0.075Δ, β = 1/Δ, and an Ohmic spectral density. The gray line marks the cutoff, 0.7 Δ−1, used to generate the GQME dynamics. Left: difference between site populations after initializing on the lower-energy site. The GQME (red dotted) recovers the HEOM reference (black), in contrast to the direct MFT result (dashed yellow). Right: representative element of the memory kernel for all methods over the first 2 Δ−1. Inset: the GQME error given the truncation of the kernel (dotted green) at time τ. Errors increase after a minimum due to the poorer convergence of the MFT trajectories at later times.

Close modal
At the technical level, the GQME displayed in Fig. 1 circumvents decomposing the auxiliary kernels into C(t) and its time derivatives and instead directly simulates the new correlation functions one obtains by rotating the initial and final conditions in the auxiliary kernels. That is, QLA=LSBA and ALQ=ALSB, where LSB=[HSB,] is the system–bath Liouvillian (see  Appendix B). For example, noting that
(8)
one finds that for the row n of Knm(3b)(t), one needs to rotate the initial condition as follows:
(9)
Thus, calculating this quantity without using the time derivatives of C(t) requires sampling the bath coupling VB (here at time zero) in addition to the original system observables, |j⟩⟨k|. Furthermore, to sample the resulting correlation functions semiclassically, one must Wigner transform the additional bath operators.86 Since the functional forms require capturing the Wigner transform of products of bath operators, one must employ the Moyal bracket,89,90
(10)
where Λ is the classical Poisson bracket and the linearization is exact when VB is a linear function of the bath coordinates. This shows that the Wigner transformation that one uses to facilitate a semiclassical treatment generates an extra term, ζB,w, containing the derivative of the coupling (see  Appendix A). For a term like Knm(1)(t), performing all rotations arising from LSB requires sampling bath operators at time t. It turns out that only time-zero sampling is significant to obtaining the true kernel,40 but here we work with the fully rotated expressions and test this observation explicitly later.

Semiclassical multitime GQME: We now consider the case where the system experiences an impulsive external field at time t1, which we choose to be the σx Pauli matrix. This corresponds to a photon-induced spin-flip of a qubit in quantum information protocols.2 

In general, time evolution causes the electronic system and bath to become entangled, making it impossible to write the full density matrix as a simple product of a system and bath component. Because this density matrix differs from the form in the projection operator—that is, multiplicatively separable with the bath always at equilibrium with the ground electronic state—introducing an interaction at intermediate times prevents the resulting multitime correlation function from being decomposable into products of 1-time correlation functions.91 In the mean-field method, the time-evolved density matrix of the entire system remains multiplicatively separable at all times. However, upon interaction with a photon, the system component of the density matrix collapses into a mixed state that cannot be separated into a simple product of a bath operator and a pure subsystem state.92 Thus, even with the simplifying approximation in the Ehrenfest method that neglects global correlation, the multitime equation does not simplify to the 1-time GQME.

Instead, we must invoke a multitime GQME. This formalism expresses the correction term using a new memory kernel-like object that takes t1 as an argument. We refer to it as the 2-time kernel Kσx(t2,t1). In the multitime GQME formalism,56,91 one can decompose the 2-time correlation function as follows:
(11)
where Σx is the matrix (A|σ̂x|A). The first term can be constructed using only 1-time data and represents the solution when the entanglement induced by the measurement at t1 is ignored. The second term isolates this 2-time contribution and requires a full 2-time simulation to obtain. For the parameter regime in Fig. 1, we expect the 2-time entanglement term to be small56 but otherwise representative of the protocol we are delineating. The 2-time kernel subject to a σ̂x interaction at t1 is
(12)
We have previously shown how to self-consistently expand this multitime kernel and decompose it into auxiliary kernels that employ only derivatives of the parent 2- and 1-time correlation functions.56 However, such decompositions cannot offer accuracy improvements when combined with semiclassics. Instead, in  Appendix B, we show how one could obtain expressions that can offer accuracy improvements by explicitly considering the LSB-induced rotations in the auxiliary kernels. The first difference with our previous work is that instead of derivatives of the dynamics appearing under the convolution integrals [as in Eq. (4)], we require the semiclassical, bath-sampled 1-time K(3f)(t) and K(3b)(t). The second difference is that we must Wigner transform the 2-time equivalent of K(1),
(13)
where X and iY are the diagonal rotation matrices representing the action of σ̂z in an anticommutator and commutator, respectively.40 This shows that one must measure bath operators VB,w and ζB,w at t = 0 and VB,w at t = t1 + t2, in analogy to the 1-time case. There is no need to sample the bath variables during the intermediate evolution (at t1) because LSB only appears at the start and end of the inner product. The same protocol holds for higher-order multitime protocols.

We now turn to our simulation results. First, we compute the benchmark 2-time correlation function (Fig. 2) using the numerically exact HEOM over a 10 × 10 Δ−1 grid of points with a 0.01 Δ−1 resolution. For the GQME, we do not need to directly simulate the entire grid, which is one of the central benefits of the method. Instead, we demarcate the first 2 × 2 Δ−1 of the grid with black boxes and first simulate this region with MFT. We chose this value because our 1-time kernel lifetime is 0.7 Δ−1 (see Fig. 1), and our previous work56 suggests the 2-time kernel goes to zero at least as fast as the 1-time kernel. A 2 Δ−1 range on both axes should be more than sufficient to fully parameterize the 2-time kernel, and thus generate the GQME dynamics for the full 10 × 10 Δ−1 grid in Fig. 2. In this and all the following figures, the 16 panels represent the matrix elements of the displayed quantity, with the i = j = 0 element in the top left.

FIG. 2.

Two-time correlation function obtained using HEOM of our spin-boson model with ϵ = Δ, ωc = 7.5Δ, λ = 0.075Δ, β = Δ, and an Ohmic spectral density. The total grid displays lightly damped oscillations, also evident in Fig. 1. The black boxes in the lower left of each panel demarcate the time window over which the 2-time memory kernels decay.

FIG. 2.

Two-time correlation function obtained using HEOM of our spin-boson model with ϵ = Δ, ωc = 7.5Δ, λ = 0.075Δ, β = Δ, and an Ohmic spectral density. The total grid displays lightly damped oscillations, also evident in Fig. 1. The black boxes in the lower left of each panel demarcate the time window over which the 2-time memory kernels decay.

Close modal

To quantify the improvement that the multitime semiclassical GQME can afford, we hone in on the difference between the full propagator and the 2-time correlation function constructed using only 1-time quantities, that is, the second term of Eq. (11). Figure 3 (left) shows the benchmark of this quantity obtained with HEOM. For these exact data, we employ a continuous-time algorithm (see  Appendix A of Ref. 56) to extract the 2-time kernel, shown in Fig. 3 (right) from the 2-time term. In this regime, the 2-time kernel has only two significant elements that are confined to small triangles of 0.5 Δ−1 base. This confirms our prediction that 2 Δ−1 is an adequate simulation time for the MFT trajectories, to which we now turn.

FIG. 3.

From the HEOM data of Fig. 6, we calculate (left) the 2-time correction to the 1-time approximation corresponding to the second term of Eq. (11), Cσx(t2,t1)C(t2)ΣxC(t1), and (right) the 2-time kernel, Kσx(t2,t1), of Eq. (12). The 0.5 Δ−1 axis limits are shown as black boxes in the left panel.

FIG. 3.

From the HEOM data of Fig. 6, we calculate (left) the 2-time correction to the 1-time approximation corresponding to the second term of Eq. (11), Cσx(t2,t1)C(t2)ΣxC(t1), and (right) the 2-time kernel, Kσx(t2,t1), of Eq. (12). The 0.5 Δ−1 axis limits are shown as black boxes in the left panel.

Close modal

Figure 4 presents both MFT and semiclassical GQME equivalents to the panels of Fig. 3. We employ the semiclassical GQME-improved version of the 1-time quantities presented in Fig. 1, which are effectively identical to the HEOM in this case. This enables us to isolate all errors in the 2-time term from the 2-time kernel. The top row of Fig. 4 shows that MFT is completely inaccurate, even at the shortest times. In particular, the 2-time term of Fig. 4(a) has no clear structure and is almost an order of magnitude too large. This means that, in this regime, discarding the 2-time term completely and just using 1-time improvements would be preferable to the result of the 2-time calculation. We construct the associated 2-time kernel in Fig. 4(b) from derivatives of the 2-time term and, therefore, obtain an object that is also of the wrong form and magnitude.

FIG. 4.

Top: MFT results for (a) 2-time term Cσx(t2,t1)C(t2)ΣxC(t1) and (b) 2-time kernel Kσx(t2,t1) calculated using the same (exact) derivative relationships as in Fig. 3. Bottom: GQME results for (c) 2-time term 0t20t1dτ2dτ1C(t2τ2)Kσx(t2,t1)C(t1τ1) where the (d) 2-time kernel is calculated with Eq. (13) and the improved 1-time quantities, as described in  Appendix B.

FIG. 4.

Top: MFT results for (a) 2-time term Cσx(t2,t1)C(t2)ΣxC(t1) and (b) 2-time kernel Kσx(t2,t1) calculated using the same (exact) derivative relationships as in Fig. 3. Bottom: GQME results for (c) 2-time term 0t20t1dτ2dτ1C(t2τ2)Kσx(t2,t1)C(t1τ1) where the (d) 2-time kernel is calculated with Eq. (13) and the improved 1-time quantities, as described in  Appendix B.

Close modal

Interestingly, if one rotates only the first QL term with LSB when obtaining Eq. (13), that is, performing bath sampling only at t = 0, then the improvement is lost. Previous work40 found that when replacing the remaining Q with 1P, the difference between the two terms was slow to converge as a function of the number of trajectories. Hence, we ascribe this observation to the increased difficulty of converging the small difference between two large terms subject to statistical noise from finite sampling.

Currently, the MFT multitime protocol is computationally expensive: our 2-time MFT simulations average over 10 000 trajectories, which has the same cost as performing 100 000 000 1-time trajectories. To obtain the 2-time analog of the 100 000 that satisfactorily converges the 1-time data in Fig. 1 would take four orders of magnitude longer. While converging the data over the entire time domain is feasible for the spin-boson model as semiclassical calculations are trivially parallelizable, performing this many simulations with an atomistic bath would become difficult as force calculations constitute the greatest expense in molecular dynamics. For this reason, we demonstrate that it is possible to obtain improved accuracy and efficiency through our semiclassical GQME even without fully converging the long-term data. Indeed, while at this level of convergence the 2-time kernel of Fig. 4(d) is still noisy compared to the HEOM result in Fig. 3 (right), the GQME 2-time term in Fig. 4(c) obtains semiquantitative agreement in form and magnitude. In addition, because it is a trajectory-based method, earlier time data is more converged, and we only need convergence in the region used in the memory kernel construction, that is, the smaller boxes in Fig. 4 (left). To be quantitative about the degree to which the GQME kernel is underconverged, in Fig. 5, we show the difference between GQME and HEOM 2-time terms on these two timescales.

FIG. 5.

Differences between the 2-time correction for the GQME, Fig. 4(c), and that of HEOM, Fig. 3 (left), (left) over the full range of Fig. 4(c) and (right) over the region needed to construct the kernel of Fig. 4(d), where the error is insignificant on the appropriate, 10−2 scale.

FIG. 5.

Differences between the 2-time correction for the GQME, Fig. 4(c), and that of HEOM, Fig. 3 (left), (left) over the full range of Fig. 4(c) and (right) over the region needed to construct the kernel of Fig. 4(d), where the error is insignificant on the appropriate, 10−2 scale.

Close modal

The fact that there is little to no error in the 0.5 Δ−1 region of Fig. 5 (right) directly represents an improvement in accuracy: while the 2-time term is not maximal in this region, one can see from the (1, 3) and (2, 0) elements of Fig. 3 (left) that it is non-zero and not trivial to obtain. The larger errors at later times are due to an accumulation of error from the underconverged tails of Kσx(t2,t1), consistent with the observation that semiclassical methods require increasingly larger numbers of trajectories to converge correlation functions evaluated at longer times. Since the size of the 2-time term itself is the same as the error, careful comparison shows this is mostly due to slightly shifted frequencies (relative positions of regions of positive and negative signal) rather than their intensities. We can therefore have some confidence that—since the magnitudes are correct, the frequencies are largely accurate, and the GQME 2-time kernel of Fig. 4(c) clearly goes to zero—it is reasonable to evolve the GQME to the full 10 × 10 Δ−1 grid of Fig. 2. Using just the data in the 0.5 × 0.5 Δ−1 region and the 1-time improvement (which is of insignificant cost in comparison), we obtain the GQME 2-time propagator and display it in Fig. 6.

FIG. 6.

2-time correlation function from the semiclassical GQME, to be compared with the HEOM equivalent of Fig. 2. In this figure, the (smaller) black boxes instead show the region of time that needed to be simulated to compute the full grid.

FIG. 6.

2-time correlation function from the semiclassical GQME, to be compared with the HEOM equivalent of Fig. 2. In this figure, the (smaller) black boxes instead show the region of time that needed to be simulated to compute the full grid.

Close modal

The agreement between the HEOM and semiclassical GQME results for the multitime correlation function Cσx(t2,t1) in Figs. 6 and 2 is excellent, even at later times. Compare, for example, the small red dots in the white “wells” of the (0, 0) element of both figures; these features represent, for example, the extent to which a quantum control sequence could be used to manipulate the coherence properties of this spin system. In contrast, Fig. 7 shows the “full MFT” multitime correlation function. To generate this figure, we do not perform the 2-time MFT simulation over the 10 × 10 Δ−1 grid, as converging it with the same resolution becomes prohibitively expensive. Instead, we employ both 1-time and 2-time GQME formulas from our previous multitime GQME work,56 which exclusively improve the efficiency but not the accuracy of semiclassics. Features such as these small dots are entirely missing in the 2-time MFT result, and the qualitative structure of many elements is incorrect [e.g., the (2, 3) element has a pattern of wells where it should be diagonal stripes], showing how critical it is to sample bath information when employing a GQME with MFT reference dynamics. Also noteworthy is that our parameter regime corresponds to weak electronic-nuclear coupling (small reorganization energy), which leads our 2-time memory kernel to offer a comparatively small correction. However, many experimental systems display intermediate to strong electronic-nuclear coupling, such as light-harvesting complexes,93,94 and small polaron-forming materials such as polymers95,96 and transition metal oxides.97,98 In such cases, a 2-time kernel that is five times too large would lead to catastrophically incorrect 2-time propagators even with exact 1-time dynamics.

FIG. 7.

GQME-extended 2-time correlation function using only the original MFT 2-time dynamics formulated to recover dynamics of MFT-level accuracy. The cutoff is 1 Δ−1. When compared with Figs. 6 and 2, significant disagreement is clear in both frequencies and amplitudes, analogous to Fig. 1(left).

FIG. 7.

GQME-extended 2-time correlation function using only the original MFT 2-time dynamics formulated to recover dynamics of MFT-level accuracy. The cutoff is 1 Δ−1. When compared with Figs. 6 and 2, significant disagreement is clear in both frequencies and amplitudes, analogous to Fig. 1(left).

Close modal

Our weak-coupling parameter regime is one where time convolutionless second-order perturbation (TCL2) theory99–101 is expected to perform well for 1-time correlation functions, that is, where max[2λ/βωc2,2λ/πωc]<1.102 There is, however, an additional Markovianity requirement for the accuracy of multitime TCL2. In particular, multitime TCL2 is consistent with the quantum regression theorem,103 so it cannot capture non-Markovian correlation104 across intermediate-time interactions.105,106 Since our semiclassical GQME correctly predicts correlation across intermediate-time interactions via the multitime kernel, our approach offers a significant advance over standard multitime TCL2 and provides a means to improve the accuracy of multitime semiclassical techniques.

In conclusion, we showed a representative example of when the MFT method cannot quantitatively or even qualitatively capture the 2-time correction to a quantum multitime correlation function reporting on light-induced interactions in the spin-boson model. We then demonstrated that our semiclassical multitime GQME dramatically improved the accuracy of the 2-time kernel by employing classical sampling of bath interactions obtained from the same qualitatively and quantitatively incorrect MFT dynamics. Computationally, the semiclassical GQME gave a modest 10/0.7 ≃ 14 times efficiency gain over the 1-time window and impressive accuracy gains that brought the semiclassical MFT into agreement with the HEOM benchmark, consistent with previous work. For the 2-time problem, the 10/0.5 = 20 times improvement was squared, meaning that to produce the figures in this paper, in which the dynamics have not yet even reached equilibrium, our multitime semiclassical GQME reduced the cost of the simulation by 400 times. Just as impressively, extending the multitime simulation to equilibrium incurs only a trivial computational cost compared with the effort to do this with either HEOM or semiclassical theory alone. Our illustrative example also recapitulates56 a general principle that since the higher-order memory kernels have lifetimes comparable to (and even along one axis, shorter than) the 1-time kernel, the multitime GQME obtains higher savings as the number of time indices in the correlation function increases. In upcoming work, we will demonstrate that a discrete-time analog of Eq. (4) can be united with the method introduced in Ref. 56 to affordably treat 3-time correlation functions. This offers a promising approach to utilizing semiclassical dynamics107,108 to efficiently obtain accurate 2D spectra for complex systems.

This work was partially supported by an Early Career Award in CPIMS program in the Chemical Sciences, Geosciences, and Biosciences Division of the Office of Basic Energy Sciences of the U.S. Department of Energy under Award No. DE-SC0024154. We also acknowledge start-up funds from the University of Colorado Boulder.

The authors have no conflicts to disclose.

Thomas Sayer: Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (equal). Andrés Montoya-Castillo: Conceptualization (lead); Supervision (lead); Writing – review & editing (equal).

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

We performed all HEOM calculations using the open-source pyrho code,109 and 2-time Ehrenfest simulations with an in-house implementation of the pure-state protocol introduced in Ref. 92. We note that while the decomposition of an off-diagonal initial condition in terms of linear combinations of subsystem density matrices is not unique, the Ehrenfest results converge to the same answer when the decomposition is done in terms of pure states.

In this paper, we choose a biased regime of ϵ = Δ with a fast bath, ωc = 7.5Δ, and weak coupling, λ = 0.075Δ, at an intermediate temperature of β = 1/Δ. We use an ohmic spectral density, which converges the MFT faster than the other common choice (Debye). We chose this regime because the memory kernel is fairly short-lived compared to the full dynamics and because the ohmic spectral density can be converged with the numerically exact reference method HEOM at a still-affordable K = 8 and L = 2 with δt = 0.01 Δ−1. This is important because the 2-time results would otherwise be too computationally expensive to converge.

The spin-boson model Hamiltonian,
(A1)
consists of system, bath, and system–bath terms. The system Hamiltonian has the form
The bath is composed of independent harmonic oscillators,
(A2)
which are antisymmetrically coupled to the two sites,
(8a)
where σz is the z Pauli matrix, q̂n and p̂n are the mass-weighted position and momentum operators for the nth harmonic oscillator in the bath, respectively, ωn is the frequency of the nth oscillator, and cn is the coupling constant to the spin. The spectroscopic initial condition we use in the Argyres–Kelley projector corresponds to an equilibrium bath described by its canonical density, ρB=eβĤB/trBeβĤB.
The spectral density, J(ω)=π2ncn2δ(ωωn)/ωn, fully determines the system–bath coupling. We adopt the common ohmic form with an exponential cutoff,
(A3)
where η is the usual Kondo parameter, ληωc/π=1π0dωJ(ω)ω is the reorganization energy that quantifies the strength of system–bath coupling, and ωc is the characteristic frequency of the bath.
To employ the mean-field Ehrenfest method to calculate correlation functions, we first perform a partial Wigner transform with respect to the bath variables, consistent with the derivation based on the quantum-classical Liouville equation,86 and then apply the standard time-dependent self-consistent field approximation under a classical treatment of the bath and a quantum treatment for the system.78,80 We adopt the protocol outlined in Ref. 40. Performing the partial Wigner transform of correlation functions for the rotated initial condition requires obtaining a Wigner transform of the commutator and anticommutator of the bath density, ρB, and the bath part of the system–bath coupling, VB. For the linear term in Eq. (10), acting the Poisson bracket on ρB,w and VB,w returns ρB,w and
(A4)
which consists of a sum over bath momenta Pj, frequencies ωj, and coupling constants cj.
For our Ehrenfest dynamics, we discretize the bath into nosc = 300 oscillators, employing the following relations:110–112 
(A5a)
(A5b)

Here, we show how to derive a closure of the self-consistently expanded multitime kernels that offers improvement in both efficiency and accuracy when using approximate semiclassical dynamics. For an explanation of the 1-time auxiliary kernels and Wigner transforms invoked in the main text, we refer the reader to Ref. 40.

Before starting to manipulate the 2-time kernel, we show that QLALSBA and ALQALSB. First, we perform the common splitting of the Hamiltonian into three parts [Eq. (A1)]:
(B1)
Since LS is a system operator,
(B2)
where Z is the static rotation between the system observables. Hence, the resulting generalized vector is part of the projector, P, such that its product with Q annihilates the state as PQ=QP=0. Since LB is a bath operator, it commutes with both the system operators, |j⟩⟨k|, and the canonical bath density, ρB=eβĤB/trBeβĤB. Hence,
(B3)
Finally, QLSBA is expanded as LSBAA(A|LSB|A) and we see immediately that the expectation of LSB removes the second term since ρB is quadratic and VB is linear, averaging to zero. The same logic up to the absence of any bath components holds for ALQ.
We can now obtain one of the key results in the paper: Eq. (13). Starting from Eq. (12), we use the idempotency property, Q=Q2, to bring down a Q to the right of the left-most L. This enables us to maintain the correct recursive structure by invoking the Dyson expansion113 on the left propagator:
(B4)
(B5)
(B6)
Thus, the 1-time auxiliary kernel, K(3b), appears where previously56 one required derivatives of the exact dynamics. We then invoke the Dyson expansion on the right-most propagator of KLσx(t2,t1):
(B7)
(B8)
(B9)
which removes the last remaining projected propagator. We insert the definition of Q=1P for the two central complementary projectors and replace the outer instances that touch L with LSB as planned,
(B10)
(B11)
where
(B12)
Here, the superscript σx denotes the modified kernels where the measurement occurs at the start or end of the propagation, as defined in the equation. Note that all the K(3f) and K(3b) that appear here also have the LSB substitutions. Only the first term of Eq. (B11) requires 2-time information. This object is the 2-time analog for the 1-time K(1),
(B13)
(B14)
(B15)
where the matrices X and Y are the simple rotations from the terms involving σz.40 We use the commutator and anticommutator (denoted with a subscript +) identities because this neatly separates the VB,w and ζB,w terms upon Wigner transformation as Eq. (13) of the main text.
1.
S.
Mukamel
,
Principles of Nonlinear Optical Spectroscopy
(
Oxford University Press
,
New York
,
1999
).
2.
L.
Viola
,
E.
Knill
, and
S.
Lloyd
, “
Dynamical decoupling of open quantum systems
,”
Phys. Rev. Lett.
82
,
2417
(
1999
).
3.
C.
Brif
,
R.
Chakrabarti
, and
H.
Rabitz
, “
Control of quantum phenomena: Past, present and future
,”
New J. Phys.
12
,
075008
(
2010
).
4.
M.
Shapiro
and
P.
Brumer
, in
Quantum Control of Molecular Processes
, 2nd ed. (
Wiley VCH
,
2012
).
5.
V. G.
Sadhasivam
,
L.
Meuser
,
D. R.
Reichman
, and
S. C.
Althorpe
, “
Instantons and the quantum bound to chaos
,”
Proc. Natl. Acad. Sci. U. S. A.
120
,
e2312378120
(
2023
).
6.
C.
Zhang
,
P. G.
Wolynes
, and
M.
Gruebele
, “
Quantum information scrambling in molecules
,”
Phys. Rev. A
105
,
033322
(
2022
).
7.
C.
Zhang
,
S.
Kundu
,
N.
Makri
,
M.
Gruebele
, and
P. G.
Wolynes
, “
Quantum information scrambling and chemical reactions
,”
Proc. Natl. Acad. Sci. U. S. A.
121
,
e2321668121
(
2024
).
8.
S. H.
Shenker
and
D.
Stanford
, “
Black holes and the butterfly effect
,”
J. High Energy Phys.
2014
,
67
.
9.
R. Q.
He
and
Z. Y.
Lu
, “
Characterizing many-body localization by out-of-time-ordered correlation
,”
Phys. Rev. B
95
,
054201
(
2017
).
10.
Y.
Tanimura
and
R.
Kubo
, “
Time evolution of a quantum system in contact with a nearly Gaussian–Markoffian noise bath
,”
J. Phys. Soc. Jpn.
58
,
101
114
(
1989
).
11.
H. D.
Meyer
,
U.
Manthe
, and
L. S.
Cederbaum
, “
The multi-configurational time-dependent Hartree approach
,”
Chem. Phys. Lett.
165
,
73
78
(
1990
).
12.
N.
Makri
, “
Numerical path integral techniques for long time dynamics of quantum dissipative systems
,”
J. Math. Phys.
36
,
2430
2457
(
1995
).
13.
M.
Thoss
,
H.
Wang
, and
W. H.
Miller
, “
Self-consistent hybrid approach for complex systems: Application to the spin-boson model with Debye spectral density
,”
J. Chem. Phys.
115
,
2991
3005
(
2001
).
14.
H.
Wang
and
M.
Thoss
, “
Multilayer formulation of the multiconfiguration time-dependent Hartree theory
,”
J. Chem. Phys.
119
,
1289
1299
(
2003
).
15.
A.
Ishizaki
and
Y.
Tanimura
, “
Quantum dynamics of system strongly coupled to low-temperature colored noise bath: Reduced hierarchy equations approach
,”
J. Phys. Soc. Jpn.
74
,
3131
3134
(
2005
).
16.
D.
Suess
,
A.
Eisfeld
, and
W. T.
Strunz
, “
Hierarchy of stochastic pure states for open quantum system dynamics
,”
Phys. Rev. Lett.
113
,
150403
(
2014
).
17.
D.
Tamascelli
,
A.
Smirne
,
J.
Lim
,
S. F.
Huelga
, and
M. B.
Plenio
, “
Efficient simulation of finite-temperature open quantum systems
,”
Phys. Rev. Lett.
123
,
090402
(
2019
).
18.
D. W.
Small
,
D. V.
Matyushov
, and
G. A.
Voth
, “
The theory of electron transfer reactions: What may be missing?
,”
J. Am. Chem. Soc.
125
,
7470
7478
(
2003
).
19.
M. M.
Waskasi
,
D. R.
Martin
, and
D. V.
Matyushov
, “
Wetting of the protein active site leads to non-Marcusian reaction kinetics
,”
J. Phys. Chem. B
122
,
10490
10495
(
2018
).
20.
D. V.
Matyushov
and
M. D.
Newton
, “
Q-model of electrode reactions: Altering force constants of intramolecular vibrations
,”
Phys. Chem. Chem. Phys.
20
,
24176
24185
(
2018
).
21.
D. T.
Limmer
,
C.
Merlet
,
M.
Salanne
,
D.
Chandler
,
P. A.
Madden
,
R.
Van Roij
, and
B.
Rotenberg
, “
Charge fluctuations in nanoscale capacitors
,”
Phys. Rev. Lett.
111
,
106102
(
2013
).
22.
J.
Bredenbeck
,
J.
Helbing
, and
P.
Hamm
, “
Solvation beyond the linear response regime
,”
Phys. Rev. Lett.
95
,
083201
(
2005
).
23.
P.
Hamm
, “
Three-dimensional-IR spectroscopy: Beyond the two-point frequency fluctuation correlation function
,”
J. Chem. Phys.
124
,
124506
(
2006
).
24.
D. M.
Packwood
and
Y.
Tanimura
, “
Non-Gaussian stochastic dynamics of spins and oscillators: A continuous-time random walk approach
,”
Phys. Rev. E
84
,
061111
(
2011
).
25.
C. L.
Degen
,
F.
Reinhard
, and
P.
Cappellaro
, “
Quantum sensing
,”
Rev. Mod. Phys.
89
,
035002
(
2017
).
26.
Y. M.
Blanter
and
M.
Büttiker
, “
Shot noise in mesoscopic conductors
,”
Phys. Rep.
336
,
1
166
(
2000
).
27.
B.
Huard
,
H.
Pothier
,
N.
Birge
,
D.
Esteve
,
X.
Waintal
, and
J.
Ankerhold
, “
Josephson junctions as detectors for non-Gaussian noise
,”
Ann. Phys.
519
(
10-11
),
736
750
(
2007
).
28.
T. P.
Purdy
,
R. W.
Peterson
, and
C. A.
Regal
, “
Observation of radiation pressure shot noise on a macroscopic object
,”
Science
339
,
801
804
(
2013
).
29.
K.
Kwac
and
E.
Geva
, “
A mixed quantum-classical molecular dynamics study of the hydroxyl stretch in methanol/carbon tetrachloride mixtures III: Nonequilibrium hydrogen-bond dynamics and infrared pump-probe spectra
,”
J. Phys. Chem. B
117
,
7737
7749
(
2013
).
30.
J.
Janoš
,
J. P.
Figueira Nunes
,
D.
Hollas
,
P.
Slavíček
, and
B. F.
Curchod
, “
Predicting the photodynamics of cyclobutanone triggered by a laser pulse at 200 nm and its MeV-UED signals—A trajectory surface hopping and XMS-CASPT2 perspective
,”
J. Chem. Phys.
160
,
144305
(
2024
).
31.
Q.
Shi
and
E.
Geva
, “
A comparison between different semiclassical approximations for optical response functions in nonpolar liquid solutions
,”
J. Chem. Phys.
122
,
64506
(
2005
).
32.
S.
Nakajima
, “
On quantum theory of transport phenomena: Steady diffusion
,”
Prog. Theor. Phys.
20
,
948
959
(
1958
).
33.
R.
Zwanzig
, “
Ensemble method in the theory of irreversibility
,”
J. Chem. Phys.
33
,
1338
1341
(
1960
).
34.
H.
Mori
, “
Transport, collective motion, and brownian motion
,”
Prog. Theor. Phys.
33
,
423
455
(
1965
).
35.
Q.
Shi
and
E.
Geva
, “
A new approach to calculating the memory kernel of the generalized quantum master equation for an arbitrary system–bath coupling
,”
J. Chem. Phys.
119
,
12063
12076
(
2003
).
36.
Q.
Shi
and
E.
Geva
, “
A semiclassical generalized quantum master equation for an arbitrary system-bath coupling
,”
J. Chem. Phys.
120
,
10647
10658
(
2004
).
37.
M. L.
Zhang
,
B. J.
Ka
, and
E.
Geva
, “
Nonequilibrium quantum dynamics in the condensed phase via the generalized quantum master equation
,”
J. Chem. Phys.
125
,
044106
(
2006
).
38.
A.
Kelly
,
N.
Brackbill
, and
T. E.
Markland
, “
Accurate nonadiabatic quantum dynamics on the cheap: Making the most of mean field theory with master equations
,”
J. Chem. Phys.
142
,
94110
(
2015
).
39.
A.
Kelly
and
T. E.
Markland
, “
Efficient and accurate surface hopping for long time nonadiabatic quantum dynamics
,”
J. Chem. Phys.
139
,
14104
(
2013
).
40.
A.
Montoya-Castillo
and
D. R.
Reichman
, “
Approximate but accurate quantum dynamics from the Mori formalism: I. Nonequilibrium dynamics
,”
J. Chem. Phys.
144
,
184104
(
2016
).
41.
A.
Kelly
,
A.
Montoya-Castillo
,
L.
Wang
, and
T. E.
Markland
, “
Generalized quantum master equations in and out of equilibrium: When can one win?
,”
J. Chem. Phys.
144
,
184105
(
2016
).
42.
E.
Mulvihill
,
A.
Schubert
,
X.
Sun
,
B. D.
Dunietz
, and
E.
Geva
, “
A modified approach for simulating electronically nonadiabatic dynamics via the generalized quantum master equation
,”
J. Chem. Phys.
150
,
34101
(
2019
).
43.
E.
Mulvihill
,
X.
Gao
,
Y.
Liu
,
A.
Schubert
,
B. D.
Dunietz
, and
E.
Geva
, “
Combining the mapping Hamiltonian linearized semiclassical approach with the generalized quantum master equation to simulate electronically nonadiabatic molecular dynamics
,”
J. Chem. Phys.
151
,
74103
(
2019
).
44.
N.
Ng
,
D. T.
Limmer
, and
E.
Rabani
, “
Nonuniqueness of generalized quantum master equations for a single observable
,”
J. Chem. Phys.
155
,
156101
(
2021
).
45.
G.
Cohen
,
E. Y.
Wilner
, and
E.
Rabani
, “
Generalized projected dynamics for non-system observables of non-equilibrium quantum impurity models
,”
New J. Phys.
15
,
073018
(
2013
).
46.
G.
Cohen
,
E.
Gull
,
D. R.
Reichman
,
A. J.
Millis
, and
E.
Rabani
, “
Numerically exact long-time magnetization dynamics at the nonequilibrium Kondo crossover of the Anderson impurity model
,”
Phys. Rev. B
87
,
195108
(
2013
).
47.
L.
Kidon
,
E. Y.
Wilner
, and
E.
Rabani
, “
Exact calculation of the time convolutionless master equation generator: Application to the nonequilibrium resonant level model
,”
J. Chem. Phys.
143
,
234110
(
2015
).
48.
L.
Kidon
,
H.
Wang
,
M.
Thoss
, and
E.
Rabani
, “
On the memory kernel and the reduced system propagator
,”
J. Chem. Phys.
149
,
104105
(
2018
).
49.
W. C.
Pfalzgraff
,
A.
Kelly
, and
T. E.
Markland
, “
Nonadiabatic dynamics in atomistic environments: Harnessing quantum-classical theory with generalized quantum master equations
,”
J. Phys. Chem. Lett.
6
,
4743
4748
(
2015
).
50.
W. C.
Pfalzgraff
,
A.
Montoya-Castillo
,
A.
Kelly
, and
T. E.
Markland
, “
Efficient construction of generalized master equation memory kernels for multi-state systems from nonadiabatic quantum-classical dynamics
,”
J. Chem. Phys.
150
,
244109
(
2019
).
51.
E.
Mulvihill
,
K. M.
Lenn
,
X.
Gao
,
A.
Schubert
,
B. D.
Dunietz
, and
E.
Geva
, “
Simulating energy transfer dynamics in the Fenna–Matthews–Olson complex via the modified generalized quantum master equation
,”
J. Chem. Phys.
154
,
204109
(
2021
).
52.
A.
Montoya-Castillo
and
D. R.
Reichman
, “
Approximate but accurate quantum dynamics from the Mori formalism. II. Equilibrium time correlation functions
,”
J. Chem. Phys.
146
,
084110
(
2017
).
53.
X.
Gao
and
E.
Geva
, “
A nonperturbative methodology for simulating multidimensional spectra of multiexcitonic molecular systems via quasiclassical mapping Hamiltonian methods
,”
J. Chem. Theory Comput.
16
,
6491
6502
(
2020
).
54.
E.
Fick
and
G.
Sauermann
,
The Quantum Statistics of Dynamic Processes
,
Springer Series in Solid-State Sciences
(
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
1990
), Vol.
86
55.
H.
Grabert
,
Projection Operator Techniques in Nonequilibrium Statistical Mechanics
(
Springer Berlin Heidelberg
,
2006
), pp.
166
.
56.
T.
Sayer
and
A.
Montoya-Castillo
, “
Efficient formulation of multitime generalized quantum master equations: Taming the cost of simulating 2D spectra
,”
J. Chem. Phys.
160
,
44108
(
2024
).
57.
T. L. C.
Jansen
and
J.
Knoester
, “
Nonadiabatic effects in the two-dimensional infrared spectra of peptides: Application to alanine dipeptide
,”
J. Phys. Chem. B
110
,
22910
22916
(
2006
).
58.
C.
Liang
and
T. L.
Jansen
, “
An efficient N3-scaling propagation scheme for simulating two-dimensional infrared and visible spectra
,”
J. Chem. Theory Comput.
8
,
1706
1713
(
2012
).
59.
R.
Tempelaar
,
C. P.
Van Der Vegte
,
J.
Knoester
, and
T. L.
Jansen
, “
Surface hopping modeling of two-dimensional spectra
,”
J. Chem. Phys.
138
,
164106
(
2013
).
60.
C. P.
Van Der Vegte
,
A. G.
Dijkstra
,
J.
Knoester
, and
T. L.
Jansen
, “
Calculating two-dimensional spectra with the mixed quantum-classical Ehrenfest method
,”
J. Phys. Chem. A
117
,
5970
5980
(
2013
).
61.
R. F.
Loring
, “
Mean-trajectory approximation for electronic and vibrational-electronic nonlinear spectroscopy
,”
J. Chem. Phys.
146
,
144106
(
2017
).
62.
J.
Provazza
,
F.
Segatta
,
M.
Garavelli
, and
D. F.
Coker
, “
Semiclassical path integral calculation of nonlinear optical spectroscopy
,”
J. Chem. Theory Comput.
14
,
856
866
(
2018
).
63.
J. R.
Mannouch
and
J. O.
Richardson
, “
A partially linearized spin-mapping approach for simulating nonlinear optical spectra
,”
J. Chem. Phys.
156
,
24108
(
2022
).
64.
R. F.
Loring
, “
Calculating multidimensional optical spectra from classical trajectories
,”
Annu. Rev. Phys. Chem.
73
,
273
297
(
2022
).
65.
T.
Begušić
,
X.
Tao
,
G. A.
Blake
, and
T. F.
Miller
, “
Equilibrium-nonequilibrium ring-polymer molecular dynamics for nonlinear spectroscopy
,”
J. Chem. Phys.
156
,
131102
(
2022
).
66.
J.
Cerrillo
and
J.
Cao
, “
Non-Markovian dynamical maps: Numerical processing of open quantum trajectories
,”
Phys. Rev. Lett.
112
,
110401
(
2014
).
67.
E. J.
Heller
, “
Time-dependent approach to semiclassical dynamics
,”
J. Chem. Phys.
62
,
1544
1555
(
1975
).
68.
M. F.
Herman
, “
Dynamics by semiclassical methods
,”
Annu. Rev. Phys. Chem.
45
,
83
111
(
1994
).
69.
G.
Stock
and
M.
Thoss
, “
Semiclassical description of nonadiabatic quantum dynamics
,”
Phys. Rev. Lett.
78
,
578
(
1997
).
70.
X.
Sun
and
W. H.
Miller
, “
Semiclassical initial value representation for electronically nonadiabatic molecular dynamics
,”
J. Chem. Phys.
106
,
6346
6353
(
1997
).
71.
J.
Shao
and
N.
Makri
, “
Forward–backward semiclassical dynamics without prefactors
,”
J. Phys. Chem. A
103
,
7753
7756
(
1999
).
72.
M.
Thoss
and
H.
Wang
, “
Semiclassical description of molecular dynamics based on initial-value representation methods
,”
Annu. Rev. Phys. Chem.
55
,
299
332
(
2004
).
73.
S.
Bonella
and
D. F.
Coker
, “
LAND-map, a linearized approach to nonadiabatic dynamics using the mapping formalism
,”
J. Chem. Phys.
122
,
194102
(
2005
).
74.
W. H.
Miller
, “
Electronically nonadiabatic dynamics via semiclassical initial value methods
,”
J. Phys. Chem. A
113
,
1405
1415
(
2009
).
75.
P.
Huo
and
D. F.
Coker
, “
Communication: Partial linearized density matrix dynamics for dissipative, non-adiabatic quantum evolution
,”
J. Chem. Phys.
135
,
201101
(
2011
).
76.
J. O.
Richardson
and
M.
Thoss
, “
Communication: Nonadiabatic ring-polymer molecular dynamics
,”
J. Chem. Phys.
139
,
31102
(
2013
).
77.
N.
Ananth
, “
Path integrals for nonadiabatic dynamics: Multistate ring polymer molecular dynamics
,”
Annu. Rev. Phys. Chem.
73
,
299
322
(
2022
).
78.
A. D.
McLachlan
, “
A variational solution of the time-dependent Schrodinger equation
,”
Mol. Phys.
8
,
39
44
(
1964
).
79.
J. C.
Tully
, “
Molecular dynamics with electronic transitions
,”
J. Chem. Phys.
93
,
1061
1071
(
1990
).
80.
G.
Stock
, “
A semiclassical self-consistent-field approach to dissipative dynamics: The spin–boson problem
,”
J. Chem. Phys.
103
,
1561
1573
(
1995
).
81.
S.
Hammes-Schiffer
, “
Multiconfigurational molecular dynamics with quantum transitions: Multiple proton transfer reactions
,”
J. Chem. Phys.
105
,
2236
2246
(
1996
).
82.
R.
Kapral
and
G.
Ciccotti
, “
Mixed quantum-classical dynamics
,”
J. Chem. Phys.
110
,
8919
8929
(
1999
).
83.
Q.
Shi
and
E.
Geva
, “
A derivation of the mixed quantum-classical Liouville equation from the influence functional formalism
,”
J. Chem. Phys.
121
,
3393
3404
(
2004
).
84.
H.
Kim
,
A.
Nassimi
, and
R.
Kapral
, “
Quantum-classical Liouville dynamics in the mapping basis
,”
J. Chem. Phys.
129
,
84102
(
2008
).
85.
C. Y.
Hsieh
and
R.
Kapral
, “
Nonadiabatic dynamics in open quantum-classical systems: Forward-backward trajectory solution
,”
J. Chem. Phys.
137
,
22
507
(
2012
).
86.
R.
Kapral
, “
Quantum dynamics in open quantum-classical systems
,”
J. Phys.: Condens. Matter
27
,
073201
(
2015
).
87.
J. E.
Subotnik
,
A.
Jain
,
B.
Landry
,
A.
Petit
,
W.
Ouyang
, and
N.
Bellonzi
, “
Understanding the surface hopping view of electronic transitions and decoherence
,”
Annu. Rev. Phys. Chem.
67
,
387
417
(
2016
).
88.
J. E.
Runeson
and
D. E.
Manolopoulos
, “
A multi-state mapping approach to surface hopping
,”
J. Chem. Phys.
159
,
94115
(
2023
).
89.
K.
Imre
,
E.
Özizmir
,
M.
Rosenbaum
, and
P. F.
Zweifel
, “
Wigner method in quantum statistical mechanics
,”
J. Math. Phys.
8
,
1097
1108
(
1967
).
90.
M.
Hillery
,
R. F.
O’Connell
,
M. O.
Scully
, and
E. P.
Wigner
, “
Distribution functions in physics: Fundamentals
,”
Phys. Rep.
106
,
121
167
(
1984
).
91.
A.
Ivanov
and
H. P.
Breuer
, “
Extension of the Nakajima–Zwanzig approach to multitime correlation functions of open systems
,”
Phys. Rev. A
92
,
032113
(
2015
).
92.
A. O.
Atsango
,
A.
Montoya-Castillo
, and
T. E.
Markland
, “
An accurate and efficient Ehrenfest dynamics approach for calculating linear and nonlinear electronic spectra
,”
J. Chem. Phys.
158
,
74107
(
2023
).
93.
T.
Mirkovic
,
E. E.
Ostroumov
,
J. M.
Anna
,
R.
van Grondelle
,
G. D.
Scholes
, and
G. D.
Scholes
, “
Light absorption and energy transfer in the antenna complexes of photosynthetic organisms
,”
Chem. Rev.
117
,
249
293
(
2017
).
94.
S. J.
Jang
and
B.
Mennucci
, “
Delocalized excitons in natural light-harvesting complexes
,”
Rev. Mod. Phys.
90
,
035003
(
2018
).
95.
R.
Ghosh
and
F. C.
Spano
, “
Excitons and polarons in organic materials
,”
Acc. Chem. Res.
53
,
2201
2211
(
2020
).
96.
T.
Nematiaram
and
A.
Troisi
, “
Modeling charge transport in high-mobility molecular semiconductors: Balancing electronic structure and quantum dynamics methods with the help of experiments
,”
J. Chem. Phys.
152
,
190902
(
2020
).
97.
N.
Iordanova
,
M.
Dupuis
, and
K. M.
Rosso
, “
Charge transport in metal oxides: A theoretical study of hematite α-Fe2O3
,”
J. Chem. Phys.
122
,
144305
(
2005
).
98.
C.
Franchini
,
M.
Reticcioli
,
M.
Setvin
, and
U.
Diebold
, “
Polarons in materials
,”
Nat. Rev. Mater.
6
,
560
586
(
2021
).
99.
F.
Bloch
, “
Generalized theory of relaxation
,”
Phys. Rev.
105
,
1206
(
1957
).
100.
A. G.
Redfield
, “
The theory of relaxation processes
,”
Adv. Magn. Reson.
1
,
1
32
(
1965
).
101.
T. M.
Chang
and
J. L.
Skinner
, “
Non-Markovian population and phase relaxation and absorption lineshape for a two-level system strongly coupled to a harmonic quantum bath
,”
Physica A
193
,
483
539
(
1993
).
102.
A.
Montoya-Castillo
,
T. C.
Berkelbach
, and
D. R.
Reichman
, “
Extending the applicability of Redfield theories into highly non-Markovian regimes
,”
J. Chem. Phys.
143
,
194108
(
2015
).
103.
M.
Lax
, “
Formal theory of quantum fluctuations from a driven state
,”
Phys. Rev.
129
,
2342
(
1963
).
104.
B. J.
Ka
and
E.
Geva
, “
A nonperturbative calculation of nonlinear spectroscopic signals in liquid solution
,”
J. Chem. Phys.
125
,
214501
(
2006
).
105.
A.
Ishizaki
and
Y.
Tanimura
, “
Nonperturbative non-Markovian quantum master equation: Validity and limitation to calculate nonlinear response functions
,”
Chem. Phys.
347
,
185
193
(
2008
).
106.
J. H.
Fetherolf
and
T. C.
Berkelbach
, “
Linear and nonlinear spectroscopy from quantum master equations
,”
J. Chem. Phys.
147
,
244109
(
2017
).
107.
G.
Hanna
and
E.
Geva
, “
Computational study of the one and two dimensional infrared spectra of a vibrational mode strongly coupled to its environment: Beyond the cumulant and condon approximations
,”
J. Phys. Chem. B
112
,
12991
13004
(
2008
).
108.
P. L.
McRobbie
,
G.
Hanna
,
Q.
Shi
, and
E.
Geva
, “
Signatures of nonequilibrium solvation dynamics on multidimensional spectra
,”
Acc. Chem. Res.
42
,
1299
1309
(
2009
).
109.
T.
Berkelbach
,
J.
Fetherolf
,
P.
Shih
, and
Iansdunn
(
2020
). “
berkelbach-group/pyrho v1.0
,” https://github.com/berkelbach-group/pyrho.
110.
N.
Makri
, “
The linear response approximation and its lowest order corrections: An influence functional approach
,”
J. Phys. Chem. B
103
,
2823
2829
(
1999
).
111.
I. R.
Craig
and
D. E.
Manolopoulos
, “
Chemical reaction rates from ring polymer molecular dynamics
,”
J. Chem. Phys.
122
,
84106
(
2005
).
112.
A.
Montoya-Castillo
and
D. R.
Reichman
, “
Path integral approach to the Wigner representation of canonical density operators for discrete systems coupled to harmonic baths
,”
J. Chem. Phys.
146
,
24107
(
2017
).
113.
R. P.
Feynman
, “
An operator calculus having applications in quantum electrodynamics
,”
Phys. Rev.
84
,
108
(
1951
).