On the Proper Derivation of the Floquet-based Quantum Classical Liouville Equation and Surface Hopping Describing a Molecule or Material Subject to an External Field
Abstract
We investigate different approaches to derive the proper Floquet-based quantum-classical Liouville equation (F-QCLE) for laser-driven electron-nuclear dynamics. The first approach projects the operator form of the standard QCLE onto the diabatic Floquet basis, then transforms to the adiabatic representation. The second approach directly projects the QCLE onto the Floquet adiabatic basis. Both approaches yield a form which is similar to the usual QCLE with two modifications: 1. The electronic degrees of freedom are expanded to infinite dimension. 2. The nuclear motion follows Floquet quasi-energy surfaces. However, the second approach includes an additional cross derivative force due to the dual dependence on time and nuclear motion of the Floquet adiabatic states. Our analysis and numerical tests indicate that this cross derivative force is a fictitious artifact, suggesting that one cannot safely exchange the order of Floquet state projection with adiabatic transformation. Our results are in accord with similar findings by Izmaylov et al. (Ref. 41), who found that transforming to the adiabatic representation must always be the last operation applied, though now we have extended this result to a time-dependent Hamiltonian. This paper and the proper derivation of the F-QCLE should lay the basis for further improvements of Floquet surface hopping.
I Introduction
A computational understanding light-matter interactions for a molecular system in a laser field is useful for interpreting spectroscopy and photochemistry, where the dynamical interplay between electronic non-adiabatic transitions and photon excitation plays an important role for many exciting phenomena, such as molecular photodissociation(Regan et al., 1999; Franks, Li, and Kong, 1999; Hilsabeck et al., 2019a, b) and coherent X-ray diffraction(Glownia et al., 2016, 2017; Lemke et al., 2017; Fuller et al., 2017). These phenomena usually involve dynamical processes in which electrons in a molecular system can make a transition through either (a) non-adiabatic coupling associated with the reorganization of nuclear configurations or (b) radiative coupling in conjunction with absorption or emission of photons. Thus, simulating these processes concurrently requires accurate theoretical treatments of both non-adiabatic molecular dynamics and light-matter interactions(Bajo, Granucci, and Persico, 2014; Hoffmann et al., 2018; Abedi, Maitra, and Gross, 2010). Over the past decades, many successful simulation schemes have been developed based on mixed quantum−classical frameworks in which the electronic wavefunction evolves according to quantum mechanics while the nuclear degrees of freedom and the laser excitation are treated as classical parameters in a time-dependent electronic Hamiltonian(Zhou et al., 2020; Richter et al., 2011; Mitrić, Petersen, and Bonačić-Koutecký, 2009; Mai, Marquetand, and González, 2018; Thachuk, Ivanov, and Wardlaw, 1996; Bajo et al., 2012; Lisinetskaya and Mitrić, 2011).
Among the myriad of semiclassical dynamics, Floquet-based fewest switch surface hopping (F-FSSH) has emerged as one of the most powerful methods especially for simulating photodissociation and ionization in a monochromatic laser field(Fiedlschuster, Handt, and Schmidt, 2016; Fiedlschuster et al., 2017; Horenko, Schmidt, and Schütte, 2001). In a nutshell, F-FSSH integrates Floquet theory with Tully’s FSSH algorithm(Tully, 2012). The general idea is to expand the electronic wavefunction in a Floquet state basis (with the electronic states dressed by for an integer and the laser frequency ), so that one can recast an explicitly time-dependent Hamiltonian into a time-independent Floquet Hamiltonian, albeit of infinite dimension. With the Floquet Hamiltonian, one can simply employ Tully’s FSSH method in the Floquet state representation with a minimal modification(Fiedlschuster, Handt, and Schmidt, 2016; Fiedlschuster et al., 2017). In addition to the standard advantages of the usual surface hopping algorithm—stability, efficiency, and ease of incorporation with electronic structure calculations—F-FSSH also yield a better estimate for both electronic and nuclear observables than other FSSH-based methods relying on instantaneous adiabatic surfaces(Zhou et al., 2020). Furthermore, given the time-independent nature of the Floquet Hamiltonian, many techniques designed to improve standard FSSH method, such as velocity reversal and decoherence(Bittner and Rossky, 1995; Jasper and Truhlar, 2005, 2003; Nelson et al., 2013; Subotnik et al., 2016; Jain, Alguire, and Subotnik, 2016; Fang and Hammes-Schiffer, 1999; Bedard-Hearn, Larsen, and Schwartz, 2005), should be applicable within the Floquet formalism. That being said, due to the fact that one cannot directly derive Tully’s FSSH—but rather only indirectly connect the equations of motion for FSSH dynamics with the quantum-classical Liouville equation (QCLE) in the adiabatic basis(Subotnik, Ouyang, and Landry, 2013; Kapral, 2016)—a proper understanding of the correct F-QCLE is crucial if we aim to make future progress in semiclassical modeling of light-matter interactions. Moreover, since it is well known that (1) FSSH performs best in an adiabatic (rather than diabatic) basis(Tully, 2012; Jasper et al., 2004), and (2) FSSH can be connected to the QCLE only in an adiabatic basis(Subotnik, Ouyang, and Landry, 2013; Kapral, 2016), we emphasize that, for the sake of future progress, we require a proper understanding of the F-QCLE in an adiabatic basis.
Unfortunately, even without a light field, the proper derivation of the correct QCLE in an adiabatic basis is tricky—one can find two different versions of the QCLE if one invokes slightly different formal derivations. Following Kapral’s approach(Kapral, 2006; Kapral and Ciccotti, 1999), the proper derivation of the standard QCLE includes two operations: (i) Wigner transformation and (ii) projection onto the adiabatic electronic state basis. First, one performs a partial Wigner transformation with respect to the nuclear degrees of freedom to obtain the operator form of the QCLE. Wigner transformation provides an exact framework to interpret the full quantum density matrix in terms of the joint electronic-nuclear probability density in the phase space of the nuclear configuration while retaining the quantum operator character of the electronic subsystem. Second, one projects the operator form of the QCLE onto the adiabatic electronic states basis obtained by diagonalizing the electronic Hamiltonian. This adiabatic representation allows the connection to electronic structure calculations in a mixed quantum-classical sense(Wong and Rossky, 2002a, b; Webster, Rossky, and Friesner, 1991; Kelly and Markland, 2013; Kim et al., 2012). This approach is called the Wigner-then-Adiabatic (WA) approach. As Izmaylov and co-workers have shown(Ryabinkin et al., 2014), however, exchanging these two operations (the Adiabatic-then-Wigner (AW) approach) leads to a different QCLE that cannot capture geometric phase effects arising from a conical intersection (even though short time dynamcis can sometimes be accurate(Ando, 2002; Ando and Santer, 2003)).
With this background in mind, the proper derivation of the F-QCLE is now even more challenging. In addition to the two operations above, there is a third step: one needs to dress the electronic states and expand the density matrix in the Floquet state basis. In the literature to date, the F-QCLE has been derived via the AW approach (projecting in the Floquet adiabatic representation, and then performing partial Wigner transformation)(Horenko, Schmidt, and Schütte, 2001). Nevertheless, as shown by Izmaylov and co-workers(Ryabinkin et al., 2014; Kapral and Ciccotti, 1999), even in the limit of a time-independent Hamiltonian, such an (incorrect) AW approach will lead to a QCLE that neglects geometric phase related features in the nuclear dynamics or introduces artificial nuclear effects. Despite recent progress, the proper derivation of the F-QCLE is still an open question.
In this paper, our goal is to explore different approaches to derive the F-QCLE as we shuffle the three key operations and quantify their differences in the context of driven electron-nuclear dynamics. By isolating the correct F-QCLE, our work will not only validate F-FSSH methods, it should also provide means to improve surface hopping methods. This paper is organized as follows. In Sec. II, we formulate the three operations that are required to derive the F-QCLE. In Sec. III, we derive F-QCLEs via approaches with different ordering of operations. In Sec. IV, we implement F-FSSH calculations corresponding to these F-QCLEs and analyze their results for a modified avoided crossing model. We conclude with an outlook for the future in Sec. V.
For notation, we denote a quantum operator by and use bold font for matrix . We use to denote the corresponding matrix in the expanded Floquet basis (infinite dimensional). The nuclear position and momentum are , where is the nuclear coordinate index. We use a shorthand notation for dot product: .
II Three Operations
In the context of driven electron-nuclear dynamics, let us formulate the three necessary operations for deriving F-QCLE. For a coupled electron-nuclei system driven by an external field of frequency , the total Hamiltonian takes the form of where is the nuclear kinetic energy and is the electronic Hamiltonian with explicit time periodicity with . Formally, the dynamics of the total system can be described by the time-dependent Schrödinger equation (TDSE) of the total wavefunction or the quantum Liouville equation (QLE) of the total density matrix . To derive F-QCLE, we need to apply the following three operations to the QLE.
II.1 Partial Wigner transformation of the nuclear degrees of freedom
To describe the dynamics in a mixed quantum-classical sense, we will follow Kapral’s approach and perform a partial Wigner transformation with respect to the nuclear degrees of freedom
(1) |
where is the dimension of the nuclear coordinate. A nuclear position eigenstate is defined as . In Eq. (1), the density matrix operator has been transformed into a Wigner wavepacket in phase space with coordinates . In what follows, we will denote the partial Wigner transformed operator by the subscript [for example ]. Note that, after the partial Wigner transformation, and remain electronic operators while and are parameters.
The equation of motion of the Wigner wavepacket can be obtained by transforming the QLE by . The Wigner transform of operator products can be expanded further by the Wigner–Moyal operator with (Kapral and Ciccotti, 1999). Then, if we expand the the Wigner–Moyal operator in Taylor series and truncate to the first order of , we obtain the operator form of the QCLE
(2) |
Here, the commutator is and the anti-commutator is . Note that Eq. (2) is exact if the partial Wigner transformed Hamiltonian is quadratic in , for example harmonic oscillators.
To propagate the Wigner wavepacket in Eq. (2), one must project the operator form of the QCLE in an electronic basis. One straightforward choice is to use a complete set of diabatic states for the electronic subsystem ; such a set does not depend on any nuclear configuration. With this electronic diabatic basis, one can derive equations of motion for the density matrix () using matrix elements of the electronic Hamiltonian, . However, for many realistic electron-nuclei systems (and certainly any ab initio calculations), this diabatic QCLE cannot be solved since finding a complete set of exactly diabatic electronic states over a large set of nuclear geometries is rigorously impossible and quite demanding in practice even for approximate diabats.
II.2 Dress the electronic basis in the Floquet formalism
Let us now focus on the Floquet formalism, according to which one solves the TDSE by transforming the time-dependent Hamiltonian into a time-independent Floquet Hamiltonian in an extended Hilbert space of infinite dimension. For the moment let us ignore all nuclear motion and focus on the electronic subsystem exclusively. According to Floquet theory, we utilize the time periodicity of the electronic Hamiltonian and dress the electronic diabatic states by a time-periodic function where is an integer formally from to . We denote the dressed state as the the Floquet diabatic state . In terms of the Floquet diabatic basis, a time periodic electronic wavefunction can be expressed as where is an infinite dimensional state vector. The electronic wavefunction coefficient must satisfy the electronic TDSE
(3) |
where the Floquet Hamiltonian operator is defined as
(4) |
Next, we close Eq. 3 by multiplying both sides by and write as a Fourier series:
(5) |
Thus, the TDSE in Eq. (3) can be solved by grouping together all terms with the same time dependence, leading to the equation of motion for
(6) |
The matrix elements of the Floquet Hamiltonian can be obtained by performing a Fourier transformation on the matrix elements
(7) |
Here, we define the double bracket projection of an electronic operator by . Given that the electronic Hamiltonian operator is periodic in time, the double-bracket projection eliminates all time dependence and the Floquet Hamiltonian matrix reads
(8) |
In the end, with this time-independent Hamiltonian, Eq. (6) can be formally solved by the exponential operator with an arbitrary initial state.
At this point, we will allow nuclei to move and turn out attention to the equation of motion for the density matrix within the Floquet diabatic basis. The Wigner-transformed density matrix in the Floquet diabatic representation is
(9) |
For a proper F-QCLE, we will need to calculate the time derivative of
(10) |
where the Floquet diabatic states depends on time explicitly. We begin by using Eq. (2) to project into a Floquet diabatic basis. For the commutator term in Eq. (2), we can divide the operator product into matrix multiplication:
(11) | |||||
Here, we have inserted the identity of the diabatic electronic basis () and expanded the time-dependent coefficients in terms of a Fourier series; see Appendix A for more details. Furthermore, if we combine Eq. (11) with the second term on the RHS of Eq. (10), we can write the sum of both terms as , i.e. we can replace with .
For the anti-commutator term, we can use the same procedure to divide the operator product
(12) | |||||
where . Note that, since the Floquet diabatic states do not depend on the nuclear coordinate, we can rewrite the derivative of the electronic Hamiltonian in terms of the Floquet Hamiltonian
(13) |
In the end, we may combine the above expressions to write down a complete diabatic F-QCLE
(14) | |||||
As a final remark, we emphasize that the Floquet Hamiltonian is a time-independent matrix, so Eq. (14) is simply the diabatic QCLE corresponding to an infinitely large electronic Hamiltonian .
II.3 Transformation to the adiabatic representation
To recast the diabatic F-QCLE in an adiabatic representation, we diagonalize the Floquet Hamiltonian matrix by solving the eigenvalue problem:
(15) |
The eigenvalues are the so-called Floquet quasi-energies with corresponding eigenvectors Since is Hermitian, we can choose the eigenvectors to be othornormal so that we have the identities , i.e. and . The Floquet adiabatic state corresponding to the quasi-energy are
(16) |
As a practical matter, although is infinite dimensional, we can truncate highly oscillating Floquet states by replacing with .
With this Floquet adiabatic state basis, the probability density can be obtained by a diabatic-to-adiabatic transformation
(17) |
in the Floquet adiabatic representation. Note that, since the eigenvectors do not depend on time explicitly, the time-derivative of the adiabatic probability density can be calculated simply to be:
(18) |
We are now ready to derive the adiabatic F-QCLE in the following section.
III Different Approaches to derive F-QCLE
In this section, we present several approaches with different orders for the three operations above; as will be shown, different orders will result in different adiabatic F-QCLEs. We summarize these approaches and the corresponding F-QCLEs in Table 1.
III.1 Wigner-Floquet-Adiabatic (WFA) approach
Our first approach follows the order presented above: we first perform partial Wigner transformation, we second project to a Floquet diabatic basis, we third transform to an adiabatic representation. For the last step, following Eq. (18), we transform the diabatic F-QCLE by sandwiching the diabatic F-QCLE (Eq. (14)) with and . The first term on the right hand side of Eq. (14) (the commutator term) becomes
(19) |
For the second term on the right hand side of Eq. (14), since the Floquet adiabatic states depend on the nuclear coordinate , the derivative of the density must yield
(20) |
where the derivative coupling is corresponding to the change of the Floquet adiabatic states with respect to the nuclear coordinate . Note that, if the Floquet Hamiltonian is real-valued, the diagonal element of the derivative coupling is zero (). For the third term on the right hand side of Eq. (14) (the anti-commutator term), the derivative of the Floquet Hamiltonian can be written in terms of the force matrix
(21) |
explicitly,
The force matrix accounts for direct changes in the nuclear momentum associated with the electronic coupling. One can understand the diagonal element as the classical force for nuclear dynamics moving along the -th Floquet quasi-energy surface in the phase space. The off-diagonal force term for corresponds to nuclear velocity rescaling associated with non-adiabatic transitions between Floquet quasi-energy surfaces(Kapral and Ciccotti, 1999; Kapral, 2006).
Finally, the F-QCLE via the WFA approach reads
(22) | |||||
We find that Eq. 22 takes exactly the same form as the standard QCLE in the adiabatic representation (for electron-nuclear dynamics without a driving laser).
III.2 Wigner-Adiabatic-Floquet (WAF) approach
For the second approach, we exchange the “adiabatic” and “Floquet” operations after the partial Wigner transformation. In this case, we directly project Eq. (2) onto the Floquet adiabatic basis . Namely, we make the diabatic-to-adiabatic transform of the Floquet electronic basis prior to the projection onto the dressed electronic states. Thus, we consider this path as the Wigner-Adiabatic-Floquet (WAF) approach.
Overall, we apply a procedure similar to what was used in Eq. 22. For the commutator term, we divide operator products using the same technique as in Appendix A. The derivative term yields a derivative coupling term as the Floquet adiabatic basis depends on explicitly. In the end, the WAF approach includes the first three terms exactly as Eq. (22). However, from the anti-commutator term of Eq. (2), the WAF approach leads to an additional cross derivative force. To see this, we focus on the derivative of the electronic Hamiltonian in the adiabatic representation
(23) |
where we have used the definition of the Floquet Hamiltonian . The derivative of the electronic Hamiltonian yields two terms: first, the same force matrix we obtained in Eq. (21):
(24) |
second, a cross derivative force comes from the explicit dependence of the Floquet adiabatic states on both the nuclear coordinates and time
and explicitly,
(25) |
Note that, unlike the derivative coupling , the diagonal element of is non-zero and real-valued.
Finally, in the same Floquet adiabatic basis as given by Eq. (16), the F-QCLE reads:
(26) | |||||
We observe that, while Eq. (22) and Eq. (26) take the same form, the “effective” force matrix (defined as the coefficients of ) includes an additional cross derivative force that indicates the difference between these two equations. In other words, even though these two QCLEs will propagate electrons equivalently, the difference in the “effective” force matrix will lead to different nuclear dynamics in phase space. Specifically, from Eq. 26, the classical force on the -th quasi-energy surface is given by , implying that the nuclear dynamics should experience the additional cross derivative force on top of the Floquet quasi-energy surface.
III.3 Adiabatic-then-Wigner (AW) approach
Finally, for completeness, we should emphasize that the first derivation of an F-QCLE was published by Horenko, Schmidt, and Schütte who used an approach that was different from either of the approaches above. In Ref. (Horenko, Schmidt, and Schütte, 2001), these authors derived their F-QCLE using an Adiabatic-then-Wigner (AW) approach. Their final equation of motion was:
(27) | |||||
As mentioned above, in the limit of a time-independent Hamiltonian, Izmaylov and co-workers have shown that the AW approach results in the loss of geometric phase related features for a 2D problem(Ryabinkin et al., 2014); the WA approach is far more accurate (see Figs in Ref. (Ryabinkin et al., 2014)). Moreover, in contrast to Eq. (22) and Eq. (26), the last term in Eq. 27 does not include any off-diagonal force term which implies no momentum rescaling. Within a surface hopping context, a lack of momentum rescaling will inevitably lead to a violation of detailed balance(Parandekar and Tully, 2006, 2005; Schmidt, Parandekar, and Tully, 2008). Note that early numerical assessments of the AW approach did not investigate either geometric phase effects or detailed balance(Ando, 2002; Ando and Santer, 2003). For these reasons, we expect the AW approach to the F-QCLE to be less accurate than the WFA and WAF approaches. Nevertheless, to satisfy the readers who is curious about the original AW approach, we will present the relevant transmission and reflection probabilities and final momentum distribution data in Appendix B which should remind one why momentum jumps are so important for scattering calculations111Interestingly, the AW approach to the time-independent QCLE was found to perform well for short times for small closed systems in 2003(Ando, 2002; Ando and Santer, 2003), but we are unaware of more systematic tests of the method beyond Ref. (Ryabinkin et al., 2014), especially within the context of long time dynamics..
IV Results
IV.1 Shifted avoided crossing model
To analyze the difference between these approaches, we consider a shifted avoided crossing model composed of a two-level electronic system coupled to a 1D nuclear motion. The diabatic electronic states are denoted as and and the electronic Hamiltonian is given by
(28) |
where the diabatic energy is given by
(29) |
(30) |
and the diabatic coupling is periodic in time
(31) |
The parameters are , , , , and the nuclear mass is . Note that this model is Tully’s simple avoided crossing model with two modifications: the diabatic coupling becomes time-periodic and the excited potential energy surface is shifted by . For simplicity, we choose the laser frequency large enough () so that the diabatic Floquet states have an avoided crossing only with (at ) and do not have any trivial crossings. The size of the Floquet basis is truncated by as the laser field is weak ().
We assume the initial wavepacket is a Gaussian centered at the initial position and momentum on diabat :
(32) |
where the normalization factor is . The width of the Gaussian is chosen to be . The wavepacket can be propagated exactly in the diabatic representation.
IV.2 F-FSSH based on WFA and WAF
To show the difference between the dynamics as obtained by the different F-QCLEs, we will simulate F-FSSH results for both the WFA and WAF approaches. Within F-FSSH, we describe the propagation of the Floquet wavepacket by a swarm of trajectories, each with its own electronic amplitudes satisfying
(33) |
where , and , representing nuclear trajectory. All nuclear trajectories move classically along an active Floquet state () obeying
(34) |
(35) |
Here, based on the connection between the QCLE and FSSH, the nuclear force in Eq. 35 is determined according to the diagonal element of the effective force matrix in the F-QCLEs.
Consistent with the standard FSSH technique, the hopping probability from active Floquet state to state is given by
where is the classical time step. After each successful hop, the velocity is adjusted to conserve the total Floquet quasi-energy. If a frustrated hop occurs, we implement velocity reversal(Jasper and Truhlar, 2003). Note that we neglect the decoherence correction since the over-coherence problem should not be severe for a simple avoided crossing model(Landry and Subotnik, 2012).
In the end, the probability to measure diabatic state can be evaluated by the density matrix interpretation(Landry, Falk, and Subotnik, 2013)
(36) |
where is the number of the trajectories that have the active surface end up on the Floquet state . Here and are the trajectory indices. We propagate trajectories for an amount of time long enough for each trajectory to pass through the coupling region ( for this parameter set).
IV.3 Effective Floquet quasi-energy surfaces for nuclear dynamics
First, we analyze the effective potential energy surfaces for nuclear dynamics by integrating Eq. 35 over for the WFA and WAF approaches respectively. For the WFA approach, the effective PES simply recovers the Floquet quasi-energy surfaces . For the WAF approach, the effective PES is obtained by where the quasi-energy surface is modified by the integration of the cross derivative force. We find that including the cross derivative force in the WAF approach increases the crossing barrier for the nuclear dynamics on the lower adiabat (see Fig. 1). Note that, in terms of an F-FSSH calculation, these changes will have a direct effect on the nuclear dynamics, but not the electronic amplitudes.

IV.4 Transmission and reflection
Next, we turn our attention to the transmission and reflection probabilities produced by the F-FSSH calculations. Overall, the WFA results are more accurate than the WAF results. In Fig. 2, we find that, according to the WFA approach, there should be a rise in transmission on the lower adiabat around , which is the momentum for which transmission should be allowed classically; see the barrier height () in Fig. 1. Indeed, such a threshold at is found according to exact wavefunction simulation as well. However, for the WAF approach with the cross derivative force, one find a higher crossing barrier energy (), and the transmission on the lower adiabat occurs (incorrectly) at . This result suggests that the cross derivative force is a fictitious term: the WAF semiclassical derivation is spurious.
Let us now focus on the WFA results in more detail. Several points are worth mentioning. First, the transmission to the upper adiabat occurs after , which agrees with the classical energy difference (see Eq. (29) for the definition of ). Second, for high initial momentum (), the F-FSSH-WFA can almost recover the correct nuclear dynamics. Third, in the intermediate momentum region , unfortunately, the F-FSSH wavepacket exhibits less transmission than the exact calculation. This discrepancy may be attributed to FSSH’s inability to capture nuclear tunneling effects.

V Conclusion and Outlook
We have analyzed three approaches for deriving the QCLE within a Floquet formalism, and found three different F-QCLEs. While these F-QCLEs take similar forms, the difference in the effective force matrix can lead to large discrepancies in nuclear dynamics. As such, in the context of driven electron-nuclear dynamics, our results reiterate the fact that one cannot change the order of the operations in the derivation of the correct QCLE. Specifically, as opposed to the WFA approach, the WAF approach [exchanging Floquet electronic basis dressing (F) and adiabatic transformation (A)] is spurious. Overall, our results are very consistent with the results of Izmaylov and Kapral(Ryabinkin et al., 2014) who find that one must be careful when deriving the QCLE even without time dependence; they have already shown that the AW approach does not include any geometric phase information. Thus, in the end, with or without a time-dependent Hamiltonian, it appears that, provided that one moves to the adiabatic representation as the very last step, one will always derive the correct semiclassical equation of motion.
Looking forward, the derivation of the F-QCLE presented here validates the F-FSSH method and paves the way to further improvements in the future. With regard to coherence and decoherence, given the time-independent nature of the Floquet Hamiltonian, we can immediately apply many decoherence schemes, including augmented moment decoherence(Jain, Alguire, and Subotnik, 2016; Landry and Subotnik, 2012; Petit and Subotnik, 2014; Schwartz et al., 1996; Prezhdo and Rossky, 1998), to the F-FSSH algorithm. As far as geometric phase effect is concerned, it is known that Berry phases are already included within the QCLE(Subotnik et al., 2019) for a time-independent electronic Hamiltonian, and so we would expect that similar effects should already be included within this proper F-QCLE for periodic (time-dependent) electronic Hamiltonians. Nevertheless, however, there is one nuance which we have conveniently neglected in the present paper. Note that, according to Eq. (7), we have every reason to believe that the F-QCLE formalism (especially for a non-monocrhomatic driving field with more than one Fourier mode in the time-dependent Hamiltonian) will necessarily introduce a complex (i.e. not real-valued) Floquet Hamiltonian. In such a case, we should find not just Berry phases, but also Berry force(Miao, Bellonzi, and Subotnik, 2019). Future research into the nature of this intrinsic magnetic Berry force—how or if it appears in the context of F-QCLE and surface hopping dynamics—is currently underway and represents an exciting new direction for non-adiabatic theory.
Lastly, it has been recently reported that novel control schemes, such as Floquet engineering(Thanh Phuc and Ishizaki, 2018; Schwennicke and Yuen-Zhou, 2020), can enhance the excitation energy transfer rate even in the presence of strong fluctuations and dissipation. Given so many potential applications for F-QCLE simulations, we believe the present manuscript should find immediate use in the physical chemistry and chemical physics community.
Acknowledgment
This material is based upon work supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Award Number DE-SC0019397. This research also used resources of the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility operated under Contract No. DE-AC02-05CH11231. We thank Abraham Nitzan for very helpful discussions.
Data Availablity
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Appendix A Rewriting the Product of Wignerized Operators As Matrix Multiplication in the Floquet Representation
Throughout this paper, we have constantly used one trick. Namely, we have consistently rewritten the product of two Wigner transformed operators (one of which must be time-periodic) into a non-standard matrix product in the Floquet representation. To see how this trick works in practice, we consider (for example) the operator product .
We insert the electronic identity operator in between and :
(37) |
Next, as in Sec. II, we express in the form of a Fourier series:
(38) |
where the double bracket projection is defined by Fourier transform. With the Fourier series, we write and obtain
(39) |
Appendix B The AW approach

Following Ref. (Horenko, Schmidt, and Schütte, 2001), one can derive the F-QCLE through an AW approach by invoking the wavefunction anstaz in the Floquet adiabatic basis given by Eq. (16), then we insert this anstaz into the TDSE and obtain the equation of motion for in the form of . Here the adiabatic Hamiltonian operator is given by
(40) |
and the nuclear momentum operator is . Thereafter, one writes the equation of motion for the density matrix () as and performs a partial Wigner transformation using the Wigner–Moyal operator:
(41) |
The Wigner transform of the adiabatic Hamiltonian operator reads
(42) |
Next, following the same procedure in Sec. II-A, we expand the Wigner–Moyal operator and truncate the equation of motion to first order of . In the end, we obtain Eq. (27), which is equivalent to Eq. (44) in Ref. (Horenko, Schmidt, and Schütte, 2001) for the case of the monochromatic external field (i.e. using the notaion of Ref. (Horenko, Schmidt, and Schütte, 2001)).
In Fig. 3, we compare the F-FSSH results for the AW v.s. WFA F-QCLEs. Note that, because of the absence of the off-diagonal force terms in Eq. (27), the corresponding AW F-FSSH trajectories neglect velocity rescaling when hopping between Floquet adiabatic surfaces. Thus, an AW trajecory can hop to the upper adiabat (the blue adiabatic potential in Fig. 1) without worrying about energy conservation. Note that, according to Fig. 3, the AW approach does outperform the WAF approach; see Fig. 2. However, the AW approach is far less accurate than the WFA approach. In particular, Fig. 3(a) shows the AW approach over-estimates the transmission probability on the upper adiabat (presumably because of the lack of velocity rescaling). Furthermore, Fig. 3(b) demostrates that the asymptotic velocities are also incorrect. Although this data necessarily relies on an F-FSSH implementation, all together Fig. 3 strongly suggests that the AW approach to the F-QCLE as given in Eq. (27) is not optimal—in agreement with the conclusions of Izmaylov et al.
References
- Regan et al. (1999) P. M. Regan, S. R. Langford, A. J. Orr-Ewing, and M. N. R. Ashfold, The Journal of Chemical Physics 110, 281 (1999).
- Franks, Li, and Kong (1999) K. J. Franks, H. Li, and W. Kong, The Journal of Chemical Physics 110, 11779 (1999).
- Hilsabeck et al. (2019a) K. I. Hilsabeck, J. L. Meiser, M. Sneha, N. Balakrishnan, and R. N. Zare, Physical Chemistry Chemical Physics (2019a), 10.1039/C8CP06107F.
- Hilsabeck et al. (2019b) K. I. Hilsabeck, J. L. Meiser, M. Sneha, J. A. Harrison, and R. N. Zare, J. Am. Chem. Soc. 141, 1067 (2019b).
- Glownia et al. (2016) J. M. Glownia, A. Natan, J. P. Cryan, R. Hartsock, M. Kozina, M. P. Minitti, S. Nelson, J. Robinson, T. Sato, T. van Driel, G. Welch, C. Weninger, D. Zhu, and P. H. Bucksbaum, Physical Review Letters 117 (2016), 10.1103/PhysRevLett.117.153003.
- Glownia et al. (2017) J. M. Glownia, A. Natan, J. P. Cryan, R. Hartsock, M. Kozina, M. P. Minitti, S. Nelson, J. Robinson, T. Sato, T. van Driel, G. Welch, C. Weninger, D. Zhu, and P. H. Bucksbaum, Phys. Rev. Lett. 119, 069302 (2017).
- Lemke et al. (2017) H. T. Lemke, K. S. Kjær, R. Hartsock, T. B. van Driel, M. Chollet, J. M. Glownia, S. Song, D. Zhu, E. Pace, S. F. Matar, M. M. Nielsen, M. Benfatto, K. J. Gaffney, E. Collet, and M. Cammarata, Nature Communications 8, 15342 (2017).
- Fuller et al. (2017) F. D. Fuller, S. Gul, R. Chatterjee, E. S. Burgie, I. D. Young, H. Lebrette, V. Srinivas, A. S. Brewster, T. Michels-Clark, J. A. Clinger, B. Andi, M. Ibrahim, E. Pastor, C. d. Lichtenberg, R. Hussein, C. J. Pollock, M. Zhang, C. A. Stan, T. Kroll, T. Fransson, C. Weninger, M. Kubin, P. Aller, L. Lassalle, P. Bräuer, M. D. Miller, M. Amin, S. Koroidov, C. G. Roessler, M. Allaire, R. G. Sierra, P. T. Docker, J. M. Glownia, S. Nelson, J. E. Koglin, D. Zhu, M. Chollet, S. Song, H. Lemke, M. Liang, D. Sokaras, R. Alonso-Mori, A. Zouni, J. Messinger, U. Bergmann, A. K. Boal, J. M. Bollinger, C. Krebs, M. Högbom, G. N. Phillips, R. D. Vierstra, N. K. Sauter, A. M. Orville, J. Kern, V. K. Yachandra, and J. Yano, Nat Methods 14, 443 (2017), number: 4 Publisher: Nature Publishing Group.
- Bajo, Granucci, and Persico (2014) J. J. Bajo, G. Granucci, and M. Persico, The Journal of Chemical Physics 140, 044113 (2014).
- Hoffmann et al. (2018) N. M. Hoffmann, H. Appel, A. Rubio, and N. T. Maitra, Eur. Phys. J. B 91, 180 (2018).
- Abedi, Maitra, and Gross (2010) A. Abedi, N. T. Maitra, and E. K. U. Gross, Physical Review Letters 105 (2010), 10.1103/PhysRevLett.105.123002.
- Zhou et al. (2020) Z. Zhou, H.-T. Chen, A. Nitzan, and J. E. Subotnik, J. Chem. Theory Comput. 16, 821 (2020).
- Richter et al. (2011) M. Richter, P. Marquetand, J. González-Vázquez, I. Sola, and L. González, Journal of Chemical Theory and Computation 7, 1253 (2011).
- Mitrić, Petersen, and Bonačić-Koutecký (2009) R. Mitrić, J. Petersen, and V. Bonačić-Koutecký, Physical Review A 79 (2009), 10.1103/PhysRevA.79.053416.
- Mai, Marquetand, and González (2018) S. Mai, P. Marquetand, and L. González, Wiley Interdisciplinary Reviews: Computational Molecular Science 8, e1370 (2018).
- Thachuk, Ivanov, and Wardlaw (1996) M. Thachuk, M. Y. Ivanov, and D. M. Wardlaw, The Journal of Chemical Physics 105, 4094 (1996).
- Bajo et al. (2012) J. J. Bajo, J. González-Vázquez, I. R. Sola, J. Santamaria, M. Richter, P. Marquetand, and L. González, The Journal of Physical Chemistry A 116, 2800 (2012).
- Lisinetskaya and Mitrić (2011) P. G. Lisinetskaya and R. Mitrić, Physical Review A 83 (2011), 10.1103/PhysRevA.83.033408.
- Fiedlschuster, Handt, and Schmidt (2016) T. Fiedlschuster, J. Handt, and R. Schmidt, Physical Review A 93, 053409 (2016).
- Fiedlschuster et al. (2017) T. Fiedlschuster, J. Handt, E. K. U. Gross, and R. Schmidt, Physical Review A 95 (2017), 10.1103/PhysRevA.95.063424.
- Horenko, Schmidt, and Schütte (2001) I. Horenko, B. Schmidt, and C. Schütte, The Journal of Chemical Physics 115, 5733 (2001).
- Tully (2012) J. C. Tully, J. Chem. Phys. 137, 22A301 (2012).
- Bittner and Rossky (1995) E. R. Bittner and P. J. Rossky, J. Chem. Phys. 103, 8130 (1995).
- Jasper and Truhlar (2005) A. W. Jasper and D. G. Truhlar, J. Chem. Phys. 123, 64103 (2005).
- Jasper and Truhlar (2003) A. W. Jasper and D. G. Truhlar, Chem. Phys. Lett. 369, 60 (2003).
- Nelson et al. (2013) T. Nelson, S. Fernandez-Alberti, A. E. Roitberg, and S. Tretiak, J. Chem. Phys. 138, 224111 (2013).
- Subotnik et al. (2016) J. E. Subotnik, A. Jain, B. Landry, A. Petit, W. Ouyang, and N. Bellonzi, Annual Review of Physical Chemistry 67, 387 (2016).
- Jain, Alguire, and Subotnik (2016) A. Jain, E. Alguire, and J. E. Subotnik, Journal of Chemical Theory and Computation 12, 5256 (2016).
- Fang and Hammes-Schiffer (1999) J. Y. Fang and S. Hammes-Schiffer, J. Phys. Chem. A 103, 9399 (1999).
- Bedard-Hearn, Larsen, and Schwartz (2005) M. J. Bedard-Hearn, R. E. Larsen, and B. J. Schwartz, J. Chem. Phys. 123, 234106 (2005).
- Subotnik, Ouyang, and Landry (2013) J. E. Subotnik, W. Ouyang, and B. R. Landry, J. Chem. Phys. 139, 214107 (2013).
- Kapral (2016) R. Kapral, Chemical Physics 481, 77 (2016).
- Jasper et al. (2004) A. W. Jasper, C. Zhu, S. Nangia, and D. G. Truhlar, Faraday Discuss. 127, 1 (2004).
- Kapral (2006) R. Kapral, Annu. Rev. Phys. Chem. (2006), 10.1146/annurev.physchem.57.032905.104702.
- Kapral and Ciccotti (1999) R. Kapral and G. Ciccotti, J. Chem. Phys. 110, 8919 (1999), publisher: American Institute of Physics.
- Wong and Rossky (2002a) K. F. Wong and P. J. Rossky, J. Chem. Phys. 116, 8429 (2002a).
- Wong and Rossky (2002b) K. F. Wong and P. J. Rossky, J. Chem. Phys. 116, 8418 (2002b).
- Webster, Rossky, and Friesner (1991) F. Webster, P. Rossky, and R. Friesner, Comput. Phys. Commun. 63, 494 (1991).
- Kelly and Markland (2013) A. Kelly and T. E. Markland, J. Chem. Phys. 139, 014104 (2013).
- Kim et al. (2012) H. W. Kim, A. Kelly, J. W. Park, and Y. M. Rhee, J. Am. Chem. Soc. 134, 11640 (2012), publisher: American Chemical Society.
- Ryabinkin et al. (2014) I. G. Ryabinkin, C.-Y. Hsieh, R. Kapral, and A. F. Izmaylov, The Journal of Chemical Physics 140, 084104 (2014).
- Ando (2002) K. Ando, Chemical Physics Letters 360, 240 (2002).
- Ando and Santer (2003) K. Ando and M. Santer, The Journal of Chemical Physics 118, 10399 (2003).
- Parandekar and Tully (2006) P. V. Parandekar and J. C. Tully, J. Chem. Theory Comput. 2, 229 (2006), detail balance.
- Parandekar and Tully (2005) P. V. Parandekar and J. C. Tully, The Journal of Chemical Physics 122, 094102 (2005).
- Schmidt, Parandekar, and Tully (2008) J. R. Schmidt, P. V. Parandekar, and J. C. Tully, The Journal of Chemical Physics 129, 044104 (2008).
- Note (1) Interestingly, the AW approach to the time-independent QCLE was found to perform well for short times for small closed systems in 2003(Ando, 2002; Ando and Santer, 2003), but we are unaware of more systematic tests of the method beyond Ref. (Ryabinkin et al., 2014), especially within the context of long time dynamics.
- Landry and Subotnik (2012) B. R. Landry and J. E. Subotnik, J. Chem. Phys. 137, 22A513 (2012).
- Landry, Falk, and Subotnik (2013) B. R. Landry, M. J. Falk, and J. E. Subotnik, The Journal of Chemical Physics 139, 211101 (2013).
- Petit and Subotnik (2014) A. S. Petit and J. E. Subotnik, The Journal of Chemical Physics 141, 014107 (2014).
- Schwartz et al. (1996) B. J. Schwartz, E. R. Bittner, O. V. Prezhdo, and P. J. Rossky, J. Chem. Phys. 104, 5942 (1996).
- Prezhdo and Rossky (1998) O. Prezhdo and P. Rossky, Phys. Rev. Lett. 81, 5294 (1998).
- Subotnik et al. (2019) J. Subotnik, G. Miao, N. Bellonzi, H.-H. Teh, and W. Dou, J. Chem. Phys. 151, 074113 (2019).
- Miao, Bellonzi, and Subotnik (2019) G. Miao, N. Bellonzi, and J. Subotnik, J. Chem. Phys. 150, 124101 (2019).
- Thanh Phuc and Ishizaki (2018) N. Thanh Phuc and A. Ishizaki, The Journal of Physical Chemistry Letters 9, 1243 (2018).
- Schwennicke and Yuen-Zhou (2020) K. Schwennicke and J. Yuen-Zhou, The Journal of Physical Chemistry C (2020), 10.1021/acs.jpcc.9b10030.