Open quantum dynamics of strongly coupled oscillators with multi-configuration time-dependent Hartree propagation and Markovian quantum jumps
Abstract
Modeling the non-equilibrium dissipative dynamics of strongly interacting quantized degrees of freedom is a fundamental problem in several branches of physics and chemistry. We implement a quantum state trajectory scheme for solving Lindblad quantum master equations that describe coherent and dissipative processes for a set of strongly-coupled quantized oscillators. The scheme involves a sequence of stochastic quantum jumps with transition probabilities determined the system state and the system-reservoir dynamics. Between consecutive jumps, the wavefunction is propagated in coordinate space using the multi-configuration time-dependent Hartree (MCTDH) method. We compare this hybrid propagation methodology with exact Liouville space solutions for physical systems of interest in cavity quantum electrodynamics, demonstrating accurate results for experimentally relevant observables using a tractable number of quantum trajectories. We show the potential for solving the dissipative dynamics of finite size arrays of strongly interacting quantized oscillators with high excitation densities, a scenario that is challenging for conventional density matrix propagators due to the large dimensionality of the underlying Hilbert space.
I Introduction
Accurate numerical simulations of open quantum systems are fundamentally important for the development of quantum technology Koch (2016). Understanding and possibly controlling system-reservoir interactions enables a diverse set of applications such as the manipulation of quantum speed limits for driven state evolution Deffner and Lutz (2013), quantum metrology with improved precision bounds Chin et al. (2012); Haase et al. (2016), quantum circuits with improved gate fidelities Di Paolo et al. (2021), quantum optics with nanophotonics Tame et al. (2013); Schmidt et al. (2016), or controlled chemistry with quantum optics Herrera and Spano (2016); Herrera and Owrutsky (2020). In many applications, the temporal correlations of the reservoir variables that couple with the system of interest decay much faster than the system-reservoir interaction times. The open quantum system dynamics can then be modeled with Markovian quantum master equations for the evolution of the reduced system density matrix Breuer et al. (2002); Manzano (2020). For a Hilbert space of dimension , the density matrix scales with , making the direct integration of quantum master equations numerically intractable for large many-body problems, as scales exponentially with the number of particles Weimer et al. (2021).
To simulate the dynamics of open quantum many-body systems, several techniques have been developed, including stochastic methods Gisin and Percival (1992); Dalibard et al. (1992a); Mølmer et al. (1993), tensor networks representations Verstraete et al. (2004); Werner et al. (2016); Orús (2019), phase space methods Carusotto and Ciuti (2005); Navez and Schützhold (2010); Schachenmayer et al. (2015), variational methods Weimer (2015), cluster expansions Cao and Berne (1990); Tanimura (2020); Tang et al. (2013) and field theory techniques Cao and Voth (1994a, b); Sieberer et al. (2016). Broadly speaking, these approaches differ in the way the density matrix and quantum master equation are represented and propagated. Advanced simulation techniques are commonly used in chemical physics for treating strongly molecules in complex reservoirs Valleau et al. (2012); Moix and Cao (2013); del Pino et al. (2018); Wang et al. (2020). Cavity quantum electrodynamics (QED) with molecules has emerged as another domain in which advanced quantum dynamics methods are useful J. et al. (2021); Feist et al. (2018); Simpkins et al. (2021); Herrera and Owrutsky (2020), as the emergence of cavity-induced single-particle and many-body correlations is believed to be relevant for explaining experiments on the modification of chemical reaction rates in optical and infrared cavities Herrera and Spano (2016); Fregoni et al. (2020a); Felicetti et al. (2020); Fregoni et al. (2020b); Campos-Gonzalez-Angulo et al. (2019); Antoniou et al. (2020); Du et al. (2018); del Pino et al. (2018); Ulusoy and Vendrell (2020); Gu and Mukamel (2020); Ahn et al. (2022).
We develop a stochastic wavefunction methodology for solving Markovian quantum master equations in the coordinate representation. The stochastic component of the method is based on the Dalibard-Castin-Molmer quantum jump technique Dalibard et al. (1992b); *Molmer93, a type of Monte Carlo method Plenio and Knight (1998) where the system wavefunction undergoes a sequence of quantum jumps with transition probabilities determined the instantaneous state of the system and the physics of the system-reservoir coupling. Between consecutive quantum jumps the wavefunction evolves deterministically according to the instantaneous Hamiltonian, which we represent in coordinate space using the multi-configuration time-dependent Hartree (MCTDH) method Meyer et al. (1990); *Beck2000; Worth et al. (2007). Observables in the quantum jump technique are guaranteed to converge to the density matrix solution of the quantum master equation by averaging over a sufficient number of wavefunction trajectories *Molmer93. The scheme is applicable to Markovian master equations in Lindblad form Manzano (2020), but extensions to more general reservoirs have been developed de Vega and Alonso (2017); Piilo et al. (2009).
Wavefunction trajectories in the coordinate representation are particularly well suited for studying strongly interacting oscillators subject to driving and dissipation, as often found in molecular cavity QED problems Feist et al. (2018); Herrera and Owrutsky (2020); J. et al. (2021). MCTDH propagators can already capture strong correlations between high-dimensional anharmonic oscillators that naturally emerge in chemical physics Raab and Meyer (2000); Nest and Meyer (2003); Andrianov and Saalfrank (2006); Vendrell and Meyer (2011); Vendrell (2018). Therefore, extending the MCTDH method beyond the use of complex potentials Ulusoy and Vendrell (2020) is a significant step toward scalable atomistic modeling of many-body molecular cavity QED systems.
II Methods
II.1 Monte Carlo Wavefunction Method

For Markovian open quantum systems Breuer et al. (2002); Carmichael (1999); *Carmichael-book2, the evolution for system density matrix in Liouville space is determined by a quantum master equation, which in Lindblad form reads ( is used throughout) Manzano (2020)
(1) |
where is the system Hamiltonian and are Lindblad jump operators that describe the interaction between the system and the -th reservoir channel. The square brackets denote the commutator and curly brackets the anticommutator. The Lindblad form of the master equation is a dynamical semi-group that ensures the positivity of the density matrix Manzano (2020).The Monte Carlo wavefunction technique avoids the direct integration of the quantum master equation by propagating an initial wavefunction over a sequence of non-Hermitian evolution intervals that are interrupted at random times by quantum jumps that encode the physics of the Lindblad operators *Molmer93.
Figure 1 summarizes the proposed Monte Carlo MCTDH (MC-MCTDH) algorithm. Starting from a reference time , the wavefunction is propagated with the MCTDH method up to with the effective non-Hermitian Hamiltonian
(2) |
For small , the wavefunction norm is reduced as
(3) |
where is determined by the instantaneous jump probabilities . At the end of the interval, a pseudo-random number is generated from a uniform distribution, and compared with . If , no quantum jump occurs and the wavefunction is renormalized as
(4) |
before another interval begins. If a quantum jump occurs and a Lindblad jump operator is chosen to act on the wavefunction. The -th reservoir channel is chosen such that the operator gives the smallest jump probability that is greater than . The new wavefunction after the jump becomes
(5) |
and a new interval begins. This algorithm (see Fig. 1) is sequentially repeated until the propagation ends, resulting in a piecewise quantum trajectory for the system wavefunction.
Observables are computed by averaging instantaneous expectation values over multiple trajectories *Molmer93. For the -th quantum trajectory, expectation values are computed at the end of each interval, after normalizing the wavefunction. The procedure is straightforward to extend for computing two-time correlation functions Breuer et al. (2002). By construction, any trajectory-averaged observable
(6) |
asymptotically converges to the density matrix solution with increasing number of trajectories . The convergence proof can be found in Ref. *Molmer93 and is reproduced in Appendix A. In practice, quantum optics problems allow for . Mean-square errors (MSE) of the average observables can be defined by comparing with the exact density matrix solutions (see Appendix A). The independence of each trajectory facilitates computational parallelization strategies.
II.2 MCTDH Propagator
We use the MCTDH method for the deterministic propagation step in Fig. 1. The method was developed by Meyer, Manthe and Cederbaum in 1990 Meyer et al. (1990) as a generalization of the time-dependent Hartree ansatz McLachlan (1964) for solving the time-dependent Schrödinger equation. MCTDH is widely used in chemical physics due to its ability for obtaining essentially exact fully-quantum results for complex molecular systems with a large number of vibrational modes and strong non-adiabatic interactions Raab et al. (1999); Meyer et al. (2009).
The standard ansatz for solving the time-dependent Schrödinger equation is an expansion in a time-independent basis with time-dependent coefficients of the form
(7) |
where are system coordinates, are dynamical expansion coefficients and are the time-independent basis functions that describe the -th degree of freedom. For example, in molecular vibration problems there would be degrees of freedom (e.g., vibrational modes) in this expansion, each described by complete basis of basis functions (e.g., vibrational eigenfunctions) represented on a one-dimensional coordinate grid using discrete variable representation (DVR) techniques Light and Carrington (2000).
MCTDH generalizes the static product basis in Eq. (7) with a linear combination of time-dependent Hartree products of the form
(8) | |||||
where the collective index labels the set of basis functions in a given tensor product configuration that contributes to the wavefunction, the tensor contains the time-dependent amplitudes of each product configuration, and denotes the instantaneous product basis configurations. The number of relevant basis states per configuration and degree of freedom in Eq. (8) is typically smaller than the number of DVR basis functions needed for convergence the static ansatz in Eq. (7). As a result of this dynamical Hilbert space contraction, the number of product configurations needed for convergence is usually smaller than the number of static configurations in the standard method because for each of the -th degrees of freedom, which becomes important when solving high-dimensional quantum dynamics problems.
The time-dependent Schrödinger equation is solved with the MCTDH ansatz [Eq. (8)] and the Dirac-Frenkel variational principle Meyer et al. (2009). Coupled non-linear equations for the and tensors are usually derived by introducing projectors over individual degrees of freedom , are complete in the limit . Projecting Eq. (8) over the -th degree of freedom gives
(9) |
which for the degree-of-freedom, for example, would give an expansion in the complementary space of the form . Variations of the time-dependent coefficients and one-dimensional time-dependent functions are given by
(10) | ||||
(11) | ||||
(12) |
Equations of motion for tensor coefficients are derived using Eqs. (8), (10) and (12) in the Dirac-Frenkel variational principle to get the set of coupled nonlinear equations
(13) | ||||
and
(14) |
with the constraint . In Eq. (14), the tensor elements per degree of freedom are denoted as .
The equations of motion for the functions are also found variationally from the time-dependent Schrödinger equation. In terms of the projection operators they read
(15) | |||
(16) |
where is a reduced density matrix, and the mean-field Hamiltonian of the -th degree of freedom. The solution of the MCTDH equations of motion preserve the norm and total energy for time-independent Hermitian Hamiltonians. In this work we use an extension of the method that includes non-Hermitian (complex) potentials. Additional details about MCTDH can be found in Refs. Meyer et al. (2009); Beck et al. (2000); Meyer (2011).
III Results
We test the MC-MCTDH methodology by solving selected open quantum system problems of interest in cavity QED. For comparison, we also solve the corresponding Lindblad quantum master equation for the density matrix using the open source Python library QuTiP Johansson et al. (2012). The same desktop machine is used for all calculations (CPU 3 GHz Intel Core i5, 8 GB RAM), unless otherwise stated.
III.1 Cavity field with finite photon lifetime

Consider a cavity mode with resonance frequency in a structure with imperfect mirror reflectivity. The mode is modeled as a quantum harmonic oscillator with annihilation operator . Cavity photons leak out to the far field at rate . A minimal decoherence model for radiative decay can be constructed with the Lindblad operator . The effective Hamiltonian for the deterministic steps of the Monte Carlo propagation method is thus given by
(17) |
The Heisenberg equation of motion for the number operator has the simple solution , i.e., exponential decay of initial occupation number with population decay time . In Fig. 2 we show the MC-MCTDH evolution of an initial Fock state with photons, together with the analytical solution. The inset shows that can be achieved with about quantum trajectories.
This one-dimensional example is not exploiting the MCTDH tensor ansatz, but demonstrates the stochastic quantum jumps on a DVR grid for an excited Fock state. In general, most photonic states of interest in quantum optics (Fock state, coherent states, squeezed light) can be accurately represented with DVR grids Triana et al. (2018), which could be advantageous in comparison with more elaborate phase-space representations Zhu et al. (2019); Koessler et al. (2022).
III.2 Vacuum Rabi oscillations
Our next case study is two bilinearly coupled quantum harmonic oscillators in the rotating wave approximation. For an initial state with a single excitation in one of the oscillators, i.e., , we expect the MC-MCTDH algorithm to describe damped Rabi oscillations of the subsystem variables. For a harmonic oscillator with annihilation operator and resonance frequency (e.g., molecular vibration) interacting with the a cavity field of frequency , the effective Hamiltonian is given by
(18) |
where is the coupling strength (Rabi frequency). Dissipation of the -oscillator at rate is modeled with the Lindblad operator , and cavity dissipation is described as before.
Figure 3 shows the evolution of the occupation number obtained with MC-MCTDH for strong resonant coupling ( and ). The exact density matrix solution is also shown for comparison. For the chosen system parameters, averaging over quantum trajectories gives good short-time accuracy, although errors tend to accumulate at longer times (). Increasing the number of trajectories () gives results that match better the exact Liouville-space solution.

III.3 Jaynes-Cummings revivals in driven cavities

We now study the coupling of a two-level atom (qubit) with a cavity field prepared in a coherent state with (real) amplitude . In the number (Fock) basis, the coherent state gives a Poissonian distribution with Carmichael (1999). Unitary dynamics is governed by the Jaynes-Cummings Hamiltonian Jaynes and Cummings (1963); Gerry and Knight (2005)
(19) |
and are the Pauli spin- operators, is the qubit frequency, and the Rabi frequency. Since spins do not have a coordinate dependence, we represent the two spin projections in a DVR grid using two flat potentials separated in energy by , i.e., . This is equivalent to having two electronic states in the MCTDH package Worth et al. (2007). Pauli operators can be constructed accordingly. Atomic relaxation at the rate is given by the Lindblad operator , and cavity decay is described before. The effective Hamiltonian for deterministic propagation is given by , with in Eq. (19).
Figure 4 shows the evolution of the inversion , where denotes level population. The qubit is initialized in the excited level () and strongly couples to a cavity that has initially photons on average. The Rabi frequency is . For the small damping rates used (), the MC-MCTDH reproduces the long-time revivals of the population inversion expected due to the exchange of coherence between qubit and Fock sub-levels that compose the coherent state Gerry and Knight (2005), using only quantum trajectories. Deviations from the exact Liouville-space solution are negligible at short times, but grow over longer timescales. The inset in Fig. 4 shows the drop of the MSE with increasing number of trajectories.
III.4 Independent quantum oscillators coupled to a common cavity field
In this example, we consider a set of independent oscillators that couple with a common quantized cavity field . The non-Hermitian effective Hamiltonian of the system is given by
(20) | ||||
with jump operators for the -oscillators , and cavity dissipation described as before. We focus on the coherent population transfer between oscillators beyond the single-excitation manifold.


Figure 5a shows the MC-MCTDH evolution of the occupation numbers and for a set of oscillators in a cavity that initially has one excitation in and another excitation in , with the cavity field in the vacuum, i.e., . The results converge to the exact Liouville-space solution for the -variables for quantum trajectories. However, Fig. 5b shows that long-time errors of about tend to accumulate for the photon number after several population transfer cycles, which could be reduced by increasing . The inset of 5b shows that the method captures the short-time rise and long-time decay of the Fock state population . Since the amount of ground state bleaching is significant (), the Hilbert space dimension needed to converge the Liouville-space solution is higher than the previous examples. For the parameters in Fig. 5, QuTiP solutions converge with a minimum Hilbert space dimension of , where is the maximum quantum number used for each of the -oscillators and is the highest Fock state included.
III.5 Strongly interacting array of quantum oscillators coupled to a common cavity field
As a final example, we compute the dynamics of a circular array of quantum harmonic oscillators of size with periodic boundary conditions. The array oscillator couple strongly to a common cavity field. We monitor the population transfer dynamics between oscillators in the array with the MC-MCTDH method assuming strong bilinear coupling between sites. The effective Hamiltonian is given by
where is the nearest-neighbor coupling and the other variables are defined as before.

In Fig. 6 we show the occupation numbers of the cavity field and the oscillator , for an array of size and an initially excited cavity with photons. Two array excitation levels are studied. In Fig. 6a the oscillator array is set to the ground state (). In this case, the initial cavity excitations are transferred rapidly to the oscillator array creating a many-particle wavepacket that eventually decays within a few vibrational lifetimes. The evolution can be converged in Liouville space with a truncated Hilbert space that includes up to excitations per site and photons, giving the Hilbert space dimension . Converged MC-MCTDH calculations involved grid points for each degree of freedom in a harmonic oscillator-DVR primitive basis, with time-dependent functions, giving 1844 equations of motion to solve. Fig. 6a shows that the MC-MCTDH expectation values agree with the converged Liouville-space results within % with only quantum trajectories.
In Fig. 6b, we informally probe the efficiency of the MC-MCTDH method by increasing the initial excitation density of the array to two excitations: one excitation in oscillator and another excitation in oscillator , again with two initial cavity photons, i.e., . For the same Hamiltonian and dissipative parameters in Fig. 6a, convergence of the Liouville space solution was not possible on the same machine where MC-MCTDH was implemented, due to RAM constraints. We obtained converged density matrix solutions with QuTiP implemented in a high-perfomance computing (HCP) workstation. The minimum Hilbert space dimension needed for convergence was found to be , which included excitations per site and cavity Fock states. Fig. 6b shows that the MC-MCTDH solution obtained in the low-RAM machine agrees well with the numerically-exact Liouville space solutions for and oscillators in the HCP workstation, using only 300 quantum trajectories.
IV Conclusions and Discussion
Motivated by current problems in molecular quantum electrodynamics Feist et al. (2018); Herrera and Owrutsky (2020); J. et al. (2021), we developed an efficient numerical methodology for computing the open system dynamics of strongly coupled quantized oscillators. The method combines deterministic non-unitary propagation of the many-particle system wavefunction in coordinate space, with a sequence of stochastic quantum jumps that model the interaction of the system with multiple reservoirs. The stochastic component of the propagator is based on the Monte-Carlo wavefunction method developed in quantum optics Dalibard et al. (1992b), which by construction converges to Lindblad semi-group dynamics *Molmer93. The deterministic steps are implemented using the multi-configuration time-dependent Hartree method (MCTDH Meyer et al. (1990)), which was originally developed to describe wavepackets with continuous-variable degrees of freedom that are relevant in chemical dynamics.
We demonstrate the applicability of the method by solving the open quantum system dynamics of selected scenarios of current interest: (i) decay dynamics of a lossy optical cavity; (ii) vacuum Rabi oscillations for strongly interacting cavity-vibration systems with photonic and material losses; (iii) population revivals for a two-level system in a driven cavity; (iv) photon-mediated population transfer between independent molecular vibrations coupled to a common cavity field; (v) quench dynamics in an array of strongly interacting vibrational oscillators with high initial excitation density. In all cases the proposed method converges to the exact Liouville-space solution with a reasonably low number of quantum trajectories. For an array of strongly coupled oscillators with high excitation density, preliminary tests suggest that the method is more efficient than currently available open-source quantum optics libraries Johansson et al. (2012) at equal machine resources.
Applications of this quantum dynamics methodology include the study of vibrational relaxation and rotational depolarization of molecular ensembles in liquid-phase infrared cavities under vibrational strong coupling Simpkins et al. (2021), which are believed to determine the dynamics of unconventinoal light-matter coherences that emerge in two-dimensional infrared cavity spectroscopy Grafton et al. (2021), and the reactive dynamics of polar molecules under vibrational ultrastrong coupling Hernández and Herrera (2019); Triana et al. (2020). The methodology can also be implemented with time-dependent Hamiltonians to study coherent control scenarios in nanophotonics Muller et al. (2018); Metzger et al. (2019); Triana et al. (2022). Future extensions of the method can be implemented to describe systems with non-Markovian coupling to multiple reservoirs Piilo et al. (2009).
Acknowledgements.
We thank Johannes Schachenmayer and Oriol Vendrell for comments. This work was supported by ANID Postdoctoral 3200565, FONDECYT Regular 1181743, Millennium Science Initiative Program ICN17-012 and Programa de Cooperación Científica ECOS-ANID ECOS200028.Data Availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Appendix A Equivalence of the Monte Carlo Wavefunction method and Lindblad quantum master equations
The time evolution of wave function inside MCWF is performed by finding the wave function at a time for enough small . At first-order approximation we obtain (in atomic units)
(22) |
like is non-Hermitian, is not normalized and hence
(23) |
with
(24) |
where describes the loss of the norm of jump operator .
Considering the operator , for a define number of realizations with different random numbers at time , the average value of is given by
(25) | ||||
with . Inserting Eqs. (4) and (22) into Eq. (25), we obtain
(26) | ||||
(27) | ||||
and considering that is given by Eq. (2), we obtain
(28) | ||||
Now, if we apply the limit , Eq. (28) reduces to
(29) |
where is the Lindblad superoperator given by
(30) |
Note that Eq. (29) is equivalent to Eq. (1). Hence, we demonstrate the validity of MCWF with the master equation in Lindblad form.
Now, the next step is to calculate the expectation value of a given operator , which according to the density operator in the limits and is equivalent to . In the MCWF method is calculated by implementing Eq. (6). However, in MCWF there are numerical errors for a finite number of trajectories . We measure the error by calculating the mean squared error at time given by
(31) |
where is the expectation value of trajectory and is the exact solution.
References
- Koch (2016) C. P. Koch, Journal of Physics: Condensed Matter 28, 213001 (2016).
- Deffner and Lutz (2013) S. Deffner and E. Lutz, Phys. Rev. Lett. 111, 010402 (2013).
- Chin et al. (2012) A. W. Chin, S. F. Huelga, and M. B. Plenio, Phys. Rev. Lett. 109, 233601 (2012).
- Haase et al. (2016) J. F. Haase, A. Smirne, S. F. Huelga, J. Kołodynski, and R. Demkowicz-Dobrzanski, Quantum Measurements and Quantum Metrology, 5, 13 (2016).
- Di Paolo et al. (2021) A. Di Paolo, T. E. Baker, A. Foley, D. Sénéchal, and A. Blais, npj Quantum Information 7, 11 (2021).
- Tame et al. (2013) M. S. Tame, K. R. McEnery, Ş. K. Özdemir, J. Lee, S. A. Maier, and M. S. Kim, Nature Physics 9, 329 (2013).
- Schmidt et al. (2016) M. K. Schmidt, R. Esteban, A. González-Tudela, G. Giedke, and J. Aizpurua, ACS Nano, ACS Nano 10, 6291 (2016).
- Herrera and Spano (2016) F. Herrera and F. C. Spano, Phys. Rev. Lett. 116, 238301 (2016).
- Herrera and Owrutsky (2020) F. Herrera and J. Owrutsky, The Journal of Chemical Physics 152, 100902 (2020), https://doi.org/10.1063/1.5136320 .
- Breuer et al. (2002) H. Breuer, P. Breuer, F. Petruccione, and S. Petruccione, The Theory of Open Quantum Systems (Oxford University Press, 2002).
- Manzano (2020) D. Manzano, AIP Advances 10, 025106 (2020).
- Weimer et al. (2021) H. Weimer, A. Kshetrimayum, and R. Orús, Rev. Mod. Phys. 93, 015008 (2021).
- Gisin and Percival (1992) N. Gisin and I. C. Percival, Journal of Physics A: Mathematical and General 25, 5677 (1992).
- Dalibard et al. (1992a) J. Dalibard, Y. Castin, and K. Mølmer, Phys. Rev. Lett. 68, 580 (1992a).
- Mølmer et al. (1993) K. Mølmer, Y. Castin, and J. Dalibard, J. Opt. Soc. Am. B 10, 524 (1993).
- Verstraete et al. (2004) F. Verstraete, J. J. García-Ripoll, and J. I. Cirac, Phys. Rev. Lett. 93, 207204 (2004).
- Werner et al. (2016) A. H. Werner, D. Jaschke, P. Silvi, M. Kliesch, T. Calarco, J. Eisert, and S. Montangero, Phys. Rev. Lett. 116, 237201 (2016).
- Orús (2019) R. Orús, Nature Reviews Physics 1, 538 (2019).
- Carusotto and Ciuti (2005) I. Carusotto and C. Ciuti, Phys. Rev. B 72, 125335 (2005).
- Navez and Schützhold (2010) P. Navez and R. Schützhold, Phys. Rev. A 82, 063603 (2010).
- Schachenmayer et al. (2015) J. Schachenmayer, A. Pikovski, and A. M. Rey, Phys. Rev. X 5, 011022 (2015).
- Weimer (2015) H. Weimer, Phys. Rev. Lett. 114, 040402 (2015).
- Cao and Berne (1990) J. Cao and B. J. Berne, The Journal of Chemical Physics 92, 7531 (1990), https://doi.org/10.1063/1.458189 .
- Tanimura (2020) Y. Tanimura, The Journal of Chemical Physics 153, 020901 (2020), https://doi.org/10.1063/5.0011599 .
- Tang et al. (2013) B. Tang, E. Khatami, and M. Rigol, Computer Physics Communications 184, 557 (2013).
- Cao and Voth (1994a) J. Cao and G. A. Voth, The Journal of Chemical Physics 100, 5093 (1994a), https://doi.org/10.1063/1.467175 .
- Cao and Voth (1994b) J. Cao and G. A. Voth, The Journal of Chemical Physics 100, 5106 (1994b), https://doi.org/10.1063/1.467176 .
- Sieberer et al. (2016) L. M. Sieberer, M. Buchhold, and S. Diehl, Reports on Progress in Physics 79, 096001 (2016).
- Valleau et al. (2012) S. Valleau, S. K. Saikin, M.-H. Yung, and A. A. Guzik, The Journal of Chemical Physics 137, 034109 (2012), https://doi.org/10.1063/1.4732122 .
- Moix and Cao (2013) J. M. Moix and J. Cao, The Journal of Chemical Physics 139, 134106 (2013), https://doi.org/10.1063/1.4822043 .
- del Pino et al. (2018) J. del Pino, F. A. Y. N. Schröder, A. W. Chin, J. Feist, and F. J. Garcia-Vidal, Phys. Rev. Lett. 121, 227401 (2018).
- Wang et al. (2020) Y.-S. Wang, P. Nijjar, X. Zhou, D. I. Bondar, and O. V. Prezhdo, The Journal of Physical Chemistry B 124, 4326 (2020).
- J. et al. (2021) G.-V. F. J., C. Cristiano, and E. T. W., Science 373, eabd0336 (2021).
- Feist et al. (2018) J. Feist, J. Galego, and F. J. Garcia-Vidal, ACS Photonics 5, 205 (2018).
- Simpkins et al. (2021) B. S. Simpkins, A. D. Dunkelberger, and J. C. Owrutsky, The Journal of Physical Chemistry C 125, 19081 (2021).
- Fregoni et al. (2020a) J. Fregoni, G. Granucci, M. Persico, and S. Corni, Chem 6, 250 (2020a).
- Felicetti et al. (2020) S. Felicetti, J. Fregoni, T. Schnappinger, S. Reiter, R. de Vivie-Riedle, and J. Feist, The Journal of Physical Chemistry Letters 11, 8810 (2020).
- Fregoni et al. (2020b) J. Fregoni, S. Corni, M. Persico, and G. Granucci, Journal of Computational Chemistry 41, 2033 (2020b).
- Campos-Gonzalez-Angulo et al. (2019) J. A. Campos-Gonzalez-Angulo, R. F. Ribeiro, and J. Yuen-Zhou, Nature Communications 10, 4685 (2019).
- Antoniou et al. (2020) P. Antoniou, F. Suchanek, J. F. Varner, and J. J. Foley, The Journal of Physical Chemistry Letters 11, 9063 (2020).
- Du et al. (2018) M. Du, L. A. Martínez-Martínez, R. F. Ribeiro, Z. Hu, V. M. Menon, and J. Yuen-Zhou, Chem. Sci. 9, 6659 (2018).
- Ulusoy and Vendrell (2020) I. S. Ulusoy and O. Vendrell, The Journal of Chemical Physics 153, 044108 (2020), https://doi.org/10.1063/5.0011556 .
- Gu and Mukamel (2020) B. Gu and S. Mukamel, Chem. Sci. 11, 1290 (2020).
- Ahn et al. (2022) W. Ahn, F. Herrera, and B. Simpkins, ChemRxiv (2022), 10.26434/chemrxiv-2022-wb6vs.
- Dalibard et al. (1992b) J. Dalibard, Y. Castin, and K. Mølmer, Phys. Rev. Lett. 68, 580 (1992b).
- Plenio and Knight (1998) M. B. Plenio and P. L. Knight, Rev. Mod. Phys. 70, 101 (1998).
- Meyer et al. (1990) H.-D. Meyer, U. Manthe, and L. Cederbaum, Chemical Physics Letters 165, 73 (1990).
- Beck et al. (2000) M. Beck, A. Jackle, G. Worth, and H.-D. Meyer, Physics Reports 324, 1 (2000).
- Worth et al. (2007) G. Worth, M. Beck, A. Jäckle, and H. Meyer, “The MCTDH Package, version 8.4,” (2007), http://mctdh.uni-hd.de.
- de Vega and Alonso (2017) I. de Vega and D. Alonso, Rev. Mod. Phys. 89, 015001 (2017).
- Piilo et al. (2009) J. Piilo, K. Härkönen, S. Maniscalco, and K.-A. Suominen, Phys. Rev. A 79, 062112 (2009).
- Raab and Meyer (2000) A. Raab and H.-D. Meyer, Theoretical Chemistry Accounts 104, 358 (2000).
- Nest and Meyer (2003) M. Nest and H.-D. Meyer, The Journal of Chemical Physics 119, 24 (2003).
- Andrianov and Saalfrank (2006) I. Andrianov and P. Saalfrank, The Journal of Chemical Physics 124, 034710 (2006).
- Vendrell and Meyer (2011) O. Vendrell and H.-D. Meyer, The Journal of Chemical Physics 134, 044135 (2011), https://doi.org/10.1063/1.3535541 .
- Vendrell (2018) O. Vendrell, Chemical Physics 509, 55 (2018), high-dimensional quantum dynamics (on the occasion of the 70th birthday of Hans-Dieter Meyer).
- Carmichael (1999) H. Carmichael, Statistical Mehods in Quantum Optics 1: Master Equations and Fokker-Planck Equations (Springer Berlin / Heidelberg, 1999).
- Carmichael (2008) H. Carmichael, Statistical Mehods in Quantum Optics 2: Non-Classical Fields (Springer Berlin / Heidelberg, 2008).
- McLachlan (1964) A. McLachlan, Molecular Physics 8, 39 (1964).
- Raab et al. (1999) A. Raab, G. A. Worth, H.-D. Meyer, and L. S. Cederbaum, The Journal of Chemical Physics 110, 936 (1999).
- Meyer et al. (2009) H. Meyer, F. Gatti, and G. Worth, Multidimensional Quantum Dynamics: MCTDH Theory and Applications (John Wiley & Sons, 2009).
- Light and Carrington (2000) J. Light and T. Carrington, Advances in Chemical Physics 114, 263 (2000).
- Meyer (2011) H. Meyer, “Introduction to MCTDH: Lecture notes,” (2011), https://www.pci.uni-heidelberg.de/cms/mctdh.html.
- Johansson et al. (2012) J. Johansson, P. Nation, and F. Nori, Computer Physics Communications 183, 1760 (2012).
- Triana et al. (2018) J. Triana, D. Peláez, and J. Sanz-Vicario, Journal of Physical Chemistry A 122, 2266 (2018).
- Zhu et al. (2019) B. Zhu, A. M. Rey, and J. Schachenmayer, New Journal of Physics 21, 082001 (2019).
- Koessler et al. (2022) E. R. Koessler, A. Mandal, and P. Huo, The Journal of Chemical Physics (2022), 10.1063/5.0099922, https://doi.org/10.1063/5.0099922 .
- Jaynes and Cummings (1963) E. Jaynes and F. Cummings, Proceedings of the IEEE 51, 89 (1963).
- Gerry and Knight (2005) C. Gerry and P. Knight, Introductory quantum optics (Cambridge University Press, Cambridge, 2005).
- Grafton et al. (2021) A. B. Grafton, A. D. Dunkelberger, B. S. Simpkins, J. F. Triana, F. J. Hernández, F. Herrera, and J. C. Owrutsky, Nature Communications 12, 214 (2021).
- Hernández and Herrera (2019) F. J. Hernández and F. Herrera, The Journal of Chemical Physics 151, 144116 (2019), https://doi.org/10.1063/1.5121426 .
- Triana et al. (2020) J. Triana, F. Hernández, and F. Herrera, The Journal of Chemical Physics 152, 234111 (2020).
- Muller et al. (2018) E. A. Muller, B. Pollard, H. A. Bechtel, R. Adato, D. Etezadi, H. Altug, and M. B. Raschke, ACS Photonics, ACS Photonics 5, 3594 (2018).
- Metzger et al. (2019) B. Metzger, E. Muller, J. Nishida, B. Pollard, M. Hentschel, and M. B. Raschke, Phys. Rev. Lett. 123, 153001 (2019).
- Triana et al. (2022) J. F. Triana, M. Arias, J. Nishida, E. A. Muller, R. Wilcken, S. C. Johnson, A. Delgado, M. B. Raschke, and F. Herrera, The Journal of Chemical Physics 156, 124110 (2022), https://doi.org/10.1063/5.0075894 .