[]sup
Dynamics with Simultaneous Dissipations to Fermionic and Bosonic Reservoirs
Abstract
We introduce a non-phenomenological framework based on the influence functional method to incorporate simultaneous interactions of particles with fermionic and bosonic thermal reservoirs. In the slow-motion limit, the electronic friction kernel becomes Markovian, enabling an analytical expression for the friction coefficient. The framework is applied to a prototypical electrochemical system, where the metal electrode and solvent act as fermionic and bosonic reservoirs, respectively. We investigate quantum vibrational relaxation of hydrogen on metal surfaces, showing that dissipation to electron-hole pairs reduces the relaxation time. Additionally, in solvated proton discharge, electronic friction prolongs charge transfer by delaying proton transitions between potential wells. This study provides new insights into the interplay of solvent and electronic dissipation effects, with direct relevance to electrochemical processes and other systems involving multiple thermal reservoirs.
I Introduction
Open systems evolve over time interacting with fermionic or bosonic reservoirs representing electrons in electrodes, lattice vibrations or electromagnetic fields. In electrochemical systems, the dynamics occurs under a moderate non-equilibrium condition influenced by a delicate balance of the multiple reservoirs. Particularly important property to characterize the systems is the rate of exchanging the energy and charge with reservoirs, which can be investigated in principle using a microscopic theory-like based on the Langevin equationLangevin (1908); Lemons and Gythiel (1997); Ford and Kac (1987); Ford et al. (1988) by incorporating the fluctuations and dissipation due to environments. The multiple reservoir effect, however, has not received much attention.
Over the years, the Langevin equation has been formally derivedFord et al. (1965); Schmid (1982) and re-derivedCaldeira and Leggett (1983a); Metiu and Schon (1984); Newns (1985); Sebastian (1987); Ford and Kac (1987); Ford et al. (1988) in various contexts, accounting for the dissipation of the impinging particle’s kinetic energy not only into substrate phonon excitations via phononic frictionNewns (1985), but also into the generation of coherent electron-hole (e-h) pairs via electronic frictionBrako and Newns (1980); Schonhammer and Gunnarsson (1980); Brandbyge et al. (1995); Head‐Gordon and Tully (1995); Plihal and Langreth (1998); Dou et al. (2017); Martinazzo and Burghardt (2022). The equation has subsequently been generalized for the formerNewns (1985); Sebastian (1987) and the latterHedegård and Caldeira (1987); Brandbyge et al. (1995), but their simultaneous operation has mostly been neglected, often justified by arguments based on the disparity of particle masses. This is, however, not always the case because the strength of the former is considerably dependent on the environments; it is reduced, for example, when the reaction center moves away from the electrode and correspondingly electrons must transfer farther away therefrom, as typically occurs with increasing pHKumeda et al. (2023); Li et al. (2022). The crossover of the dominant dissipation channels thus occurs in nature.
To address this issue, we reformulate the particle dynamics using the influence functional path integralFeynman and Vernon (2000) framework, which yields without phenomenological assumptions a quasiclassical Langevin equation that incorporates non-Markovian effects of both reservoirs. The dissipation kernel and fluctuations become Markovian in the limit of slow particle motion, allowing analytical evaluation of the electronic friction coefficient. We demonstrate this through numerical studies of (1) quantum vibrational relaxation of hydrogen on metal surfaces and (2) solvated proton discharge dynamics on metal electrodes. Considering the generality of this scheme, it should extend the frontier of the research on quasi-equilibrium systems.
II Theory
The total Hamiltonian of a system comprising a particle coupled to two thermal reservoirs is expressed as , where is the particle Hamiltonian, represents the composite reservoirs () with accounting for inter-reservoir interactions, and describes particle-reservoir coupling. The formalism is initially developed using a bosonic representation, with the case of a fermionic reservoir recovered subsequently. We represent as interacting sets of harmonic oscillators and where is the frequency, and and are sets of modes of and , respectively. We describe the and interaction by the mode-resolved coupling . Eliminating (Sec. LABEL:sup:tot_Ham of Supplementary Materialsup ) results in
(1) |
Here, and are the boson annihilation (creation) operators of and , respectively. The eigenvalue equation, forms the basis of our functional integral scheme. We factorize as , where and are the eigenstates of and , respectively. Denoting the particle coupling with and as and , respectively, with being the particle coordinate, the total Hamiltonian then reads
(2) |
where the second and third terms describe the particle dynamics. At the initial time we assume the reservoirs are in thermal equilibrium and the subsystems are noninteracting. This state is described by
where denotes the particle’s initial density matrix, and is the partition function. Rapid subsystem mixing at justifies neglecting initial correlationsBrandbyge et al. (1995). The total density operator
governs the system’s evolution from to , where and its corresponding adjoint. We express as a reduced density matrix defined as
(3) |
where is thermal average over the complete sets of bosonic states in Eq. (1), , and . and are the forward and backward moving paths in a contour that defines the time evolution of . Expressing these operators in Feynman path integral representationFeynman and Hibbs (2005) gives where the transition amplitude is given by
(4) |
with the particle action . The influence functionals () are expressed (interaction picture) as
(5) |
where, denotes the time (anti-time) ordering operator and
The bosonic trace is worked out in Sec. LABEL:sup:bos_trace of sup . We notice that the product of the influence functionals in Eq. (4) suggests that the reservoirs are essentially noninteractingFeynman and Hibbs (2005). We work with the influence phase (Eq. (LABEL:eq:inf_phase)) instead of and introduce a transformation in terms of the average and relative coordinates. This is formally parallel to the generating functional methodSchmid (1982) where and are the and fields obtained from Keldysh rotations of and residing on the forward and backward branches of the Keldysh contourKamenev (2011). Applying this transformation which has a Jacobian of unity modifies the position variables in Eq. (4) as and , with their time-dependence implied. The potentials contained in are not always conveniently defined in this transformationSchmid (1982); Newns (1985) requiring a functional Taylor expansion in terminated at some order. The expansion gives
(6) |
where we neglected the terms that contain the quantum effects of the wavepacket spreadingNewns (1985). For harmonic , the higher derivatives are identically zero and is undoubtedly quantum. For general , ignoring these terms constitutes the quasiclassical approximationSchmid (1982). We now split the influence phase as , where and whose functional expansions are given in Sec. LABEL:supp:fte of sup . Consequently, the transition amplitude may then be expressed as
(7) |
where the superscript denotes the order of expansion. In Eq. (LABEL:eq:Sigma2) one can deduce that provides the stochastic fluctuations to the dynamics via the random force autocorrelation function defined as
(8) |
Similarly, one can show [Eq. (LABEL:eq:Pi1)] that contains the dissipation through the nonlocal friction coefficient kernel
(9) |
and a potential shift that renormalizes . Except for the presence of a sign function, the form of above closely resembles the friction kernels previously obtained by Head-Gordon and TullyHead‐Gordon and Tully (1995). Expressions for and are retrieved by replacing in Eq. (9) and Eq. (8) their corresponding variables. The transition amplitude now takes the form
(10) | ||||
where . The summations over in Eq. (10) implies the existence of an effective random force that accounts for the stochastic fluctuations coming from both reservoirs. has the properties and . After stochastic averaging of Eq. (10) over , the path integral can be done explicitly (Sec.LABEL:supp:stoav of sup ). This yields a Dirac delta functional (Eq. (LABEL:eq:supp_J)) where we deduce the generalized quasiclassical Langevin equation
(11) |
describing the motion of the particle interacting simultaneously with two thermal reservoirs.
III Interactions with Fermionic and Bosonic Environments
Let us now turn to the case where one of the reservoirs is fermionic. Specifically, we consider a system consisting of a particle interacting with a metal surface immersed in a solvent. The solvent modes and the electronic manifold of the metal represent the bosonic () and fermionic () reservoirs, respectively. The frequency and eigenstates are renormalized by the inter-reservoir interaction, but for simplicity, we assume that the simulation parameters have been already renormalized. We also assume that the particles couple linearly with the bosonic bath, thereby yielding the featureless bosonic friction coefficient in the Caldeira-Leggett frameworkCaldeira and Leggett (1983b, a) (Eq. (LABEL:eq:fric_kerR1_2) of sup ). The dynamics of solvated particles on a metal is typically described by the time-dependent Newns-Anderson-Schmickler (TD-NAS) modelSchmickler (1986); Arguelles and Sugino (2024). In this model, the electronic coupling is treated in the wide-band limit in terms of , with being the time-dependent hopping integrals and is the band energy. The particle level is renormalized as by the solvent reorganization energy ( is the particle charge). We note that the TD-NAS model reduces to the usual time-dependent Anderson-Holstein (AH) HamiltonianAnderson (1961); Holstein (1959) for when is re-interpreted as an average electron-phonon coupling. Using the TD-NAS model, we obtained an effective HamiltonianBrako and Newns (1981, 1989); Arguelles and Sugino (2024) for the metal electron system perturbed by the slowly approaching particle. We then mappedTomonaga (1950); Brako and Newns (1989); Mahan (2000) it to to obtain in Eq. (2). This effective Hamiltonian is given by Eq. (LABEL:eq:Heh1) in sup , where (after bosonization) we identified . Inserting into the version of Eq. (9), yields the electronic frictional force where (Eq. (LABEL:eq:fric_kerR2_2) of sup )
(12) |
is the electronic friction coefficient. Here, is the particle local density of states. Eq. (12) agrees with different forms of that have been derived over the years with varying levels of generalizationsNourtier, A. (1977); Makoshi (1983); Brako and Newns (1981); Schonhammer and Gunnarsson (1980); Head‐Gordon and Tully (1995); Plihal and Langreth (1998); Trail et al. (2003); Mizielinski et al. (2005); Dou et al. (2017). Eq. (11) then reduces to
(13) |
where and is the effective random force with a corresponding autocorrelation function , where
(14) | ||||
III.1 Vibrational Relaxation of H on metal surfaces
We first demonstrate the vibrational relaxation of a hydrogen (H) atom adsorbed on a metal surface using a harmonic potential of the form , as depicted in Figure 1a. We assume that H is located initially at a position away from its equilibrium position ( is the Bohr radius) by . H moves towards the equilibrium position in an oscillatory manner, while dissipating energy, via and , into the excitation of the thermal bath phonons and the creation of coherent e-h pairs in the metal, respectively. Because of smallness of the amplitude of H oscillation, has been set to be a constant . ( fs) is set within the typical relaxation time of e-h excitationsAlducin et al. (2017); Auerbach et al. (2021). The quantum effects are considered by the quantum fluctuations of . For a practical reason, we adopt a Gaussian white noise approximation to the dissipation kernel K(t-s) as introduced recently by Furutani and SalasnichFurutani and Salasnich (2021, 2023), the use of which was validated in the low friction limit for a harmonic potential system. This amounts to rewriting , where . We solved Eq. (13) () using the symplectic BAOAB methodLeimkuhler and Matthews (2013) and obtained the averaged quantities by , where is the number of simulations. In the simulations, we set eV, eV and .




Figure 1b shows that the main effect of the electronic friction is to induce slightly faster vibrational relaxation by enhancing the damping of the quantum dynamics. Figure 1c and Figure 1d which are obtained using the quantum kernel and the classical one , respectively, show the probability distributions on top of the plot of the average. The peaks of density are not normalized and are plotted with between 0 and 1.0 for illustrative purposes. The figures show significant discrepancies existing between the classical and quantum simulations: although the average positions look similar, the probability distribution is much narrower and sharper for the classical case than for the quantum case. For the quantum case, the probability distribution may be interpreted as where is the wavefunction of the H nuclear position. In terms of this, one can recognize that the quantum fluctuations, as introduced by using instead of , affected the spread of wavefunction as a purely quantum effect.
III.2 Proton Discharge on metal electrodes
Let us now consider the dynamics of electrochemical proton discharge which occurs on metal electrodes in the presence of solvents. We study the simplest case corresponding to the Volmer stepLevich (1970) represented as , which involves dissociation of from and subsequent deposition on electrode surface as after accepting an electron. As schematically presented in Figure 2a we use a potential energy surface with double minima ( and ) and a barrier () to simulate a movement of across . The electronic crossing point at which or equivalently, is given by (see Figure 2a caption for details). The movement of H experiences a collective drag due to excitation of the solvent bath phonons () and creation of the e-h pairs in the metali (). To describe the -dependence of , we represent and as Gaussian functionsYoshimori and Makoshi (1986); Nakanishi et al. (1988); sup whose parameters were chosen so that their maxima are located at .




We set eV, where is the electron solvation energy in waterZhan and Dixon (2003). We note that in Eq. (12), diverges in proportion to for , but for our chosen parameters, sharply peaked at (see Figure 2a). We carried out Langevin simulations at K for the duration of ( fs), using the stochastic variable that satisfies , where we set eV (). was placed initially at with a zero initial velocity. The resulting depicted in Figure 2b shows that EF slightly delays the crossing towards due to being significant only within a narrow range of . The distribution of the H position initially peaked at gradually approaches without destroying the peak structure for long . This suggests that recrossing back to side of the PES is highly unlikely, and can be discerned from the extremely assymetric nature of . Using and , we evaluated the the time evolution of the nonadiabatic electron occupation as shown in Figure 2c for cases with and without EF. In slow particle motion limit, can be decomposedArguelles and Sugino (2024) as , where, and are the adiabatic and nonadiabatic components, respectively. In both cases, initially increases exponentially with time and then shows highly oscillatory overpopulation () as manifested by the Langevin velocities. The amplitude of the oscillation is described by the shaded region, of which the width has been plotted in proportion to the variance of the density distribution. The oscillatory overpopulation quickly decays to the equilibrium after crossing the Fermi level mainly due to the nonadiabatic effect, or by . Thereafter, the charge transfer occurs over a short time, as represented by the partial occupation occurring as early as (197 fs). It is worth pointing out that deviates from even at slow speeds, whereas the effect of the electronic friction (EF), is seen to slightly delay the change in the occupation of the energy level. The delay effect can be also seen in Figure 2b wherein the time evolution of is suppressed by an additional drag due to EF. Finally, we show in Figure 2d the average energy dissipation rate and to the bath phonon and e-h pair excitations, respectively, as functions of . The dissipation to the bath phonon excitations is seen to take effect immediately after the particle starts moving and persists over a wide range of . as at . In contrast, is highly localized within the time interval when crosses FL and is peaked. In other words, the energy dissipation towards e-h excitations only occurs within this short range of . The total average dissipated energy eV where the individual contributions are eV and eV. can be thought as the (average) energy gained by the particle from thermal fluctuations generated by , allowing it to overcome the potential barrier of eV at and move to .
IV Discussions and Conclusions
Using the influence function methodFeynman and Vernon (2000), we have derived a non-phenomenological equation of motion for a particle accounting for simultaneous interactions with bosonic and fermionic thermal reservoirs. As an example of the fermionic reservoir, we take electrons in a metal electrode, which dissipate the particle energy into the creation of electron-hole pairs thereby causing electronic friction (EF). An explicit expression for the local dissipation kernel (Markovian kernel) is given in the limit of slow particle motion providing thereby a way to compute the rate of energy transfer via a stochastic simulation. For demonstration purposes, we applied the framework to prototypical electrochemical systems where a hydrogen (H) atom moves in contact with a metal electrode and a solvent and investigated the interplay of the reservoirs using two setups: (1) quantum vibrational relaxation of a hydrogen (H) confined on a metal surface and (2) solvated electrochemical proton discharge. In the former case we have seen that the electronic friction enhances the vibrational friction, and in addition, quantum fluctuation effect manifests as a spread of the wave function of H within the Gaussian white noise approximationFurutani and Salasnich (2023). In the latter case, we have seen that, contrary to the phononic friction, the electronic friction is active at a short time interval of electronic level crossing and contributes to delaying of the electron and proton transfers. On the other hand, the average energy dissipation is comparable between the two reservoirs despite the electronic friction being local near the level crossing. This work marks the first explicit consideration of friction from b oth solvent modes and electronic excitations in the metal during electrochemical proton discharge. Although initially motivated by electrochemistry, the presented framework is be broadly applicable to systems where interactions with multiple thermal baths play a significant role.
V Acknowlegments
This research is supported by the New Energy and Industrial Technology Development Organization (NEDO) project, MEXT as “Program for Promoting Researches on the Supercomputer Fugaku” (Fugaku battery & Fuel Cell Project) (Grant No. JPMXP1020200301, Project No.: hp220177, hp210173, hp200131), Digital Transformation Initiative for Green Energy Materials (DX-GEM) and JSPS Grants-in-Aid for Scientific Research (Young Scientists) No. 19K15397. Some calculations were done using the supercomputing facilities of the Institute for Solid State Physics, The University of Tokyo.
References
- Langevin (1908) P. Langevin, C. R. Acad. Sci. Paris 146, 530 (1908).
- Lemons and Gythiel (1997) D. S. Lemons and A. Gythiel, American Journal of Physics 65, 1079 (1997).
- Ford and Kac (1987) G. W. Ford and M. Kac, Journal of Statistical Physics 46 (1987).
- Ford et al. (1988) G. W. Ford, J. T. Lewis, and R. F. O. Connell, Physical Review A 37, 4419 (1988).
- Ford et al. (1965) G. W. Ford, M. Kac, and P. Mazur, Journal of Mathematical Physics 6, 504 (1965).
- Schmid (1982) A. Schmid, Journal of Low Temperature Physics 5 (1982).
- Caldeira and Leggett (1983a) A. O. Caldeira and A. J. Leggett, ANNALS OF PHYSICS 149, 374 (1983a).
- Metiu and Schon (1984) H. Metiu and G. Schon, Physical Review Letters 53, 13 (1984).
- Newns (1985) D. M. Newns, Surface Science 154, 658 (1985).
- Sebastian (1987) K. L. Sebastian, Proc. Indian Acad. Sci. (Chem. Sci.) 99, 53 (1987).
- Brako and Newns (1980) R. Brako and D. M. Newns, Solid State Communications 33, 713 (1980).
- Schonhammer and Gunnarsson (1980) K. Schonhammer and O. Gunnarsson, Physical Review B 22, 1622 (1980).
- Brandbyge et al. (1995) M. Brandbyge, P. Hedegård, T. F. Heinz, J. A. Misewich, and D. M. Newns, Phys. Rev. B 52, 6042 (1995).
- Head‐Gordon and Tully (1995) M. Head‐Gordon and J. C. Tully, The Journal of Chemical Physics 103, 10137 (1995).
- Plihal and Langreth (1998) M. Plihal and D. C. Langreth, Physical Review B 58, 2191 (1998).
- Dou et al. (2017) W. Dou, G. Miao, and J. E. Subotnik, Phys. Rev. Lett. 119, 046001 (2017).
- Martinazzo and Burghardt (2022) R. Martinazzo and I. Burghardt, Physical Review Letters 128 (2022), 10.1103/PhysRevLett.128.206002.
- Hedegård and Caldeira (1987) P. Hedegård and A. O. Caldeira, Physica Scripta 35, 609 (1987).
- Kumeda et al. (2023) T. Kumeda, L. Laverdure, K. Honkala, M. M. Melander, and K. Sakaushi, Angewandte Chemie International Edition 62, e202312841 (2023).
- Li et al. (2022) P. Li, Y. Jiang, Y. Hu, Y. Men, Y. Liu, W. Cai, and S. Chen, Nature Catalysis 5, 900 (2022).
- Feynman and Vernon (2000) R. P. Feynman and F. L. Vernon, Annals of Physics 281, 547 (2000).
- (22) See Supplemental Material at URL-will-be-inserted-by-publisher.
- Feynman and Hibbs (2005) R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals, emended edition ed., edited by D. F. Styer, Vol. 3 (McGraw-Hill, 2005).
- Kamenev (2011) A. Kamenev, Field theory of nonequilibrium systems (Cambridge University Press, 2011).
- Caldeira and Leggett (1983b) A. Caldeira and A. Leggett, Physica A: Statistical Mechanics and its Applications 121, 587 (1983b).
- Schmickler (1986) W. Schmickler, J. Electroanal. Chem 204, 31 (1986).
- Arguelles and Sugino (2024) E. F. Arguelles and O. Sugino, The Journal of Chemical Physics 160, 144102 (2024).
- Anderson (1961) P. W. Anderson, Phys. Rev. 124, 41 (1961).
- Holstein (1959) T. Holstein, Annals of Physics 8, 325 (1959).
- Brako and Newns (1981) R. Brako and D. M. Newns, J. Phys. C: Solid State Phys 14, 3065 (1981).
- Brako and Newns (1989) R. Brako and D. M. Newns, Rep. Prog. Phys 52, 655 (1989).
- Tomonaga (1950) S. Tomonaga, Progress of Theoretical Physics 5, 544 (1950).
- Mahan (2000) G. D. Mahan, Many-particles physics, 3rd ed., Vol. II (Plenum, 2000).
- Nourtier, A. (1977) Nourtier, A., J. Phys. France 38, 479 (1977).
- Makoshi (1983) K. Makoshi, J. Phys. C: Solid State Phys , 3617 (1983).
- Trail et al. (2003) J. R. Trail, D. M. Bird, M. Persson, and S. Holloway, Journal of Chemical Physics 119, 4539 (2003).
- Mizielinski et al. (2005) M. S. Mizielinski, D. M. Bird, M. Persson, and S. Holloway, Journal of Chemical Physics 122 (2005), 10.1063/1.1854623.
- Alducin et al. (2017) M. Alducin, R. Díez Muiño, and J. Juaristi, Progress in Surface Science 92, 317 (2017).
- Auerbach et al. (2021) D. J. Auerbach, J. C. Tully, and A. M. Wodtke, Natural Sciences 1, e10005 (2021), https://onlinelibrary.wiley.com/doi/pdf/10.1002/ntls.10005 .
- Furutani and Salasnich (2021) K. Furutani and L. Salasnich, Physical Review B 104 (2021), 10.1103/PhysRevB.104.014519.
- Furutani and Salasnich (2023) K. Furutani and L. Salasnich, AAPPS Bulletin 33 (2023), 10.1007/s43673-023-00087-2.
- Leimkuhler and Matthews (2013) B. Leimkuhler and C. Matthews, The Journal of Chemical Physics 138, 174102 (2013).
- Levich (1970) V. Levich, Physical Chemistry. An Advanced Treatise, edited by H. Eyring, D. Henderson, and W. Jost, Vol. Xb (Academic Press, 1970).
- Yoshimori and Makoshi (1986) A. Yoshimori and K. Makoshi, Progress in Surface Science 21, 251 (1986).
- Nakanishi et al. (1988) H. Nakanishi, H. Kasai, and A. Okiji, Surface Science 197, 515 (1988).
- Zhan and Dixon (2003) C. G. Zhan and D. A. Dixon, Journal of Physical Chemistry B 107, 4403 (2003).