semiclassical monte carlo: a first principles approach to ...serg/postscript/jcp.4900988.pdf ·...

12
THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first principles approach to non-adiabatic molecular dynamics Alexander J. White, 1,2 Vyacheslav N. Gorshkov, 1,3 Ruixi Wang, 1,4 Sergei Tretiak, 1,2,5, a) and Dmitry Mozyrsky 1, b) 1 Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA 2 Center for Nonlinear Studies (CNLS), Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA 3 National Technical University of Ukraine, Kiev 03056, Ukraine 4 Department of Chemistry, Wayne State University, 5101 Cass Avenue, Detroit, Michigan 48202, USA 5 Center for Integrated Nanotechnologies (CINT), Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA (Received 10 September 2014; accepted 23 October 2014; published online 10 November 2014) Modeling the dynamics of photophysical and (photo)chemical reactions in extended molecular sys- tems is a new frontier for quantum chemistry. Many dynamical phenomena, such as intersystem crossing, non-radiative relaxation, and charge and energy transfer, require a non-adiabatic descrip- tion which incorporate transitions between electronic states. Additionally, these dynamics are often highly sensitive to quantum coherences and interference effects. Several methods exist to simulate non-adiabatic dynamics; however, they are typically either too expensive to be applied to large molec- ular systems (10’s-100’s of atoms), or they are based on ad hoc schemes which may include severe approximations due to inconsistencies in classical and quantum mechanics. We present, in detail, an algorithm based on Monte Carlo sampling of the semiclassical time-dependent wavefunction that involves running simple surface hopping dynamics, followed by a post-processing step which adds little cost. The method requires only a few quantities from quantum chemistry calculations, can sys- tematically be improved, and provides excellent agreement with exact quantum mechanical results. Here we show excellent agreement with exact solutions for scattering results of standard test prob- lems. Additionally, we find that convergence of the wavefunction is controlled by complex valued phase factors, the size of the non-adiabatic coupling region, and the choice of sampling function. These results help in determining the range of applicability of the method, and provide a starting point for further improvement. © 2014 AIP Publishing LLC.[http://dx.doi.org/10.1063/1.4900988] I. INTRODUCTION Advances in theoretical methods as well as in computa- tional power have positioned quantum chemistry as a pow- erful tool in studying various properties of molecules and molecular complexes. 1, 2 Accurate, but numerically expen- sive, wavefunction based approaches enable detailed descrip- tion of electronic properties in relatively small molecules. In contrast, efficient Density functional theory (DFT) methods in combination with molecular dynamics (MD) allow one to deduce valuable information on ground state properties for molecules with hundreds of atoms in size. 3 Further advances in time-dependent density functional theory (TDDFT) have made it possible to carry out efficient calculations for exited states in these systems and thus predict their susceptibilities, absorption, and emission spectra, etc. 4 At present quantum chemistry faces a new frontier: Now it aims not only at computations of the equilibrium (static) molecular properties, but also at modeling dynamics of pho- tophysical processes and (photo)chemical reactions in molec- ular systems. 5, 6 The latter are characterized by different re- action pathways that lead to different reaction products. For a) [email protected] b) [email protected] example, upon absorbing a photon, a molecule may undergo radiative relaxation processes, 7 such as fluorescence, 8 and phosphorescence; non-radiative relaxation processes, such as photodissociation 911 or photoisomerization; 1215 charge; 1619 and energy transfer; 2022 or intersystem crossing. 2325 Such dynamics are obviously accompanied by electronic transi- tions, involving complex electron-phonon interaction 26, 27 and possibly many electronic states, and therefore cannot be de- scribed within the standard adiabatic or Born-Oppenheimer approximation. That is, while in the traditional adiabatic ap- proximation the nuclei move along a given potential energy surface (PES) of a molecule, description of photophysical or photochemical processes requires accounting for the transi- tions between different PESs. Such transitions typically occur in the relatively small regions of the phase space where the PESs closely approach or cross each other, so that the energy separation between relevant PESs becomes comparable with inverse time scales of the nuclear motion (in units of ¯), i.e., phonon frequencies. Thus the assumption of separation be- tween electronic and nuclear timescales breaks down in the vicinity of electronic PES crossings and dynamics becomes non-adiabatic. Several approaches exist in the literature to treat the non-adiabatic dynamics within the framework of quantum chemistry. 2836 The most well known of these are mixed 0021-9606/2014/141(18)/184101/12/$30.00 © 2014 AIP Publishing LLC 141, 184101-1 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 204.121.6.216 On: Tue, 26 May 2015 05:32:56

Upload: others

Post on 28-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014)

Semiclassical Monte Carlo: A first principles approach to non-adiabaticmolecular dynamics

Alexander J. White,1,2 Vyacheslav N. Gorshkov,1,3 Ruixi Wang,1,4 Sergei Tretiak,1,2,5,a)

and Dmitry Mozyrsky1,b)

1Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA2Center for Nonlinear Studies (CNLS), Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA3National Technical University of Ukraine, Kiev 03056, Ukraine4Department of Chemistry, Wayne State University, 5101 Cass Avenue, Detroit, Michigan 48202, USA5Center for Integrated Nanotechnologies (CINT), Los Alamos National Laboratory, Los Alamos,New Mexico 87545, USA

(Received 10 September 2014; accepted 23 October 2014; published online 10 November 2014)

Modeling the dynamics of photophysical and (photo)chemical reactions in extended molecular sys-tems is a new frontier for quantum chemistry. Many dynamical phenomena, such as intersystemcrossing, non-radiative relaxation, and charge and energy transfer, require a non-adiabatic descrip-tion which incorporate transitions between electronic states. Additionally, these dynamics are oftenhighly sensitive to quantum coherences and interference effects. Several methods exist to simulatenon-adiabatic dynamics; however, they are typically either too expensive to be applied to large molec-ular systems (10’s-100’s of atoms), or they are based on ad hoc schemes which may include severeapproximations due to inconsistencies in classical and quantum mechanics. We present, in detail,an algorithm based on Monte Carlo sampling of the semiclassical time-dependent wavefunction thatinvolves running simple surface hopping dynamics, followed by a post-processing step which addslittle cost. The method requires only a few quantities from quantum chemistry calculations, can sys-tematically be improved, and provides excellent agreement with exact quantum mechanical results.Here we show excellent agreement with exact solutions for scattering results of standard test prob-lems. Additionally, we find that convergence of the wavefunction is controlled by complex valuedphase factors, the size of the non-adiabatic coupling region, and the choice of sampling function.These results help in determining the range of applicability of the method, and provide a startingpoint for further improvement. © 2014 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4900988]

I. INTRODUCTION

Advances in theoretical methods as well as in computa-tional power have positioned quantum chemistry as a pow-erful tool in studying various properties of molecules andmolecular complexes.1, 2 Accurate, but numerically expen-sive, wavefunction based approaches enable detailed descrip-tion of electronic properties in relatively small molecules. Incontrast, efficient Density functional theory (DFT) methodsin combination with molecular dynamics (MD) allow one todeduce valuable information on ground state properties formolecules with hundreds of atoms in size.3 Further advancesin time-dependent density functional theory (TDDFT) havemade it possible to carry out efficient calculations for exitedstates in these systems and thus predict their susceptibilities,absorption, and emission spectra, etc.4

At present quantum chemistry faces a new frontier: Nowit aims not only at computations of the equilibrium (static)molecular properties, but also at modeling dynamics of pho-tophysical processes and (photo)chemical reactions in molec-ular systems.5, 6 The latter are characterized by different re-action pathways that lead to different reaction products. For

a)[email protected])[email protected]

example, upon absorbing a photon, a molecule may undergoradiative relaxation processes,7 such as fluorescence,8 andphosphorescence; non-radiative relaxation processes, such asphotodissociation9–11 or photoisomerization;12–15 charge;16–19

and energy transfer;20–22 or intersystem crossing.23–25 Suchdynamics are obviously accompanied by electronic transi-tions, involving complex electron-phonon interaction26, 27 andpossibly many electronic states, and therefore cannot be de-scribed within the standard adiabatic or Born-Oppenheimerapproximation. That is, while in the traditional adiabatic ap-proximation the nuclei move along a given potential energysurface (PES) of a molecule, description of photophysical orphotochemical processes requires accounting for the transi-tions between different PESs. Such transitions typically occurin the relatively small regions of the phase space where thePESs closely approach or cross each other, so that the energyseparation between relevant PESs becomes comparable withinverse time scales of the nuclear motion (in units of ¯), i.e.,phonon frequencies. Thus the assumption of separation be-tween electronic and nuclear timescales breaks down in thevicinity of electronic PES crossings and dynamics becomesnon-adiabatic.

Several approaches exist in the literature to treat thenon-adiabatic dynamics within the framework of quantumchemistry.28–36 The most well known of these are mixed

0021-9606/2014/141(18)/184101/12/$30.00 © 2014 AIP Publishing LLC141, 184101-1

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 2: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-2 White et al. J. Chem. Phys. 141, 184101 (2014)

quantum classical treatments, the Ehrenfest and surface hop-ping methods. The Ehrenfest method is easy to implement.It requires one to run a single trajectory of the nuclei, onan average PES, with electronic populations being evaluatedon the fly.37 The simplicity of the implementation of themethod is flawed by its poor accuracy, related to its meanfield nature. Usually different electronic configurations areassociated (“entangled”) with different paths of the nucleiand so the mean field approach becomes inadequate whenthese paths are very dissimilar. In order to fix this problem,Tully introduced a surface hopping approach, which accountsfor the trajectory branching by running multiple trajectoriesthat can hop between PESs.38 Fewest switches surface hop-ping (FSSH) has become the most popular approach for prob-lems at various scales, from molecules to nanoclusters.39–45

However, the method is build on ad hoc assumptions and,as a result, frequently fails to properly describe correlationsbetween the trajectories of nuclei and electronic states. Inparticularly, it does not take into account decoherence aris-ing due to spacial separation between the components of thenuclear wavefunctions corresponding to different electronicstates (PES) and, as a results, does not properly accountsfor the interference effects, etc.46–48 Many approaches existin the literature for overcoming this decoherence problem.Some attempt to add decoherence into Tully’s surface hop-ping procedure,47, 49–61 while others involve rigorous treat-ments at the density matrix level, e.g., quantum classicalLiouville equation (QCLE)62–68 and the Meyer-Miller-Stock-Thoss36, 69, 70 formalism. While these approaches all improveupon the FSSH algorithm, they are often either too costly, leadto complicated algorithms, or may not be accurate in certainscenarios.

In this paper, we describe a method based on Monte Carlosampling of the semiclassical time-dependent wavefunctionof a molecule. Unlike common surfaces hopping approaches,the SCMC is not ad hoc by its construction, and is able toaccount for the quantum coherence effects between differ-ent photoinduced pathways. The main idea is based on writ-ing the full molecular wavefunction as a series of terms thatcan be viewed as contributions of “partial” wavefunctions ofthe system, each corresponding to a certain number of tran-sitions (hops) between electronic PES. These partial wave-functions can be sampled by a stochastic process similar tothat used in the surface hopping method: “on-the-fly” propa-gation of nuclei according to the classical equations of motionwithout prior knowledge of the PESs involved. The accumu-lated phase information characterizing each classical trajec-tory (e.g., the action along the trajectory) is further used toperform “post-processing” evaluation of multidimensional in-tegrals that correspond to each partial wavefunction using theMonte Carlo technique.

The method was first reported in Ref. 71. It was shownthat the results obtained with our method are in very goodagreement with numerically exact results for three stan-dard sample problems (the so-called Tully problems).38 Ad-ditionally, comparison of the results for 1-D and 2-D testproblems,71 demonstrated that the convergence of the methoddoes not deteriorate with an increase in the number of nucleardegrees of freedom. This is a consequence of a fact that the

stochasticity of the Monte Carlo procedure is related purely tothe non-determinism in the subspace of the electronic states(between the hops nuclei propagate according to the classi-cal equations of motion, i.e., deterministically). In this paper,we present a detailed algorithmic description of the SCMCprocedure, which should make its implementation straight-forward. Additionally, we show that the Monte Carlo proce-dure used in our method is reasonably well convergent for theproblems involving relatively few level crossings. Finally, wedemonstrate and discuss how the convergence of the wave-function is controlled by complex valued phase factors, thesize of the non-adiabatic coupling region, and the choice ofsampling function.

The QCLE procedure developed by Kapral et al.62–65

is similar in spirit to our method. The largest difference inthe approach presented here is that, while Kapral’s approachinvolves mixed classical-quantum dynamics of the densitymatrix using the Wigner-Liouville equation, our approachinvolves direct calculation of the dynamics of the wavefunc-tions. We believe our approach provides a more “straightfor-ward” path to calculating accurate non-adiabatic moleculardynamics, with minimal deviation from the commonly usedFSSH procedure. For example, such as FSSH, the dynamics inour approach always follow a single PES, whereas dynamicsused to sample QCLE involve following single and averagedPES.62 Additionally, QCLE dynamics may require branchedtrajectories.63

The paper is organized as follows: In Sec. II, we will for-mulate the problem and describe the theory behind the numer-ical method. In Sec. III, we describe, in detail, how to carryout the SCMC simulation, and provide an algorithm for a sys-tem with two PES. In Sec. IV, we analyze the results of theSCMC method when applied to Tully’s 1D test problems andan additional three electronic levels, 1D, test problem.

II. THEORY

Consider a Hamiltonian with nuclear and electronic de-grees of freedom

H = P2

2M+ Hel(R), (1)

where P is the N-component nuclear momentum operator(N is the number of independent nuclear coordinates, R≡ (R1, . . . , RN)T), M is the nuclear mass (for simplicity ofnotation we assume that all nuclear degrees of freedom havethe same mass, extension to different masses is straightfor-ward) and Hel is the electronic Hamiltonian (which includeselectronic kinetic energy and all interactions and dependsparametrically on R). The dynamics of the nuclear degreesof freedom can be conveniently represented in the adiabaticbasis in terms of the effective Hamiltonian (see Refs. 72 and73 for details)

Heff = (P − d(R))2

2M+ E(R). (2)

Here E(R) is diagonal matrix with elements being the eigen-states of the Hamiltonian Hel and d(R) is the non-adiabatic(derivative) coupling matrix, dij(R) = ¯ 〈� i(R)|∇R|� j(R)〉.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 3: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-3 White et al. J. Chem. Phys. 141, 184101 (2014)

In the absence of the non-adiabatic coupling, the PESsare decoupled and the nuclei propagate on a potential land-scape given by a particular Ei(R). Formally, in order to de-scribe quantum dynamics for a system of N particles cor-responding to Hamiltonian H0 = P2

2m+ E(R), one needs to

solve an N + 1 dimensional Schrodinger equation (for a givenPES, i.e., Ei(R)), which is an impossible task for N � 1. How-ever, in the classical limit, the problem reduces to solving Ncoupled Newton’s equations, which is not too difficult, evenfor a reasonably large number of degrees of freedom. Further-more, in this classical or semiclassical limit, one can easilyconstruct a wavefunction of the nuclei by utilizing a gaussianapproximation

�i(R, t) ∼N∏

α=1

e− (Rα−Rα

c (t))2

2σ2α (t)

+iP αc (t)(Rα−Rα

c (t))+iSi(t)

. (3)

That is, the gaussian wavepacket with average momentumPc(t) is centered around positions Rc(t), is evaluated accord-ing to the classical equations of motion, with Si(t) being theclassical action of the system produced during a time, t. Forsimplicity in this paper, we assume that the dispersion of thegaussian wavepacket corresponds to that of a free particle,σ 2(t) = σ 2(0) + i¯t/M. While this approximation is clearlyviolated for PESs with sufficient curvature, for dynamics inrealistic molecules this is practically never the case. Indeed, atypical spread of the nuclear wavepacket is of the order of nu-clear zero point motion associated with the vibrational groundstates of a molecule, which is at least an order of magnitudesmaller than the size of a typical electronic bond (which setsthe length scale of Ei(R)) due to large mass, M, of the nu-clei. Furthermore, this approximation did not lead to signif-icant deviation from exact solution in the scattering resultsconsidered below. Thus, instead of solving a N + 1 dimen-sional Schrodinger equation, a problem reduces to solving Nordinary second order differential equations.

Unfortunately, for d �= 0 this is no longer the case. Oneway to proceed is to use the mean field approach and ne-glect the spacial dependence of the wavefunction. Such ap-proach is easy to implement, but it is well known that it leadsto significant errors when branching of trajectories is impor-tant. Another approach was proposed by Pechukas in the late1960s,74, 75 which is an attempt to apply the semiclassical ap-proximation to the d �= 0 situation. Unfortunately in such caseone can only formulate an iterative procedure (which is notnecessarily well convergent) that leads to the saddle point(i.e., semiclassical) solution to the Schrodinger equation forthe molecule. Instead we propose a different approach: Wewrite down solution to the Schrodinger equation as a pertur-bative expansion in powers of d,

|�(t)〉= e−iH0t/¯|�(0)〉 +∫ t

0dt1e

−iH0(t−t1)/¯

× (−i)

2m¯(dP+Pd)e−iH0t1/¯|�(0)〉+

∫ t

0dt1

∫ t1

0dt2 . . . .

(4)

Equation (4) is an infinite series with second and higher orderterms being time ordered integrals and containing second and

higher orders of perturbation. Here we neglected the d2/2M

term, which is small (compared to the systems kinetic energy)in the semiclassical approximation. Moreover, this term is di-agonal and therefore it may be included in the definition ofH0 if needed.

Suppose that the state |�(0)〉 is a gaussian wavepacket lo-calized on the PES 1. Furthermore, let us assume that the elec-tronic subspace is two dimensional, i.e., there are only twoclose PES, the remaining PESs are separated by large gapsso that transition probabilities are infinitesimally small due torapid oscillations of the integrands in Eq. (4). In fact such sit-uation is usually the case; typically only two PES cross in arelevant region of the phase space. Then d = d12(R)σy , whereσy is a Pauli matrix and d12 is a scalar function of R. Accord-

ing to the above discussion of the d = 0 situation, the firstterm in the right hand side (RHS) of Eq. (4) is just a gaussianwavepacket that have propagated along the first PES accord-ing to classical equations of motion.

Let us consider the second term in the RHS of Eq. (4)which corresponds to a “piece” of the wavefunction that hashopped once at time t1 from PES 1 to PES 2. It is clear thatthe wavepacket right before the hopping event at t1 can beapproximated as a gaussian. It is reasonable to assume thatafter the scattering event at t1 the wavepacket retains its gaus-sian shape, but acquires an amplitude c12 (proportional tod12 at Rc(t1)). Similar assumptions apply to the higher or-der integrals. That is, we approximate the wavefunction inEq. (4) as

|�(R, t)〉

= N(

e− (R−R(0)

c (t))2

2σ2(t)+iP(0)

c (t)(R−R(0)c (t))+iS(0)(t)|1〉

+∫ t

0dt1c12

[R(0)

c (t1)]e− (R−R(1)

c (t))2

2σ2(t)+iP(1)

c (t)(R−R(1)c (t))+iS(1)(t)|2〉

+∫ t

0dt1

∫ t1

0dt2 c21

[R(0)

c (t1)]c12

[R(1)

c (t2)]

× e− (R−R(2)

c (t))2

2σ2(t)+iP(2)

c (t)(R−R(2)c (t))+iS(2)(t)|1〉 + . . .

), (5)

where the normalization constant N = π−N/4[σ (t)]−N/2 andfrom now on we will set ¯ = 1 unless stated otherwise. InEq. (5), the superscript indices for the “classical” variables,Rc, Pc, and Sc, indicate the number of jumps. For example,R(2)

c is a classical position for a trajectory of the nuclei withtwo hops: At time t1 from PES 1 to PES 2 and back at time t2,so that R(2)

c is a function not only of t, but also of t1 and t2.Equation (5) is incomplete on itself: One needs to spec-

ify the c12/21[Rc(t)] as well as the value of the momentumright after the hop. (Here we assume that neither the posi-tion nor the dispersion of the wavepacket changes immedi-ately after the hop, which is easy to check by a simple cal-culation similar to the one discussed below.) For that let usdivide the time integrals in Eqs. (4) and (5) into integralsover smaller time intervals �t, where �t is smaller than thetimescale at which the particle moves over the distance atwhich potential E1,2(R) varies significantly. Let us considerdynamics of the wavepacket at such interval �t. Assume that

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 4: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-4 White et al. J. Chem. Phys. 141, 184101 (2014)

at the beginning of the interval the shape of the wavepacket(for the nuclei) is φ1(R, 0) (which is not necessarily gaus-sian) with electrons being in state 1. Furthermore, we assumethat the wavepacket is in the semiclassical regime: On onehand it is sufficiently localized so that its width, σ , is smallerthan the distance at which potential E1,2(R) varies signifi-cantly; on the other hand its momentum is well defined sothat it is much greater in magnitude than its uncertainty (σ−1).Then the parameters c12/21[Rc(t)] and P2 (the momentum af-ter the hop, i.e., at the PES 2) must be such that the descrip-tions of the wavefunction in Eqs. (4) and (5) match. Thatis, Eq. (4) on the time interval, �t, describes dynamics of awavepacket according to an effective Hamiltonian in Eq. (2),where, within a small �t, we may determine the wavefunction

as |�ex(t)〉 = e−itH

eff φ1(R, 0)|1〉 by expanding in powers of tand setting d and E to be independent of R. Similarly, we mayevaluate the wavefunction for the gaussian approximation inEq. (5). Up to the second order in t, which is sufficient for ourpurposes, we may write

|�g(t)〉 = (1 + c21c12t2/2)φ1(R, t)|1〉 − ic12tφ2(R, t)|2〉,

(6)

where φ2(R, t) is a normalized wavepacket at the second PES.The t2 term in Eq. (6) arises due to two hop contribution inEq. (5). Note that the wavefunction in Eq. (6) is normalized upto O(t2). Since c12/21 and P2 = ∫

dRφ∗2 (R, t)Pφ2(R, t) define

occupation probabilities and system’s momentum, we requirethat

〈�g(t)|σz|�g(t)〉 = 〈�ex(t)|σz|�ex(t)〉, (7)

〈�g(t)|P|�g(t)〉 = 〈�ex(t)|P − d12σy |�ex(t)〉. (8)

Equations (7) and (8) provide the matching (boundary) con-ditions that determine values of c12/21 and P2. Note that in theRHS of Eq. (8) we have used the “true” momentum of thesystem, which is given by its velocity MR = iM[Heff , R]

= P − d12σy . The left hand side (LHS) of Eq. (8) gives1 + 2c21c12t2, while the RHS results in 1 − 2(d12P1t)2/M2,where P1 = ∫

dRφ∗1 (R, t)Pφ1(R, t). Then we obtain that

c12(R) = (d12(R) · P1)/M = d12(t)

and

c21(R) = (d21(R) · P2)/M = d21(t), (9)

with P1 being the momentum right before the hop.Similarly, the LHS of Eq. (8) gives P1 + c21c12t2(P2

− P1), where we have assumed that t is small enough andso the momenta P1 and P2 do not change during the evolu-tion over the time interval t. The RHS yields P1 − d12(R)(E2− E1)(d12(R) · P1)t2/M, so that

P2 − P1 = M(E2 − E1)d12(R)

d12(R) · P1

. (10)

Equation (10) gives a condition for the re-scaling of momen-tum of the wavepacket which, at first glance, may seem tobe in contradiction with that proposed by Herman76 and usedby Tully38 in his surface hopping algorithm. Indeed, while in

Eq. (10) the change in the momentum at the hop is along d12,Eq. (10) does not conserve the total energy at the hop. How-ever, careful inspection reveals that Eq. (10) is simply the ze-roth and first order terms in an expansion, in powers of

E2−E1

P21/2M

,

of the full energy conservation condition. This is, presumably,an artifact of the short time approximation that we used in thederivation of Eq. (10). Indeed, it is easy to check that Eq. (10)conserves energy approximately, up to the leading order in�E = E2 − E1. Furthermore, we have run the test problemsconsidered below using condition of Eq. (10) as well as usingthe exact energy conservation condition at the hop and foundvery good agreement between the two in a very broad range ofinitial momenta. This is a consequence of a fact that the maincontribution to the integrals in Eq. (5) comes from the regionwhere �E is small and so the energy conservation at the hopis satisfied. Therefore, we conclude that the two conditionsare essentially identical and in the following we will be us-ing the Herman/Tully criterion to re-scale the wavepacket’smomentum at the hop. In Sec. III, we will describe an effi-cient numerical algorithm that evaluates the wavefunction inEq. (5) using Monte Carlo technique.

III. SCMC CALCULATION

The SCMC provides a first principles method for calcu-lating non-adiabatic molecular dynamics, with similar cost tosurface hopping methods. The calculation occurs in two mainsteps: (A) the non-adiabatic propagation of classical trajecto-ries, and (B) the use of trajectory data to sample the integralsin Eq. (5), with cij[Rc(t)] = dij(t). The classical propagationis similar to standard surface hopping algorithms.29, 38, 76 Forsimplicity, we assume that we have only two electronic statesto consider. Generalization to higher numbers of states isstraightforward.71 In Subsections III A and III B, we describe,in algorithmic fashion, the surface-hopping dynamics, A, andthe Monte Carlo calculation of the time-dependent semiclas-sical wavefunction, B. Figure 1 shows a diagrammatic map ofthe algorithm.

A. Semiclassical molecular dynamics

A1: All trajectories are initialized with the same positionand momentum. Various tracked values are initialized(see Fig. 1). We begin propagation of the nuclearwavepackets.

A2: Begin time step. Supplied with a molecular ge-ometry, Rc, quantum chemistry methods (e.g.,DFT and TDDFT) are used to calculate theelectronic state energies, Ei(t), the Hellmann-Feynman forces (electronic gradients), Fc, thefirst order non-adiabatic coupling vectors (NACV),dij (Rc) = 〈�i(Rc)| ∂

∂Rc

�j (Rc)〉, and scalars dij (t)

= 〈�i(Rc(t))| ∂∂t

�j (Rc(t))〉 = dij (Rc) · Rc(t).A3: The classical particles are propagate for one time step,

∂t, on the current adiabatic surface, i, using New-ton/Lagrange equations of motion,

∂tPc(t) = Fc(t) = − ∂

∂Rc

Ei(Rc(t)), (11)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 5: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-5 White et al. J. Chem. Phys. 141, 184101 (2014)

Begin Surface Hoppingk=0

Initialize Trajectory kRk = R(0), Pk = P(0), Sk = 0

k = 0, Dk = 1, k = 1t = 0, mk = 0, sk = s(0)

Quantum ChemistryEk=Esk

(Rk)Fk=- Esk

(Rk)/ Rkdk=dsksk

(Rk)

Equations of Motiont = t + tRk=Pk t/M + Rk

Pk=Fk t + Pk

Sk= (Pk2/2M - Ek) t + Sk

Pk2/2M + Es(Rk) > Es(Rk) ?

t > tmax?

NO

YES

k=1+kk = N ?

YES

YES NO

NO

Calculate sksk(Rk) sksk

(Rk)=0

k= skskt + k

Generate RNRN< sksk

?

YES

NO

Dk = dk Dk

k = sksk(t) k

mk = 1 + mk

sk = sk

Process Trajectory Datak=0For all m & s:Nm,s = 0, Am,s = 0

Nmk,sk= 1 + Nmk,sk

Amk,sk= exp[- k] + Amk,sk

k = 1 + k , k = N ?

YES NO

For all m & s:Am,s = (Nm,s/N) / Am,s

Cs=0

Calculate Wavefunctionsk=0

(R)= {((M )2/[ (M )

4+t

2])

0.25

exp[iPkR + (R-Rk)2/(

2+it/M)]}

Cs(R)=(DkAmk,sk/ k) exp[i(Sk-PkRk)] (R) + Cs(R)

k = 1 + k , k = N ?

YES NO

Calculate Desired Expectation Values

A1:

A2:

A3:

A4:

A5:

A6:

B1:

B2:

B3:

B4:

FIG. 1. Example for semiclassical Monte Carlo surface hopping algorithm for a two electronic state system. Green boxes – Part A (dynamics). Purple boxes –Part B (Wavefunction calculation). Definitions are same as text.

by some numerical method, e.g., Verlet, Runga-Kutta.The classical action along the trajectory,

S(t) =∫ t

0dt ′

N∑α

P 2α (t ′)

2Mα

− E(t ′)

=N∑α

∫ Rαc (t)

Rαc 0

dRαc Pα(Rα

c ), (12)

is calculated for subsequent use in part B. We as-sume our propagated gaussian nuclear wavepackets are“free,” i.e., they broaden with time, independent of thePES.

A4: At the end of the time step, we determine whether ornot to hop to a new state, j. First, we determine whetherthe hop is allowed or frustrated by, the assumed, con-servation of energy, i.e.,

P2c(t ′)2M

+ Ei(Rc) ≥ Ej (Rc). (13)

If the hop is allowed the probability, γ i→j(t), is calcu-lated. If not γ i→j(t) = 0. Unlike other surface hoppingmethods,29, 38, 76 the final result is formally independentof the choice of hopping rate, assuming that the hoppingrate is nonzero everywhere dij(t) is non-zero. However,the choice of hopping rate is crucial to achieve rapidconvergence of the result.

A5: At each time step, the hopping probability to the newstate is added to the integral

(t) =l=m∑l=0

∫ tl+1

tl

dt ′ γsl→s

l+1[Rc(t ′)]. (14)

Here, m is the number of hops over the course of thetrajectory, and tl is the time at which the lth hop occurs(tm + 1 ≡ t, sm + 1 ≡ j). If RN < γ i→j, where RN is agenerated random number between 0 and 1, then thetrajectory hops to state j.

A6: If a hop occurs, the products

D(tm−1 . . . t0) =m−1∏l=0

dslsl+1

(tl), (15)

and

�(tm−1 . . . t0) =m−1∏l=0

γsl→s

l+1(tl) (16)

are multiplied by dij(t) and γ i→j(t), respectively. Thenumber of hops, m, is increased by one.

The procedure is repeated for the next time step on ei-ther the original or new state. These dynamics continue forthe desired time. Once completed, the final position, Rc(t),and momentum, Pc(t), of the trajectory are stored. The pro-cess is repeated for the specified number of trajectories. In

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 6: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-6 White et al. J. Chem. Phys. 141, 184101 (2014)

calculations of realistic molecular systems, A2, i.e., quantumchemistry calculation of state energies, forces, and non-adiabatic couplings, will be the most computationally de-manding step. However, due to the independence of thedynamical trajectories, the loop beginning at A1 is triviallyparallelizable. This will greatly aid in the application of themethod to larger systems.

B. Post-processing

Information collected by the surface hopping dynamicsis further analyzed in the second step at negligible computa-tional expense when compared to part A. Namely, we calcu-late the nuclear wavefunctions,

�i(R, t) =∑m

∫dmt Im

i (R, t), (17)

or probabilities, Pi = ∫dR |� i(R, t)|2. Here t ≡ {t1, . . . tm,

t} is the set of time integration variables and∫

dmt≡ ∫ t

0 dtm∫ t

m

0 dtm−1 . . .∫ t2

0 dt1. We independently sample in-tegrals,

∫dmt Im

i (R, t), which differ in the number of hops,m, and final state, i. Additionally, if there is significant sep-aration between wavepackets, e.g., wavepackets which arereflected or transmitted in 1D, then integrals contributing tothose wavepackets are also sampled separately. While not re-quired formally, this consideration of spacial overlap signif-icantly improves convergence rates with minimal loss of in-formation. Thus, trajectories are sorted by the number of hopsand final electronic state, and grouped into spatially separatedcontributors of the partial wavefunctions.

Using the standard Monte Carlo approach, the integral∫dmt Im

i (R, t), can be expressed in terms of an expectationvalue ∫

dmt Imi (R, t) =

∫dmt

Imi (R, t)ρm

i (t)× ρm

i (t)

≈ 1

Nmi

×Nm

i∑k=0

Imi (Rk, tk)

ρmi (tk)

. (18)

Here, k is the index of a trajectory from step 1, Nmk is the

total number of trajectories with m hops and ending in statei. ρm

i (tk) is the value of the probability distribution, ρmi (t), for

the trajectory k.The kth value of the probability distribution can be ex-

pressed as ρmi (tk) = �m

i (tk)

Ami

. Here �mi (tk) is given by Eq. (16),

and Ami = ∫

dmt �i(t) is the normalization coefficient for tra-jectories with m hops finishing in state i. Am

i can be deter-mined using statistical considerations of the stochastic pro-cess defined by the rate γ . We consider the probability of asingle trajectory finishing in state i with m hops for all valuesof t (see the supplementary material of Ref. 71),

pi =∫

dmt �i(t)e−

i(t). (19)

This probability can also be expressed as an expectationvalue, using the same probability distribution, ρm

i (t) = �i(t)

Ami

,

from Eq. (18),

pi = Ami

Nmi

Nmi∑

k=0

e−i(t

k). (20)

i(ti) is given by Eq. (14). We seek the value of Ami . Thus, we

assume the number of trajectories, Nmi , is sufficiently large so

that we can replace pi with Nmi /N , and rearrange

Ami ≈ (

Nmi

)2[N

Nmi∑

k=0

e−mi (t

k)

]−1

,

and

ρmi (tk) = �m

i (tk)N(

Nmi

)2

Nmi∑

k=0

e−mi (t

k). (21)

B1: We first determine Nmi and calculate the sum in Eq. (21)

for every i/m pair.B2: We calculate each Am

i .B3: For a given trajectory Im

i (Rk, tk) is easily calculated, seeEq. (5), from the surface hopping data, Eqs. (11), (12)and (15),

Imi (Rk, tk) = Dk(tk) × ei[Si (t

k)−Pk (t)Rk

c (t)]

×φ(R, Rk

c, Pkc, t

), (22)

where φ is a “free” gaussian wavepacket,

φ(R, Rk

c,Pkc, t

) = 4

√M2σ 2

π [M2σ 4 + t2]eiPk (t)Re

[R−Rkc (t)]2

σ2+ itM ,

(23)

with the final position and momentum of the trajectoryk, and an initial width, σ . At this point the nuclear wavefunctions can be calculated by summing over all the tra-jectories, Eq. (18).

B4: Once � i(R, t) is known, expectation values (e.g., x, P,density matrix) can be calculated. When the number oftrajectories is low, the random error in � i(R, t) can bereduced by normalization. However, the effect is mini-mal in a well converged calculation.

Some notes on the case of more than 2 electronic states:(1) The rate in Eq. (14) would be replaced with a sum overthe rates to all possible new states. (2) Eqs. (15) and (16) re-main unchanged. (3) Determination of which state to hop tocan be determined as it is in Ref. 38. (4) For every possiblepath of intermediate electronic excited states, there is an inte-gral which must be sampled independently. Thus each distinctpath has a different probability distribution, Eq. (21).

IV. ANALYSIS OF SCMC METHOD

In order to test the SCMC procedure, we consider threetwo-level models proposed by Tully.38 Figure 2 shows the en-ergy eigenstates and NACVs of these three one-dimensionalproblems. For the exact form of the Hamiltonians see Ref. 38.For these three problems, scattering probabilities calculatedusing the SCMC, FSSH, and Ehrenfest methods are com-pared to the exact time-dependent Schrödinger equation inGorshkov et al.71 In Figure 3, we represent those scattering

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 7: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-7 White et al. J. Chem. Phys. 141, 184101 (2014)

-0.2

-0.1

0

0.1

0.2

-10 -5 0 5x

-0.01

0

0.01

-2 0 2x

-0.05

0.05

-5 0 5x

-0.05

0.05

-10 -5 0 5 10x

TU TU

TURU

RL

TLRLTLRL

TLE1

E2

|d12|

|d12|15

|d12|100|d12|100

d12100

d32100

E1

E2

E3TU

RM

TLRL

FIG. 2. Test problems. (Top Left) Tully Problem 1-Single avoided crossing.(Top Right) Tully Problem 2 – Double avoided crossing. (Bottom Left) TullyProblem 3 – Extended coupling with reflection. (Bottom Right) Three LevelModel X. Green dotted (Purple dashed/Blue dashed-dotted) line is lower (up-per/middle) electronic eigenstate. Black (red) solid line is the non-adiabaticcoupling vector, d12(23)(x). All values are in atomic units.

results. For Problems 1 and 3: 25 000 and 10 000 trajectorieswere used for the SCMC and FSSH approaches, respectively.For Problem 2: 75 000 trajectories were used in the SCMCcalculation, while only 10 000 were used in the FSSH cal-culation. Here we seek to provide in-depth analysis of theseresults and insight into the method. All parameters and resultsin this section are in atomic units.

0.2

0.4

0.6

0.8

1.0 FSSHSCMC

0.2

0.4

0.6

0.8

10 15 20 25 30

EhrenfestSchodinger

0.2

0.4

0.6

20 30 40 50

0.2

0.4

0.6

0.8

1.0

5 10 15 20 25

k

0.2

0.4

5 10 15 20 25 30

20 30 40 50

k

0.6

0.8

1.0

0 10 20 30 40 50

(c) (d)

(a)(b)

TL

TU

RL

RU

TU

TL

TL

TU

FIG. 3. Scattering probabilities for Tully’s problem set, using time-dependent Schrödinger equation (black solid line), Ehrenfest method (bluedotted line), fewest switching surface hopping (purple circles), and SCMCmethod (green squares). (a) Problem 1 – Transmission on lower (inset-upper) surface. (b) Problem 2 – Transmission on lower (inset-upper) surface.(c) Problem 3 – Reflection on lower (inset-upper) surface. (d) Problem3 – Transmission on upper (inset-lower) surface. Results previously reportedin Ref. 71.

Problem 3 provides a simple model problem where thestandard implementation of the FSSH quantitatively and qual-itatively fails to reproduce the exact TD-Schrödinger result,see Figs. 3(c) and 3(d). It is well known that this failure isdue to the lack of decoherence.38, 77 The equation of motionfor the electronic density matrix, or complex expansion co-efficients, does not include the effect of bifurcation of nu-clear wavepackets. Corrections can be introduced by handinto the FSSH algorithm to correct for this problem.47, 49–61

In the SCMC approach, we do not need to propagate the elec-tronic density matrix in order to determine hopping rates. Wefind the simple hopping rate γ 1 → 2(t) = |d12(t)| leads to ac-curate results for all of the test problems, see Figs. 3 and 4.Thus, since we do not utilize the electronic density matrix,we do not encounter the “over coherence” problem. The left(right) column of Figure 4 shows the results of the SCMC(time-dependent Schrödinger) calculation for Problem 3 withan initial wavefunction

�(t = 0, x) =⎛⎝ (πσ 2)−

14 × exp{ikx − (x−x0)2

2σ 2 }0

⎞⎠ ,

R e[ 1](x)| 1|

2(x)x=39.8

p=29.9

| 1|2

x>0

=0.69

R e[ 1](x)| 1|

2(x)x=-12.4

p=-10.0

| 1|2x<0

=0.12

R e[ 1](x)| 1|

2(x)x=39.7

p=29.9

| 1|2x>0

=0.70

R e[ 1](x)| 1|

2(x)x=-12.4

p=-10.0

| 1|2x<0

=0.09

R e[ 2](x)| 2|

2(x)x=-12.1

p=-9.8

| 1|2x<0

=0.21

R e[ 2](x)| 2|

2(x)x=-12.1

p=-9.8

| 1|2x<0

=0.19

FIG. 4. Scattered wavepackets problem 3 at t = 4059. (Left column) SCMCmethod using 25 000 trajectories. (Right column) TD Schrödinger equation.(Top (Center)) Reflected wavepacket on upper (lower) electronic eigenstate.(Bottom) Transmitted wavepacket on lower electronic eigenstate. (Coloredsolid) Wavefunction. (Black dashed) Probability density.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 8: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-8 White et al. J. Chem. Phys. 141, 184101 (2014)

with initial position: x0 = −12, momentum: k = 10, andwidth: σ = √

200/k, starting on the lower eigenstate: |ψ1〉.The snapshot is taken after the wavepacket has gone throughthe interaction region with a portion being transmitted on thelower adiabatic surface (bottom), or reflected on either thelower (middle) or upper (top) surface. The SCMC wavepack-ets have the same momentum, p, and positions, x, as the exactsolution, indicating that our semiclassical dynamics, i.e., as-sumption of conservation of energy and Newtonian equationsof motion is reasonable. The transmitted wavepacket (bottom)calculated by SCMC shows significantly less broadening thanthe exact result. This is due to our “free” gaussian approxima-tion, which ignores the increase in broadening due to the neg-ative second derivative in the energy of the lower surface fromx = −5 → 0. In principle, this is simple to account for,78 how-ever calculation of the second derivative of the PES would beprohibitively expensive in a many-dimensional problem.

To test the limits of our method, we consider the threelevel crossing “Model X” problem from Ref. 48 (see Ref. 48for Hamiltonian). The PES can be seen in Fig. 2(d). We con-sider a particle entering from negative x on the middle surface.The particle has a well-defined momentum (plane wave forthe exact solution, σ = √

200 000/k for SCMC). The largevalue of σ , as well as the multiple crossings, and possibility ofreflection makes “Model X” a very difficult problem for semi-classical and surface hopping methods. However, with enoughtrajectories, 5 × 105, the SCMC can approach the exact solu-tion. Figure 5 shows the unnormalized scattering probabilitiesfor this problem, with SCMC compared to the exact solution.The most difficult part of the exact dynamics to recover isthe period and phase of the oscillations in the transmission on

0

0.1

0.2

0 10 20 300

.05

0.1

0 10 20 30

0.0

0.2

0.4

0.6

0.8

0 10 20 30

k

0

0.1

0.2

0 10 20 30

k

RL

TL

RM

TU

FIG. 5. Scattering probabilities for three level crossing problem (Fig. 2(d)),using time-dependent Schrödinger equation (purple dotted line), and SCMCmethod (green solid line). (Top Left) Reflection on lower surface. (Top Right)Reflection on middle surface. (Bottom Left) Transmission on lower surface.(Bottom Right) Transmission on upper surface. Schrödinger equation resultswere previously reported in Ref. 48.

the lower surface.48 The SCMC method correctly reproducesthis phase and period. Within the Fewest Switches framework,careful corrections to decoherence, e.g., augmented-FSSH,79

as well as a phase correction, e.g., Shenvi phase correction,59

are required to correctly achieve the period and phase of theoscillation.48

While the SCMC method does a good job of reproduc-ing the exact result, it does so at the cost of a high numberof trajectories. For a problem like “Model X” significant im-provement in the convergence could be achieved by using theSCMC only in the vicinity of the strong NAC, then adiabat-ically propagating wavepackets with the normalization, posi-tion, momentum, and phase given by SCMC. This is simi-lar, in essence, to the branching scheme recently proposed byHerman, which significantly improved convergence in a twocrossing problem.80

We have shown how the details of the problem, number,and nature of level crossings, can affect the number of trajec-tories required, i.e., 2.5 × 104 for a simple avoided crossing(∼1%−2% absolute error) with 3 × as many required for thedouble avoided crossing. In Subsection IV A, we will discussthe factors which lead to this increased cost in trajectories.

A. Role of phases

While the SCMC method does a good job of reproducingthe exact result, including when FSSH and Ehrenfest methodsfail, it does so at an increased cost in trajectories. In order todetermine the areas where SCMC is most applicable, we seekto understand how the convergence rate is affected by differ-ent models and parameters. First, we look at the simple singlelevel crossing, Problem 1. Figure 6 (Inset) shows the relativeerror, δrel, in the calculated probabilities of transmission onthe upper (lower) surfaces, TU (TL), as a function of the to-tal number of trajectories, N. For all momentums and bothprobabilities, the convergence obeys an approximate power

5

15

rel(%)2.5

7.5

5 25 5

5 25 5 50 7

0 75

N ( 1000)

k=10k=20k=30k=10k=20k=30rel(%)

N ( 1000)

TL

TU

FIG. 6. Average relative error percentage (δRel) of SCMC calculation, with-out normalization, for Problem 1 as a function of the number of trajecto-ries (N) for k = 10, 20, and 30. Twenty calculations per data point. Filled(Empty) Markers – Transmission probability on the upper (lower) surface.Solid (Dashed) lines are respective fits to the data of the form δrel = a√

N+ b.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 9: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-9 White et al. J. Chem. Phys. 141, 184101 (2014)

law of the form δRel = a√N

+ b (see line fits). This O( 1√N

)dependence of the error is standard to Monte Carlo meth-ods. For TU, we see faster convergence with increasing initialmomentum.

This k dependence can be understood by looking at thephase in Eq. (22). Assuming a one-dimensional system withtwo electronic states, we can calculate the phase for a trajec-tory which begins on the lower surface and hops once,

F1(tf , t1)

= P2(tf )

[ ∫ t1

0dt

P1(t)

M+

∫ tf

t1

dtP2(t)

M+ x0

]

+∫ t1

0dt

[P1(t)2

2M− E1(t)

]+

∫ tf

t1

dt

[P2(t)2

2M− E2(t)

].

(24)

Since all trajectories have the same final and initial times, wecan subtract off any term which does not depend on the timeor position of the hop. Additionally, if we assume that theinitial momentum, k, is large enough that we do not have anyreflection, and tf is large enough that P2(tf) is a constant, wecan rewrite Eq. (24) as a relative phase that depends on theposition of the hop,

F1(x1) =P2f

∫ x1

x0

dx

[1 − P2(x)

P1(x)

]

+∫ x1

x0

dx

[P1(x) − P2(x)2

P1(x)

]. (25)

With the same considerations we can write the relative phasefor a trajectory with two hops,

F2(x1, x2) = P1f

∫ x2

x1

dx

[1 − P1(x)

P2(x)

]

+∫ x2

x1

dx

[P2(x) − P1(x)2

P2(x)

]. (26)

In Figure 7, we can see that the number of oscillations,which must be sampled, is reduced in the one (a) and (b) andtwo (c)–(f) hop integrals when we go from the low momen-tum, k = 10, to the high momentum k = 30. Indeed for inte-grals of all numbers of hops the number of oscillations willdecrease with an increasing momentum. The “smoother” in-tegrands associated with higher momentum are easier to sam-ple via Monte Carlo methods. For the TL in Figure 6, the ef-fect competes with constant error sources, due to, e.g., thefree gaussian approximation, which appears as an increasingrelative error with decreasing average probability (probabil-ity of TL decreases with increase k). This leads to the non-monotonic k dependence.

When going from a single (Problem 1) to double (Prob-lem 2) crossing systems the integrated area of |d12(x)| in-creases from 1.57 to 2.60. The number of trajectories requiredto sample an integral with m hops will approximately scale

0

1.5

-1.5 0 1.5

x1

x2-x1

-1.5 0 1.5

x1

0

1.5

x2-x1

-1.5

0

1.5(a) (b)

(e)

(c)

(f)

(d)

k=10 k=30

FIG. 7. Relative Phases, Problem 1, for single (a) and (b), Eq. (25), and dou-ble (c)–(f), Eq. (26), hop trajectories for k = 10 (left) and k = 30 (right).(a) and (b) Black solid – |d12(x)|, Purple dotted – d12(x) × Re[F1(x)], Greendashed – d12(x) × Im[F1(x)]. (c) and (d) Color map of

∣∣d12(x1)d21(x2)×Re[F2(x1, x2)]

∣∣. (e) and (f) Color map of∣∣d12(x1)d21(x2)×Im[F2(x1, x2)]

∣∣.∫dx |d12(x)| � 1.57.

with∫

dx |d12(x)| as

1

m!

[ ∫dx |d12(x)|

]m

. (27)

However, integrals with more hops, generally, will contributeless to the total wavefunction. Additionally, oscillations dueto the phases, Eqs. (25) and (26), increase for Problem 2 (seeFig. 8). These two, not independent, effects result in a muchslower convergence rate for Problem 2 as compared to Prob-lem 1 or 3. It was shown in Ref. 71 that ∼3 times the numberof trajectories is required to obtain similar precision for Prob-lem 2 as for Problems 1 and 3.

B. Choice of hopping rate

In principle, the hopping rate, γ i→j, is arbitrary. It is sim-ply the choice of a probability distribution for sampling theintegral, Eq. (5). A better probability distribution leads to lesssamples being required to accurately calculate the correct re-sult. The closer the probability distribution is to the abso-lute value of the normalized integrand, the better the sam-pling. If we knew the normalization factor for each integrandImi , then we would not need to run simulations. Even if we

could make a good estimate of these normalization factors,we would have to run separate sets of trajectories, with differ-ent hopping rates, to sample different outcomes. Additionally,it is not clear, in general, how to include the complex phasesinto a probability distribution, and it is difficult to know thephases in Eq. (22) a priori. However, by choosing γ i→j(t)= |dij(t)| we can at least include the factor D(t) into the prob-ability distribution. D(t) provides bounds and amplitude tothe integrands in Eq. (5). While there are methods to improvesampling of complex valued integrands, e.g., stationary phaseapproximation,81, 82 they emphasize sampling in the same

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 10: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-10 White et al. J. Chem. Phys. 141, 184101 (2014)

0

3

-3 0 3

x1

x2-x1

0

3

x2-x1

-0.75

0.0

0.75 (a)

(c)

(b)

k=15

FIG. 8. Problem 2 – Relative Phases, for single (a), Eq. (25), and dou-ble (b) and (c), Eq. (26), hop trajectories for k = 15. (a) Black solid– |d12(x)|, Green dotted – d12(x) × Re[F1(x)], Purple dashed – d12(x)× Im[F1(x)]. (b) Color map of

∣∣d12(x1)d21(x2) × Re[F2(x1, x2)]∣∣. (c) Color

map of∣∣d12(x1)d21(x2) × Im[F2(x1, x2)]

∣∣. ∫dx |d12(x)| � 2.60.

region as D(t), i.e., regions with small �E between PES. Thisleads to minimal or no improvement. Investigating other hop-ping rate choices, which would effectively include the effectsfrom the phases, is an ongoing project.

One choice which may come to the readers mind, is thefewest switches hopping rate from Tully,38

γ FSi→j (t) = 2Im[aji(t)Vji] − 2Re

[aji(t)dji

aii(t)

]

= 0 if < 0,

with

∂taji(t) =

∑k

[i(ajk(t)Vki − Vjkaki(t))

+ ajk(t)dki − djkaki(t)

]. (28)

Here a is the electronic density matrix. Figure 9 shows the rel-ative error of the SCMC calculation for the different hoppingrates, γ FS

i→j and dji. In both the cases where FSSH methodworks (Problem 1 – Fig. 9(a)), and when the FSSH methodfails (Problem 2 – Fig. 9(b)), the error is extreme (see Fig. 9(b)inset) if the results are unnormalized, and still high even whennormalized (Fig. 9). There is a problem with using γ FS

i→j . Afterone surface hop, the returning hop is frustrated, for a time, dueto the resulting negative sign in γ FS

i→j . This leads to regions ofvanishing hop probability in trajectories with more than one

0

10

20

5 25 50 75

d21 k=15d21 k=40d21 k=50

FSk=15

FSk=40

FSk=50

5

15

d21 k=10d21k=20d21 k=30

FSk=10

FSk=20

FSk=30

rel(%)

rel(%)

N ( 1000)

50

100

150

5 25 50 75

N ( 1000)

50

60

5 25 50 75

N ( 1000)

rel(%)

rel(%)

(a)

(b)

N ( 1000)

FIG. 9. Averaged relative error percentage (δRel) of SCMC calculation asa function of the number of trajectories (N). (a) Problem 1, Transmissionprobability on the upper surface with normalization for k = 10, 20, and 30.(b) Problem 2, Transmission probability on the lower surface with normaliza-tion for k = 15, 30, and 50. (b) – (inset) Averaged relative error percentage ofunnormalized results. Filled markers are for γ 1 → 2 = |d21|. Empty markersare for γ1→2 = γ FS

1→2. Solid (Dashed) lines are respective fits to the data ofthe form δrel = a√

N+ b.

hop. Even without the phases this would, in principle, lead toerroneous results.

Additionally, when the integrals are unevenly sampledwithout consideration of the phases, as they are when usingγ FS

i→j , the sensitive constructive/destructive behavior of thephases leads to the large error. We believe there will alwaysbe situations where any ad hoc hopping rate will lead to poorconvergence due to the imperfect accounting of the phases.Thus, while we have repeatedly referred to the SCMC proce-dure as surface hopping with a post-processing step, it shouldnot be blindly applied to any surface hopping procedure, suchas FSSH. However, it remains to be seen if other surface hop-ping probabilities do provide a good probability distributionfor Monte Carlo sampling. The difference in the result before(Fig. 6 and inset of Fig. 9(b)) and after (Figs. 9(a) and 9(b))normalization tells us how well the method is working. If thechange is minimal (compare Figs. 6 and 9(a)), the SCMC pro-cedure is working well for the system. If the change is large(compare Fig. 9(b) body and inset), the SCMC procedure isfailing.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 11: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-11 White et al. J. Chem. Phys. 141, 184101 (2014)

V. CONCLUSION

The semiclassical Monte Carlo method provides an ap-proach, which is not ad hoc, for calculating non-adiabaticdynamics in extended molecular systems. It is based on anexpansion of the full molecular wavefunction into partialwavefunctions corresponding to the number of transitions be-tween electronic states. The procedure involves standard sur-face hopping dynamics with an additional post-processingstep (calculation of the wave functions from collected data).While in the surface hopping method the hopping probabilitycontrols all of the dynamics; in SCMC it is primarily inter-ference between trajectories with different (or similar) phasesthat determines the final result. For this reason, the SCMCmethod does not suffer from the same “over coherence” prob-lem as the FSSH approach.

Some notes on the applicability of this method: (1) Whilethe phases lead to difficulty in sampling for small momen-tum, for higher momentums the sampling becomes easier. (2)Additionally, we believe that convergence rate does not in-crease with increased number of system coordinates.71 (3) In-creasing the number of crossing regions increases the num-ber of trajectories required for sampling, see Eq. (27), dueto an increase in the area of non-adiabatic coupling and anincrease in the number of oscillations due to the phases. (4)While it has been shown that, in some systems, FSSH ap-proach requires complete coupling, i.e., the quantum mechan-ical electronic dynamics requires calculation of NACV for allcombinations of states ( 1

2Ns[Ns − 1] NACV calculations),83

the SCMC method with a hopping rate of |di, j(t)| requiresonly partial coupling (Ns − 1 NACV calculations). WhenNs becomes large the calculation of NACVs can become thebottleneck of the calculation. Thus improved scaling in Nsmay make up for the additional trajectories needed for sam-pling the wavefunction. (5) Alternatively, one could utilizethis method only in the vicinity of non-negligible NAC, inessence attacking the multi-crossing problem one crossing ata time. We leave development of such a procedure, with com-parison to full SCMC and TD-Schrödinger calculations, aswell as continued efforts in improving convergence rates, andapplication of the SCMC approach to realistic systems for fu-ture work.

ACKNOWLEDGMENTS

We acknowledge the support of the U.S. Department ofEnergy through the Los Alamos National Laboratory (LANL)LDRD Program. LANL is operated by Los Alamos NationalSecurity, LLC, for the National Nuclear Security Administra-tion of the U.S. Department of Energy under Contract No.DE-AC52-06NA25396. We acknowledge the support of theCenter for Nonlinear Studies (CNLS) and the Center for Inte-grated Nanotechnology (CINT) at LANL. We also thank J.E.Subotnik for sharing his numerical results reported in Ref. 48.

1C. J. Cramer, Essentials of Computational Chemistry: Theories and Models(John Wiley and Sons, 2004).

2A. Szabo and N. S. Ostlund, Modern Quantum Chemistry: Introduction toAdvanced Electronic Structure Theory (McGraw-Hill, 1989).

3“Ab initio molecular dynamics,” in Computational Methods in Catalysisand Materials Science (Wiley-VCH, 2009), pp. 93–120.

4C. Ullrich, Time-Dependent Density-Functional Theory: Concepts and Ap-plications (Oxford University Press, 2012).

5A. Nitzan, Chemical Dynamics in Condensed Phases: Relaxation, Transferand Reactions in Condensed Molecular Systems (Oxford University Press,2006).

6G. C. Schatz and M. A. Ratner, Quantum Mechanics in Chemistry (Dover,2002).

7C. J. Smallwood, W. B. Bosma, R. E. Larsen, and B. J. Schwartz, J. Chem.Phys. 119, 11263 (2003).

8G. Cui and W. Fang, J. Phys. Chem. A 115, 1547 (2011).9J. M. Hostettler, A. Bach, and P. Chen, J. Chem. Phys. 130, 034303 (2009).

10A. Matsugi, J. Phys. Chem. Lett. 4, 4237 (2013).11M. H. Kim, L. Shen, H. Tao, T. J. Martinez, and A. G. Suits, Science 315,

1561 (2007).12J. Clark, T. Nelson, S. Tretiak, G. Cirmi, and G. Lanzani, Nat. Phys. 8, 225

(2012).13A. J. Neukirch, L. C. Shamberger, E. Abad, B. J. Haycock, H. Wang, J.

Ortega, O. V. Prezhdo, and J. P. Lewis, J. Chem. Theor. Comput. 10, 14(2014).

14A. Kazaryan, Z. Lan, L. V. Schäfer, W. Thiel, and M. Filatov, J. Chem.Theor. Comput. 7, 2189 (2011).

15R. Mathies, C. Brito Cruz, W. Pollard, and C. Shank, Science 240, 777(1988).

16R. A. Marcus, J. Phys. Chem. 67, 853 (1963).17P. F. Barbara, T. J. Meyer, and M. A. Ratner, J. Phys. Chem. 100, 13148

(1996).18P. Peumans, S. Uchida, and S. R. Forrest, Nature (London) 425, 158 (2003).19S. M. Falke, C. A. Rozzi, D. Brida, M. Maiuri, M. Amato, E. Sommer, A.

De Sio, A. Rubio, G. Cerullo, E. Molinari, and C. Lienau, Science 344,1001 (2014).

20T. G. Goodson, Acc. Chem. Res. 38, 99 (2005).21E. Collini and G. D. Scholes, Science 323, 369 (2009).22H. Lee, Y.-C. Cheng, and G. R. Fleming, Science 316, 1462 (2007).23M. Reufer, M. J. Walter, P. G. Lagoudakis, A. B. Hummel, J. S. Kolb, H.

G. Roskos, U. Scherf, and J. M. Lupton, Nat. Mater. 4, 340 (2005).24S. Mai, P. Marquetand, and L. González, J. Chem. Phys. 140, 204302

(2014).25A. Cannizzo, A. M. Blanco-Rodríguez, A. El Nahhas, J. Sebera, S. Záliš,

A. Vlek, Jr., and M. Chergui, J. Am. Chem. Soc. 130, 8967 (2008).26R. S. Becker, A. P. Pelliccioli, A. Romani, and G. Favaro, J. Am. Chem.

Soc. 121, 2104 (1999).27S. V. Kilina, D. S. Kilin, and O. V. Prezhdo, ACS Nano 3, 93 (2009).28L. Wang, D. Trivedi, and O. V. Prezhdo, J. Chem. Theor. Comput. 10, 3598

(2014).29J. C. Tully, J. Chem. Phys. 137, 22A301 (2012).30W. H. Miller, Classical-Limit Quantum Mechanics and the Theory of

Molecular Collisions (John Wiley and Sons, Inc., 2007), pp. 69–177.31T. J. Martnez, Acc. Chem. Res. 39, 119 (2006).32N. Makri, Annu. Rev. Phys. Chem. 50, 167 (1999).33K. Drukker, J. Comput. Phys. 153, 225 (1999).34R. Kapral, Annu. Rev. Phys. Chem. 57, 129 (2006).35R. Car and M. Parrinello, Phys. Rev. Lett. 55, 2471 (1985).36H. Meyer and W. H. Miller, J. Chem. Phys. 70, 3214 (1979).37S.-I. Sawada, A. Nitzan, and H. Metiu, Phys. Rev. B 32, 851 (1985).38J. C. Tully, J. Chem. Phys. 93, 1061 (1990).39M. Barbatti, M. Ruckenbauer, F. Plasser, J. Pittner, G. Granucci, M. Per-

sico, and H. Lischka, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 4, 26(2014).

40A. V. Akimov and O. V. Prezhdo, J. Chem. Theor. Comput. 9, 4959 (2013).41A. V. Akimov and O. V. Prezhdo, J. Am. Chem. Soc. 136, 1599 (2014).42T. Nelson, S. Fernandez-Alberti, V. Chernyak, A. E. Roitberg, and S. Tre-

tiak, J. Phys. Chem. B 115, 5402 (2011).43I. Schapiro, M. N. Ryazantsev, L. M. Frutos, N. Ferr, R. Lindh, and M.

Olivucci, J. Am. Chem. Soc. 133, 3354 (2011).44E. Tapavicza, I. Tavernelli, and U. Rothlisberger, Phys. Rev. Lett. 98,

023001 (2007).45C. F. Craig, W. R. Duncan, and O. V. Prezhdo, Phys. Rev. Lett. 95, 163001

(2005).46E. R. Bittner and P. J. Rossky, J. Chem. Phys. 103, 8130 (1995).47J. E. Subotnik and N. Shenvi, J. Chem. Phys. 134, 244114 (2011).48J. E. Subotnik, J. Phys. Chem. A 115, 12083 (2011).49H. M. Jaeger, S. Fischer, and O. V. Prezhdo, J. Chem. Phys. 137, 22A545

(2012).50J.-Y. Fang and S. Hammes-Schiffer, J. Phys. Chem. A 103, 9399 (1999).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56

Page 12: Semiclassical Monte Carlo: A first principles approach to ...serg/postscript/JCP.4900988.pdf · THE JOURNAL OF CHEMICAL PHYSICS 141, 184101 (2014) Semiclassical Monte Carlo: A first

184101-12 White et al. J. Chem. Phys. 141, 184101 (2014)

51O. V. Prezhdo and P. J. Rossky, Phys. Rev. Lett. 81, 5294 (1998).52B. J. Schwartz, E. R. Bittner, O. V. Prezhdo, and P. J. Rossky, J. Chem.

Phys. 104, 5942 (1996).53F. Webster, E. T. Wang, P. J. Rossky, and R. A. Friesner, J. Chem. Phys.

100, 4835 (1994).54M. J. Bedard-Hearn, R. E. Larsen, and B. J. Schwartz, J. Chem. Phys. 123,

234106 (2005).55Y. L. Volobuev, M. D. Hack, M. S. Topaler, and D. G. Truhlar, J. Chem.

Phys. 112, 9716 (2000).56M. D. Hack and D. G. Truhlar, J. Chem. Phys. 114, 2894 (2001).57C. Zhu, A. W. Jasper, and D. G. Truhlar, J. Chem. Phys. 120, 5543 (2004).58C. Zhu, S. Nangia, A. W. Jasper, and D. G. Truhlar, J. Chem. Phys. 121,

7658 (2004).59N. Shenvi, J. E. Subotnik, and W. Yang, J. Chem. Phys. 134, 144102 (2011).60J. E. Subotnik, J. Chem. Phys. 132, 134112 (2010).61N. Shenvi and W. Yang, J. Chem. Phys. 137, 22A528 (2012).62R. Kapral and G. Ciccotti, J. Chem. Phys. 110, 8919 (1999).63S. Nielsen, R. Kapral, and G. Ciccotti, J. Chem. Phys. 112, 6543 (2000).64C.-Y. Hsieh and R. Kapral, J. Chem. Phys. 137, 22A507 (2012).65D. Mac Kernan, G. Ciccotti, and R. Kapral, J. Phys. Chem. B 112, 424

(2008).

66W. Boucher and J. Traschen, Phys. Rev. D 37, 3522 (1988).67O. V. Prezhdo and V. V. Kisil, Phys. Rev. A 56, 162 (1997).68C. C. Martens and J.-Y. Fang, J. Chem. Phys. 106, 4918 (1997).69H. Meyer and W. H. Miller, J. Chem. Phys. 72, 2272 (1980).70G. Stock and M. Thoss, Phys. Rev. Lett. 78, 578 (1997).71V. N. Gorshkov, S. Tretiak, and D. Mozyrsky, Nat. Commun. 4, 2144

(2013).72M. Baer, Beyond Born-Oppenheimer: Electronic Nonadiabatic Coupling

Terms and Conical Intersections (Wiley-Interscience, 2006).73M. Baer, Phys. Rep. 358, 75 (2002).74P. Pechukas, Phys. Rev. 181, 166 (1969).75P. Pechukas, Phys. Rev. 181, 174 (1969).76M. F. Herman, J. Chem. Phys. 81, 754 (1984).77B. R. Landry and J. E. Subotnik, J. Chem. Phys. 135, 191101 (2011).78E. J. Heller, J. Chem. Phys. 62, 1544 (1975).79J. E. Subotnik and N. Shenvi, J. Chem. Phys. 134, 024105 (2011).80M. F. Herman, J. Phys. Chem. B 118, 8026 (2014).81N. Makri and W. H. Miller, Chem. Phys. Lett. 139, 10 (1987).82J. Doll, D. Freeman, and M. Gillan, Chem. Phys. Lett. 143, 277 (1988).83T. Nelson, S. Fernandez-Alberti, V. Chernyak, A. E. Roitberg, and S. Tre-

tiak, J. Chem. Phys. 136, 054108 (2012).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

204.121.6.216 On: Tue, 26 May 2015 05:32:56