<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by Scholar Commons - Institutional Repository of the University of South Carolina

University of South Carolina Scholar Commons

Faculty Publications Chemistry and Biochemistry, Department of

1-15-2004 Conserving Approximations to the Quantum Potential: Dynamics with Linearized Quantum Force Sophya Garashchuk University of South Carolina--Columbia, [email protected]

Vitaly A. Rassolov University of South Carolina - Columbia, [email protected]

Follow this and additional works at: https://scholarcommons.sc.edu/chem_facpub Part of the Chemistry Commons

Publication Info Published in Journal of Chemical Physics, Volume 120, Issue 3, 2004, pages 1181-1190. http://jcp.aip.org/ © 2004 by American Institute of Physics

This Article is brought to you by the Chemistry and Biochemistry, Department of at Scholar Commons. It has been accepted for inclusion in Faculty Publications by an authorized administrator of Scholar Commons. For more information, please contact [email protected]. JOURNAL OF CHEMICAL PHYSICS VOLUME 120, NUMBER 3 15 JANUARY 2004

Energy conserving approximations to the quantum potential: Dynamics with linearized quantum force Sophya Garashchuka) and Vitaly A. Rassolov Department of Chemistry & Biochemistry, University of South Carolina, Columbia, South Carolina 29208 ͑Received 29 August 2003; accepted 21 October 2003͒ Solution of the Schro¨dinger equation within the de Broglie–Bohm formulation is based on propagation of trajectories in the presence of a nonlocal quantum potential. We present a new strategy for defining approximate quantum potentials within a restricted trial function by performing the optimal fit to the log-derivatives of the density. This procedure results in the energy-conserving dynamics for a closed system. For one particular form of the trial function leading to the linear quantum force, the optimization problem is solved analytically in terms of the first and second moments of the weighted trajectory distribution. This approach gives exact time-evolution of a correlated Gaussian wave function in a locally quadratic potential. The method is computationally cheap in many dimensions, conserves total energy and satisfies the criterion on the average quantum force. Expectation values are readily found by summing over trajectory weights. Efficient extraction of the phase-dependent quantities is discussed. We illustrate the efficiency and accuracy of the linear quantum force approximation by examining a one-dimensional scattering problem and by computing the wavepacket reaction probability for the hydrogen exchange reaction and the photodissociation spectrum of ICN in two dimensions. © 2004 American Institute of Physics. ͓DOI: 10.1063/1.1633263͔

I. INTRODUCTION for systems of large mass and because trajectories can be Quantum-mechanical effects are essential for under- propagated independently of each other. These methods per- standing of a multitude of physical and chemical phenomena, formed well in various model problems in reactive scatter- such as zero-point energy, tunneling, interference and nona- ing, nonadiabatic dynamics, dynamics in clusters and spec- diabatic behavior. In particular, time-dependent quantum troscopic applications combined with the filter-diagonal- 4–9 wavepacket techniques are widely used to study molecular ization techniques. Remarkably, recent applications ex- phenomena. Traditional exact methods of solving the time- tend to biological systems that are well beyond the full 10,11 dependent Schro¨dinger equation that are based on spatial quantum-mechanical theoretical studies due to their size. grids, basis sets of functions or discrete variable For example, in Ref. 12 intramolecular proton transfer in representation1 scale exponentially with the dimensionality photoexcited 2-(2Ј-hydroxyphenyl)-oxazole was modeled N of a system. This scaling makes them inapplicable to sys- in full-dimension of 35 degrees of freedom. These calcula- tems beyond eight or so dimensions. In contrast, methods of tions demonstrated importance of coupling of the reaction molecular dynamics for which a phase space ensemble of modes to the vibrational modes and predicted absorption classical trajectories represents a density distribution, are spectrum. The drawbacks of the semiclassical initial value routinely used in studies of large systems of thousands of representation methods are the expensive stability analysis particles, such as liquids and biomolecules, though these that scales as N2, the notorious sign problem of the oscilla- methods are not capable of describing quantum effects. tory phase space integrals resulting in a very large number of Therefore, intensive research efforts go into development of trajectories,12–14 and also difficulties in assessing the semi- new approaches incorporating leading quantum effects into classical error and in systematically improving semiclassical trajectory-based methods. description toward full . The most popular techniques of this kind are the semi- A new area of active research deals with approaches classical initial value representation methods2–4 based on the based on quantum trajectories that solve the hydrodynamic stationary phase approximation (ប→0) to the Schro¨dinger or de Broglie–Bohm form of the Schro¨dinger equation.15 equation. Trajectories sampling the phase space of an initial The wave function of a system is represented in a set of wavepacket evolve in time according to classical mechanics; ‘‘particles’’ that move according to the classical equation of quantum effects are derived from the classical and motion and carry along certain density. The nonlocal from the stability of the final phase space variables with re- quantum-mechanical character enters this formulation spect to their initial values. The semiclassical initial value through the quantum potential, which depends on the wave representation methods are very appealing in the context of function density and its derivatives. The quantum potential nuclear motion, because the semiclassical limit is appropriate governs the dynamics of the particles along with the classical external potential. Since solution of the Schro¨dinger equation a͒Electronic mail: [email protected] is based on trajectories, rather than grid points, the scaling

0021-9606/2004/120(3)/1181/10/$22.001181 © 2004 American Institute of Physics

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 1182 J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 S. Garashchuk and V. A. Rassolov

bottleneck is avoided. Unlike semiclassical initial value rep- are described in Sec. II. One- and two-dimensional examples resentation methods, quantum trajectories sample the coordi- are presented in Sec. III along with the analysis of the LQF nate space. In the last few years several practical ways of method. Section IV concludes. using quantum trajectories have been suggested, such as lo- cal least-square fit, adaptive and moving grids;16–19 a meth- odology of using quantum trajectories within the Wigner rep- II. THE ENERGY CONSERVING APPROXIMATE resentation and in dissipative and nonadiabatic dynamics QUANTUM POTENTIAL have been also developed.20–22 Application to multidimen- sional model problems performed with a small number of A. The quantum trajectory formalism 23 trajectories is encouraging. However, for general problems First of all, we briefly describe the quantum trajectory accurate implementation of the hydrodynamic Schro¨dinger formalism in many dimensions emphasizing the most general equation, which is a nonlinear equation resulting in a com- features of the quantities involved. Let us consider a quan- plicated and rapidly varying quantum potential, is challeng- tum mechanical system described in N-dimensional Carte- ing. Several intricate examples of quantum trajectories can sian coordinates with the Hamiltonian Hˆ ϭPˆ †M Ϫ1Pˆ /2ϩV, be found in Refs. 24 and 25, where they serve as an inter- ϭ (n) 33 where M is a diagonal matrix of masses, ͕M nn͖ m . pretive tool. Substituting expressions Recently, we have introduced the idea of using more practical approximate quantum potentials ͑AQP͒26,27 that can ␫ ␺͑xជ,t͒ϭA͑xជ,t͒expͩ S͑xជ,t͒ͪ , A͑xជ,t͒ϭͱ␳͑xជ,t͒, ͑1͒ provide accurate description of quantum effects in semiclas- ប sical systems. With AQP derived from the global fit to the ជ ជ density we could balance numerical effort vs level of de- where amplitude A(x) and phase S(x) are real functions, ¨ scription of quantum effects by choosing the number of fit- into the time-dependent Schrodinger equation and making a ting Gaussian functions. Complete basis of fitting functions transformation into the Lagrangian frame of reference, one obtains gives exact quantum potential and full quantum-mechanical description, while zero functions gives a ‘‘classical’’ limit. dS͑xជ,t͒ pជ TM Ϫ1pជ ϭ ϪVϪU, ͑2͒ An intermediate number of functions can be considered as a dt 2 semiclassical description with a single fitting function repro- ducing exact quantum dynamics of a Gaussian wavepacket d␳͑xជ,t͒ ϭϪٌជ vជ ␳͑xជ,t͒. ͑3͒ in a locally quadratic potential. In principle, the accuracy of dt • the fit controls the degree in which semiclassical description approaches exact quantum dynamics. This was illustrated on U is a nonlocal time-dependent quantum potential a one-dimensional scattering on a barrier for which accurate ប2 ٌជ TM Ϫ1ٌជ A͑xជ,t͒ transmission probabilities were obtained. Nevertheless, the UϭϪ . ͑4͒ total energy of the system was not conserved. In this paper 2 A͑xជ,t͒ we describe a different approach to define an approximate pជ ϭMvជ ϭٌជ S(xជ,t) is a classical momentum. Equation ͑2͒ is quantum potential. This new method of the linearized quan- the Hamilton–Jacobi equation describing dynamics of a tra- tum force ͑LQF͒, first described in Ref. 28, is based on the jectory in the presence of a classical potential V and a quan- optimal fit within a trial function to the log-derivatives of the tum potential U. Position and momentum of a trajectory, xជ wave function density and it conserves total energy in a ϭxជ(t) and pជ ϭpជ (t), can be found from Hamilton’s equations closed system. We interpret the fitted function as a ‘‘nonclas- of motion. sical’’ component of the momentum operator. A simple The difference between the action of a quantum me- implementation with a linear trial function described below ˆ is rigorous, free of empirical parameters, and cheap. Optimal chanical momentum operator, P, on a wave function in co- ជ parameters of the trial function are found from the first and ordinate space and that of classical p is second moments of distribution of trajectories, whose ٌជ A͑xជ,t͒ weights are constant in time. The procedure produces a linear ͑Pˆ Ϫpជ ͒␺͑xជ,t͒ϭϪ␫ប ␺͑xជ,t͒. ͑5͒ A͑xជ,t͒ quantum force and it is formulated solely in terms of sum- mation over trajectories. This new method provides exact The prefactor Ϫ␫បٌជ A(xជ,t)/A(xជ,t) in front of ␺(xជ,t)is quantum-mechanical evolution of a Gaussian in a quadratic imaginary and on the order of ប. We interpret it as a nonclas- potential, as do the most successful semiclassical methods, sical component of Pˆ vanishing in the ប→0 limit, with pជ and it is capable of describing quantum mechanical effects being its classical component. The nonclassical momentum for general semiclassical systems. As an illustration we ob- is proportional to the slope of the amplitude or density and tained the photodissociation spectra of ICN and the probabil- this is the quantity that we will use below to approximate the ity of the wavepacket transmission for H and its isotopic 3 quantum potential substitutions in two dimensions. A multidimensional derivation of the energy conserving ٌជ A͑xជ,t͒ ٌជ ␳͑xជ,t͒ approximations to the quantum potential, details of the lin- rជ͑xជ,t͒ϭ2 ϭ . ͑6͒ A͑xជ,t͒ ␳͑xជ,t͒ earized quantum force approximation and strategies for the wavepacket analysis from weighted trajectory distributions The total energy of a system in this formulation is

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 Energy conserving approximations 1183

pជ TM Ϫ1pជ Eϭ ͵ ͩ ϩV͑xជ ͒ϩU͑xជ,t͒ͪ ␳͑xជ,t͒dxជ. ͑7͒ I(n)ϭ ͵ ͑r(n)͑xជ,t͒Ϫg(n)͑xជ,sជ͒͒2␳͑xជ,t͒dxជ 2 g(n)͑xជ,sជ͒ץ ,The quantum potential U(xជ,t), which is on the order of ប2 ϭI(n)ϩ ͵ ͩ 2 ϩg(n)͑xជ,sជ͒2 ͪ ␳͑xជ,t͒dxជ. (x(nץ can be considered as a nonclassical contribution to the ki- 0 netic energy operator, though it enters equations of motion ͑13͒ on equal footing with potential V(xជ). Atomic unit of បϭ1is (n) ជ used below. I0 abbreviates a term that does not depend on s. After Equation ͑3͒ is a for a normalizable trajectory-discretization of the initial density ␳(xជ,0), I(n) is wave function density, which is conserved in a closed system expressed as a weighted sum over trajectories g(n)͑xជ ,sជ͒ץ (n)ϭ (n)ϩ ͩ i ϩ (n)͑ជ ជ͒2 ͪ ͑ ͒ lim ␳͑xជ,t͒ϭ0 and ͵ ␳͑xជ,t͒dxជϭ1. ͑8͒ I I0 ͚ wi 2 (n) g xi ,s . 14 xץ i ͉xជ͉→ϱ ␳ ជ Initial density ␳(xជ,0) can be discretized in a set of particles The fact that neither (x,t) nor its derivatives are involved is moving along trajectories. A certain amount of density within of crucial importance for efficient global implementation, ␦ជ ϭ␦ (1) ␦ (2) ␦ (3) which in this case will be linear with respect to the number a volume element, xi(t) xi (t) xi (t) xi (t)• ..., is associated with the ith trajectory. This quantity—the weight of trajectories. The optimal values of sជ are found from minimization of of a trajectory wi—remains constant in the course of dynamics27 the combined functional I N I(n) d ϭ ͑ ͒ ͑␳͑xជ ,t͒␦xជ ͑t͒͒ϭ0or␳͑xជ ,t͒␦xជ ͑t͒ϭw . ͑9͒ I ͚ (n) , 15 dt i i i i i nϭ1 m We can use this expression to establish an important property by solving a system of K equations Iץ of the quantum potential for a closed system following from ϭ0 for kϭ1...K. ͑16͒ (s(kץ :the conservation of the total energy and equations of motion U͑xជ,t͒ The mass factors were introduced in Eq. ͑15͒ in order toץ dE ϭ ͵ ␳͑xជ,t͒dxជϭ0. ͑10͒ .t relate I to the corresponding quantum potential. Once Eqsץ dt ͑16͒ are solved, an approximate quantum potential U˜ , and its Another property of the quantum potential following from its ͗˜ ͘ (n) ជ ជ definition ͑4͒ is that it exerts zero average force average U can be expressed in terms of g (x,s) using ͒ ␳͑ជץ ជ ͒ 1 x,t͑ ץ U x,t Ϸg(n)͑xជ,sជ͒ and ͑17͒ (x(nץ F(n)ϭϪ͵ ␳͑xជ,t͒dxជϭ0. ͑11͒ ␳͑xជ,t͒ (x(nץ g(n)͑xជ,sជ͒ץ 2␳͑xជ,t͒ץ 1 These two general properties, Eqs. ͑10͒ and ͑11͒, can be used Ϸ ϩ (n)͑ជ ជ͒2 ͑ ͒ (n)2 (n) g x,s , 18 xץ xץ to assess the quality of semiclassical methods based on quan- ␳͑xជ,t͒ tum trajectories. as g(n)͑xជ,sជ͒ץ N 1 ˜ ϭϪ ͩ (n)͑ជ ជ͒2ϩ ͪ ͑ ͒ U ͚ (n) g x,s 2 (n) 19 xץ ϭ 8m B. Approximating the nonclassical momentum n 1 (n) As follows from Eq. ͑4͒, the quantum potential does not 1 I0 and ͗U˜ ͘ϭϪ ͩ IϪ͚ ͪ . ͑20͒ depend on the amplitude of a wave function, only on the 8 m(n) of the amplitude. Local interpolation of ␳(xជ,t), ␳ ជ For approximate quantum potential evaluated at the optimal evaluation of derivatives of (x,t) and subsequent determi- values of sជ, the condition of the energy conservation, Eq. nation of the quantum potential and quantum force become ͑10͒, is satisfied as can be verified using Eqs. ͑9͒ and ͑16͒ expensive and inaccurate in the region of low density. There- (I ds(kץ U˜ ds(k) 1 Kץ fore, we will look for the optimal approximation to the log- dE K ϭ ͵ ␳͑ជ ͒ ជ ϭϪ ជ ជ ͑ ͒ ͚ (k) x,t dx ͚ (k) s dtץ s dt 8 kϭ1ץ derivative of the density, r(x,t) of Eq. 6 , rather than to dt kϭ1 ␳(xជ,t). Each spacial component of rជ(xជ,t), ϭ0. ͑21͒ ␳͑xជ,t͒ץ 1 r(n)͑xជ,t͒ϭ , ͑12͒ x(n) C. Linearized quantum force approximationץ ␳͑xជ,t͒ is approximated by a function of xជ of general form, A solution of Eq. ͑16͒ will generate the energy conserv- g(n)(xជ,sជ), with time-dependent parameters sជ. Vector sជ is a ing dynamics of a closed system with any functional form of combined list of parameters of all g(n)s. In general, the di- g(n). Naturally, one is interested only in physically meaning- mensionality K of sជ is not directly related to the dimension- ful and practical forms for g(n). A simple functional form of ality N of xជ. For each dimension we define a functional, g(n) yielding analytical solutions of Eq. ͑16͒ is highly desir- which after differentiation by parts becomes able, but it is g(n) represented in terms of a complete ͑or a

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 1184 J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 S. Garashchuk and V. A. Rassolov

sufficiently large͒ basis, that will lead to accurate quantum ͑iv͒ since U˜ is based on approximation to the non- dynamics for any system. Considering that we target classical component of the momentum operator, quantum-mechanical effects in semiclassical systems and which is small in the semiclassical limit, we expect that the most successful semiclassical methods give exact that the LQF method can adequately describe domi- time-evolution of a Gaussian wave function in a quadratic nant quantum-mechanical effects in semiclassical sys- potential, we start by assuming that our g(n)s are generated tems. from a Gaussian density det͑A͒ N/2 ␳͑xជ,t͒Ϸͩ ͪ exp͑Ϫ͑xជϪxជ ͒TA͑xជϪxជ ͒͒. ͑22͒ ␲ 0 0 D. The wavepacket analysis A is a symmetric matrix of the widths parameters, and vector So far we have formulated our approach in terms of ជ x0 is the center of the Gaussian density. The corresponding g trajectories with their weights determined by the initial den- is a linear function of xជ. Note here that we do not restrict the sity. In order to propagate quantum trajectories in the pres- actual density to be a Gaussian. We simply look for the best ence of the approximate quantum potential, we do not need linear approximation of rជ(xជ,t) as outlined in Sec. II B, that is the densities or wave functions explicitly except at tϭ0. consistent with a Gaussian density and results in a linearized Computation of U˜ is formulated using just the trajectory quantum force ͑LQF͒.IfEq.͑22͒ becomes a poor approxi- weights, Eq. ͑9͒. The weights and trajectory positions are mation to the actual density, then the matrix elements of A also sufficient to compute expectation values of the and the resulting U˜ go to zero and dynamics becomes purely coordinate-dependent operators. For example, an expression classical. With g(n)s derived from Eq. ͑22͒, gជ (xជ,sជ)ϭ2A(xជ for the reaction probability of a wavepacket is Ϫជ ͑ ͒ x0), the functional 15 becomes prod ͑ ͒ϭ ͵ ˆ ␳͑ជ ͒ ជϷ ͑ ͒ P t P x,t dx ͚ wi . 26 ͑ជ͒ϭ ͵ ͑͑ជϪជ ͒T T Ϫ1 ͑ជϪជ ͒ i I s x x0 A M A x x0

Ϫ Ϫ1͒͒␳ ជ ͒ ជϩ ͑ ͒ ˆ Tr͑AM ͑x,t dx I0 . 23 Operator P projects the density onto products and can be represented as a Heaviside function in reaction coordinate. The last term, I , is independent on parameters sជ. The list of 0 The summation goes over trajectories, which are located in parameters sជ consists of N(Nϩ1)/2 elements of A and of N the product region of a potential surface at time t. Other components of xជ . Minimization of this functional gives a 0 quantities used to define LQF are the moments of the trajec- unique solution which for a normalized density is tory distribution. Using the hermiticity of the Hamiltonian, we can also directly compute a less trivial phase-sensitive ជ ϭ͗ ͘ϭ ͵ ជ␳͑ជ ͒ ជ ͑ ͒ x0 x x x,t dx, 24 quantity—the time-dependent auto-correlation function of a real wavepacket, which is the central object for calculating and introducing matrix B of the second moments bij photodissociation spectra ϭ͐ (i)Ϫ (i) ( j)Ϫ ( j) ␳ ជ ជ (x x0 )(x x0 ) (x,t)dx C͑t͒ϭ͗␺͑xជ,0͉͒␺͑xជ,t͒͘ϭ͗␺Ã͑xជ,t/2͉͒␺͑xជ,t/2͒͘ ϭ 1 Ϫ1 ͑ ͒ A 2 B . 25 ϭ ␫ ជ ͒͒ ͑ ͒ The corresponding approximate quantum potential and ͚ wi exp͑2 S͑xi ,t/2 . 27 its force have several attractive features. From the point of i view of implementation: For analysis of arbitrary wavepackets we need a more ͑ ͒ general type of integrals—a correlation function of station- i Optimization has a simple analytical solution; ␹ ជ ͑ii͒ the first and second moments of the density distribu- ary, (x,0), and evolving wavepackets. The correlation func- tion are expressed as single sums over trajectories, tion xជ ϭ͚ w xជ and b ϭ͚ w (x(i)Ϫx(i))(x( j)Ϫx( j)); 0 i i ij i i 0 0 Ã ͑iii͒ since our approach is global, the determination of C͑t͒ϭ ͵ ␹ ͑xជ,0͒␺͑xជ,t͒dxជ these two moments is the only addition to the classical trajectory propagation at each time step; Ã ϭ ͵ ␹ ͑xជ,0͒exp͑␫S͑xជ,t͒͒ͱ␳͑xជ,t͒dxជ, ͑28͒ ͑iv͒ the propagation method is linear with the number of dimensions. requires knowledge of either the volume element dxជ or of ␳ ជ 26 From the theoretical point of view: (x,t). In one-dimensional applications we found dxi from adjacent trajectories, which would be cumbersome to imple- ͑i͒ Time-evolution of a Gaussian wavepacket in a locally ment in many dimensions. Therefore, we resort to approxi- quadratic potential is solved exactly ͑including corre- mation of ␳(xជ,t) and possibly of S(xជ,t)or␺(xជ,t) in a local- lated density͒; ized region of space, where ͉␹(xជ,0)͉ evaluated at positions of ͑ii͒ the total energy is conserved; trajectories at time t is appreciable, ͉␹(xជ,0)͉Ͼ␦. Several ͑iii͒ the average quantum force is zero: Using Eqs. ͑19͒ strategies are available. One can approximate the density and ͑24͒, the average linearized quantum force is Fជ with a trial function on a subspace of a given ‘‘window’’ ϭ Ϫ1 ជϪជ ϭ ជ ␳ ជ Ϸ ជ ជ AM A͗x x0͘ 0; function h(x), (x,t) G(a,x), by minimizing

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 Energy conserving approximations 1185

III. EXAMPLES AND DISCUSSION Fϭ ͵ ͑␳͑xជ,t͒ϪG͑aជ ,xជ,t͒͒2h͑xជ ͒dxជ, ͑29͒ Our first example illustrates the LQF method in one di- mension. We apply it to describe the wavepacket dynamics with respect to parameters aជ . Computationally, the most for the Eckart barrier emphasizing the difference of this straightforward choice is h(xជ)ϭ͉␹(xជ,0)͉ in conjunction with method with the approximate quantum potential method, a polynomial form of G in xជ, which will result in a single based on the global density fitting.27 As a multidimensional ͑for a given time͒ fitting solved by a system of linear equa- illustration we apply the LQF approximation to a benchmark tions for aជ . The correlation function in terms of trajectories test for nuclear dynamics—calculation of the reaction prob- is ability for collinear H3 system, and to photodissociation of a collinear ICN, which is also a standard model problem for ͒Ϸ ␹Ã ជ ͒ ␫ ͒ ͱ ជ ជ ͒ ͑ ͒ C͑t ͚ ͑x j,0 exp͑ S j w j / G͑a,x j . 30 studies of approximate propagation methods. The purpose of j our testing is twofold: To check the numerical efficiency of LQF and to probe the quality of description it provides. We used this expression with G as the second order polyno- mial in the numerical examples. A. The Eckart barrier A more controllable, but more expensive option is to We consider scattering of a wavepacket on the Eckart approximate ␳(xជ,t) individually for each trajectory, using a barrier, VϭD coshϪ2(Zx), which serves as a simple one- narrow window function centered at a trajectory. This proce- dimensional model for the hydrogen exchange reaction. In dure scales as the number of trajectories squared, but only a mass-scaled units the Hamiltonian is Hˆ ϭpˆ 2/2ϩV. The val- ͉␹ ជ ͉Ͼ␦ subset of contributing trajectories, satisfying (xi,0) , ues of the parameters of V are Dϭ16 and Zϭ1.3624. The need to be considered. Similar to the convolution procedure initial wave function is a Gaussian wavepacket ជ ϭ Ϫ ជ of Ref. 26, for jth trajectory we chose h j(x) exp( (x Ϫxជ )T␤(xជϪxជ )). G(aជ ,xជ) was chosen to be either a constant or 2␣ 1/4 j j ␺͑ ͒ϭ ͑Ϫ␣͑ Ϫ ͒2ϩ␫ ͑ Ϫ ͒͒ͩ ͪ ͑ ͒ ជϪជ ␤ 0 exp x q0 p0 x q0 , 31 a quadratic form in (x x j). The widths parameter can be ␲ a number or a matrix related to the localization of ͉␹(xជ,0)͉. C(t) with several values of ␤ can be computed simulta- located to the left of the barrier in the asymptotic region of V neously in order to monitor convergence with respect to the with a positive initial momentum. window size. C(t) is found from Eq. ͑30͒ with the exception Earlier we considered the same problem and obtained ជ ϭ (0) energy resolved transmission probabilities using the AQP that G(x j) a j is the zero-order coefficient of the polyno- mial expansion for each trajectory. We used this formulation method, in which quantum potential and quantum force were for comparison to Eq. ͑27͒ in the ICN example. obtained by fitting the density in terms of a linear combina- 27 Yet another possibility is to approximate tion of Gaussian functions. It was shown that having full ␹Ã(xជ,0)2␺(xជ,t)2ϷG(aជ ,xជ), and to find complex integrand of quantum mechanics as the limit of the large number of C(t), ␹Ã(xជ,0)␺(xជ,t), directly. The importance of approxi- Gaussian functions and trajectories, the AQP method, in mating product of squares of the wavepackets is that once principle, could describe tunneling and interference effects ͑ ͒ again we can obtain equations on aជ in terms of trajectory the latter becoming numerically intensive . It was also weights. For a Gaussian G with complex parameters, equa- shown that the dynamics of the transmitted part of a wave- tions on aជ can be solved analytically by equating the lowest packet, including the reactant-product wavepacket correla- moments, and the resulting ͱG(aជ ,xជ) can be integrated ana- tion function and its energy spectrum could be fairly accu- lytically to obtain C(t). The square root of a complex num- rately obtained with a single Gaussian density fit. The ber should be taken such that the real and imaginary parts of underlying working equation for obtaining its parameters ͕ ͖ C(t) are continuous functions of time. As a caution, we note c,a,x0 , that a more sophisticated trial forms of G, that might be Ϫ 2 Ϫ 2 desired to check the adequacy of the Gaussian G, will result IAQPϭ ͵ ͑␳͑x͒Ϫc2e a (x x0) ͒2dx, ͑32͒ in more complicated equations for aជ that may not have ana- lytical solutions. was solved by the gradient minimization technique for each All of the methods mentioned above can be used to find time step. The AQP formulation with a single Gaussian den- the wave function at a given point, if the stationary wave- sity fit looks similar to the LQF method, which is also related packet is chosen to be a limiting form of the Dirac ␦-function to a Gaussian density via Eq. ͑22͒. The parameters of Eq. centered around this point. However, we will try to ͑22͒ solve a different minimization problem, Eq. ͑15͒, and with correlation functions, rather than wave functions them- therefore, can be different from their counterparts of Eq. selves, because the storage and analysis of correlation func- ͑32͒. Here we choose parameters of the initial wavepacket ␺ tions are cheaper. Most quantities in chemical dynamics that clearly show the difference between the LQF and a ͑energy-resolved reaction probabilities, etc.͒ can be found by single-Gaussian AQP methods. ␺ ␣ϭ ϭϪ analyzing correlation functions. Obtaining averaged quanti- For the initial parameters of taken as 6, q0 3 ϭ ␺ ties is often less demanding numerically and, for approxi- and p0 4, the classical initial energy of is half the barrier mate methods, more accurate. In case of the quantum trajec- height. If the quantum potential were set to zero, then all tory approach, calculation of a correlation function on a trajectories would remain to the left of the barrier resulting in restricted subspace is significantly cheaper. zero transmission. In the presence of the quantum potential

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 1186 J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 S. Garashchuk and V. A. Rassolov

methods begin to differ as ␺ bifurcates. The LQF position, ϭ x0 ͗x͘, coincides with the quantum-mechanical expectation of x very well. The unphysically large quantum force of the AQP method is also manifested in the total energy of the wave- packet, which deviates from its initial value as ␳ becomes non-Gaussian in the course of dynamics. The total energy is plotted on Fig. 1͑c͒: The LQF energy is conserved by con- struction; the AQP energy changes significantly once the un- physically large quantum force is present. Introduction of the cutoff helps in this respect as well as in stabilizing the AQP method. Despite this energy nonconservation, the total amounts of the transmitted density for the two approximate methods are essentially the same ͑about 10% less than the quantum-mechanical value͒, since the wavepacket transmis- sion probability is the half-space averaged and phase- independent quantity for this simple one-dimensional sys- ␺ tem. For of initial energy higher than the barrier top (p0 ϭ6 as in Ref. 27͒, optimization procedure of the two meth- ods gave very similar results, including the reactant-product wavepacket correlation functions. In the AQP method the quantum force was approaching zero for long times, and propagation was numerically stable without a cutoff, though the total energy in AQP was not conserved. The results were obtained using 199 trajectories for both methods. Computationally, the LQF calculation was 15 times faster than the AQP calculation, because the latter involves a nonlinear minimization. The convergence of the LQF param- ͑ ͒ eters with respect to the number of trajectories was better FIG. 1. Dynamics on the Eckart barrier. a LQF and AQP approximations Ϫ5 of density at tϭ0.8. ͑b͒ The width and the position of the center of the fit as than 10 . This suggests, that it might be possible to prede- functions of time. ͑c͒ Total energy of the system as a function of time. termine parameters for the quantum potential using small number of trajectories and to use these parameters to run large number of trajectories with a sampling adapted to a some of the trajectories gain energy sufficient to surmount specific problem. the barrier, and the wavepacket splits into the reflected and transmitted parts. Let us examine the wavepacket density ␳ B. Wavepacket reaction probability for collinear H3 at a particular instant of time, tϭ0.8, when this splitting ϩ The collinear hydrogen exchange reaction, HA HBHC occurs. Gaussian For the given choice of initial ␺ we obtain → ϩ HAHB HC , is a standard test in reaction dynamics. This qualitatively different approximations of ␳. Figure 1͑a͒ model system of light nuclei exhibits large quantum effects shows the density and its approximations obtained with the and provides a fairly stringent test on the quality of approxi- LQF and AQP methods. The first apparent difference is that mate dynamical methods. We find the transition probability while the LQF density covers both reflected and transmitted of a wavepacket for several values of initial parts, the AQP density tends to fit the most prominent peak for a system of three hydrogen nuclei, as well as for deute- that happens to be to the left of the barrier. This narrow rium (D3) and tritium (T3), which allows one to examine the Gaussian severely underestimates density on the tails and classical limit, Uϭ0, and the isotopic effect on the wave- leads to unphysically large quantum force resulting in the packet probability. The system is described in the Jacobi co- numerical breakdown of the calculation. This problem could ordinates of reactants where the kinetic energy is diagonal. be circumvented by using a convoluted density instead of the The Hamiltonian, coordinates and potential surface are the optimized density below a certain cutoff of the fitted density, same as in Ref. 6. which ensures that quantum force will go to zero at the tails We start by defining an initial wavepacket of the distribution. The LQF method is more sensitive to the low-amplitude region: The Gaussian width parameter 2a and 2 ␺͑0͒ϭͱ ͑␣ ␣ ͒1/2 exp͑Ϫ␣ ͑RϪR ͒2Ϫ␣ ͑rϪr ͒2 the corresponding quantum force go to zero, as the bifurca- ␲ 1 2 1 0 2 0 tion of ␺ progresses. The AQP width parameter also de- ϩ␫p ͑RϪR ͒͒, ͑33͒ creases from 12 to 0.8 as the wavepacket spreads, but after 0 0 tϭ0.7 it starts to increase, as the density optimization begins where r is the vibrational coordinate of the diatomic, the to reproduce the dominant peak forming on the left of the distance between HB and HC , and R is the translational de- barrier. The width parameter and the position of a fitting gree of freedom—the distance between HA and the diatomic. ͑ ͒ ␺͑ ͒ Gaussian are shown on Fig. 1 b . Positions x0 for the two 0 is a product of the Gaussian wavepacket in r approxi-

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 Energy conserving approximations 1187

mating the ground vibrational state with the Gaussian wave- packet in R located in the reactant region and with the mo- mentum directed toward the interaction region. Values of the parameters in atomic units ͑the unit of time is 918 a.u.͒ are ϭ ϭ ␣ ϭ ␣ ϭ ϭ Ϫ Ϫ R0 4.5, r0 1.3, 1 4, 2 9 and p0 ͓ 15, 1͔. The ជ ϭ initial positions for quantum trajectories x(0) ͕Ri ,ri͖ are chosen on a rectangular grid with spacings dRϭ0.029 and drϭ0.019. Trajectory with the small weights ⑀Ͻ10Ϫ6 are ជ ϭ not included. The initial momenta are p(0) ͕p0,0͖ and the ϭ Ϫ initial classical actions are Si(0) p0(Ri R0). In the course of dynamics the trajectories arrive either to the product Ͻ (HAHB HBHC) or back to the reactant region (HAHB Ͼ HBHC). The wavepacket reaction probability is a value of P(t) of Eq. ͑26͒ when the reactant/product separation of the wavepacket is complete and P(t) reaches a plateau. Trajec- ϭ tories were propagated up to time tmax 1.8 with the incre- ment dtϭ2.5ϫ10Ϫ3. Probabilities were obtained using 3569 trajectories for each value of p0 . The results are compared with the quantum-mechanical calculations performed with the split-operator method29 on a 256ϫ256 grid and with the classical calculations for which the quantum potential was set identically to zero, Uϭ0. The probability calculation were repeated for deuterium and for tritium using 2241 and 1481 trajectories, respectively. The wavepacket parameters for D3 and T3 systems are the same as for H3 with the ex- ϭ Ϫ Ϫ ception of the range for p0 ; p0 ͓ 2, 22͔ for D3 and p0 ϭ Ϫ Ϫ ͓ 3, 27͔ for T3 . Some details of the LQF calculation for a single value ϭϪ p0 9.5 for the H3 system are shown on Fig. 2. The wave- FIG. 2. The collinear hydrogen exchange reaction: ͑a͒ The wavepacket re- packet reaction probability as a function of time, obtained action probability, P(t), as a function of time; ͑b͒ the LQF position param- with the LQF method, is similar to the quantum result. The eters and the center of the quantum wavepacket, ͗R(t)͘ vs ͗r(t)͘, for times average position in the LQF calculation agrees well with the tϭ͓0.0,1.0͔; ͑c͒ the LQF width parameters, i.e., the diagonal matrix ele- ments aRR and arr , and the off-diagonal matrix element aRr , as a function quantum-mechanical expectation value for short times until of time. the wavepacket ‘‘turns around the corner’’ at tϭ0.35 and splits into transmitted and reflected components. This is also 8–10% and shifted by about 0.08 eV for all three isotopes, reflected in the LQF width parameters, i.e., in the elements while the overall shape of the LQF curves are similar to the of the matrix A whose elements have their second maximum quantum results. The difference between the quantum and at about the same time. The off-diagonal element, a , be- Rr LQF probabilities reduces as the quantum potential decreases comes comparable in magnitude to the diagonal elements showing large amount of correlation between the two coor- dinates. For later times the elements of A decrease and the quantum force goes to zero because the wave function be- comes diffuse in the process of scattering. The behavior of the width parameters suggests that the LQF, due to its simple analytical form, describes the zero-point energy effect for short times when the wave function is mostly localized in the reactant region of the potential, and does not account for it in the product region, something that can be corrected by a more sophisticated trial function g. The correct short time description might be still adequate for computation of aver- aged quantities in semiclassical systems. Figure 3 shows the wavepacket probabilities as a func- tion of the initial total energy. The classical probabilities are the same for all three isotopes ͑shown with a single curve͒ and are quite different from the quantum results. This differ- ence, however, tends to decrease for heavier isotopes, as it FIG. 3. Transmission P of a wavepacket for a collinear system of three should vanish in the infinite mass limit. The agreement be- nuclei for H3 ,D3 , and T3 as a function of initial total energy E, obtained with the LQF method ͑symbols͒ and quantum mechanically ͑lines͒.The tween the LQF and quantum probabilities for the three sys- dotted line represents quantum trajectory calculations in the absence of the tems is qualitative: The maximal P(E) are overestimated by quantum potential, Uϭ0, for the three isotopes.

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 1188 J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 S. Garashchuk and V. A. Rassolov

TABLE I. Parameters of the ground and excited potential surfaces and of the initial wavepacket for ICN in ¯ ¯ atomic units. RCN and RCI are the equilibrium distances for CN and CI stretches.

Ϫ Ground k ϭ0.97345 Ha 2 ¯ ϭ ϭ Ϫ2 ¯ ϭ CN 0 RCN 2.1732 a0 kCI 0.1863 Ha0 RCI 4.006 a0 Ϫ Ϫ Excited k ϭ0.6576 Ha 2 ¯ ϭ Aϭ7349.9 H aϭ3.535 a 1 CN 0 RCN 2.3295 a0 0 ␺͑ ͒ ␣ ϭ Ϫ2 ␣ ϭ Ϫ2 ␣ ϭ Ϫ2 ϭ 0 1 55.1525 a0 2 42.2053 a0 12 4.3893 a0 Ntraj 167

relative to the kinetic energy, i.e., for larger mass and higher namics. C(t) decays on the time scale of about one and a initial total energy. The discrepancy for the close to half oscillations of the CN stretch. The LQF spectrum agrees the top of the barrier ͑around 0.4 eV͒ persists for all masses, with the quantum result very well as was reported in Ref. 28. which can be explained as follows. In order to have the same The second value of the repulsion parameter, which is three Uϭ0 limit we used identical wavepackets for all three iso- times larger than its physical value, yields a predissociation topes, while the ground state depends on the mass. Com- process ͑system II in Ref. 31͒. Numerical values of the pa- pared to the ground state, our initial wavepackets were wider rameters of the surfaces and ␺͑0͒ are organized in Table I and the quadratic approximation to the potential was less and presented in atomic units. Trajectories sampling ͉␺(0)͉2 accurate for heavier isotopes. are initialized on a rectangular grid with zero initial momenta Thus far we conclude that LQF may be suitable to ana- and phase. A total of 167 trajectories whose weights exceed lyze isotope effects in semiclassical molecular systems. For ␦ϭ10Ϫ4 are propagated with a time step dtϭ1.25 a.u. for more detailed information, such as the energy-resolved reac- up to 2500 a.u. The LQF parameters are shown on Fig. 4. tion probability, one should be able to account for the zero- Figure 4͑a͒ shows the average position of the wavepacket point energy effects on the product side of the potential. and initial positions of some of the trajectories sampling Choice of the coordinate system should also play a role on ͉␺(0)͉2; the average position, plotted as ͗y͘ vs ͗x͘, illus- the LQF description of the zero-point energy effect. For ex- trates how the distance between I and center of mass of CN ample, using the Cartesian coordinates for this problem increases in the course of dissociation as the CN stretch un- would be of great interest. Also having a criterion on the dergoes three vibrations. Figure 4͑b͒ shows the matrix ele- applicability of semiclassical methods would be highly use- ments of A: a11 represents the changing width of the wave- ful for the development of a rigorous semiclassical propaga- tion method. For the quantum trajectory method with ap- proximate quantum potential, a cumulative over time global quantity, which accounts for local nonharmonicity, might serve this purpose. Such criterion may be applicable to a wide range of semiclassical methods.

C. Photodissociation cross section of ICN In the final example—the photodissociation of ICN—we compute a phase-dependent quantity in two different ways: With and without approximation to the time-dependent den- sity. We follow works of Heller30 and Coalson and Karplus,31 where collinear ICN was treated within the Beswick–Jortner model.32 A wavepacket representing an ICN molecule is ex- cited by a laser from the ground to the excited electronic state, where it dissociates into I and CN. The Hamiltonian and the Jacobi coordinate system are described in Ref. 31. ␺ ϭ ͱ␣ ␣ ␲ 1/2 Ϫ␣ An initial wave function (0) (2 1 2/ ) exp( 1(y Ϫ 2Ϫ␣ Ϫ 2ϩ ␣ Ϫ Ϫ y0) 2(x x0) 2 12(y y 0)(x x0)) is defined as the lowest eigenstate of the ground electronic surface with zero momentum. The ground state potential is composed of two harmonic potentials in CN and CI stretches. Thus, ␺͑0͒ is a correlated Gaussian wavepacket in the Jacobi coordinates lo- cated on the repulsive wall of the excited surface. The pho- todissociation cross section is computed from the Fourier transform of the wavepacket auto-correlation function C(t) ϭ͗␺(0)͉␺(t)͘

␴͑␻͒ϭ␻Rͩ ͵ C͑t͒exp͑␫␻t͒dtͪ . ͑34͒ FIG. 4. LQF parameters for ICN: ͑a͒ The average position, ͗y͘ vs ͗x͘,of The physical value of the repulsion parameter of the ex- ͑ ͒ the wavepacket and initial positions of trajectories; b width parameters a11 ͑ ͒ ͑ ͒ ͑ ͒ cited potential surface yields a rather simple dissociation dy- solid line , a22 dashed line , and a12 dot–dashed line .

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 Energy conserving approximations 1189

packet in CN mode, a22 describes the spreading in the dissociation mode. a12 , which is one-half of the coupling term, also shows oscillations and changes sign in the course of dynamics. Working with a real ␺͑0͒ has the advantage for the LQF formulation that its auto-correlation function can be easily found by simply summing over the trajectory weights with the phase factors, as given by Eq. ͑27͒. In this expression C(t) depends on the quality of the wave function at half-time t/2 on the whole space. We can compare accuracy of the LQF method to a situation when the trajectories are propa- gated for the whole time t, but C(t) is obtained from a localized region of space where ␺͑0͒ is appreciable using Eq. ͑30͒. The LQF method is exact for a Gaussian wavepacket in the presence of a locally quadratic potential. For a general potential, the quality of the LQF description will deteriorate as ␺ becomes non-Gaussian as time increases and as the wave function becomes diffuse. Comparison of the two cor- relation functions shows the interplay of shorter time dynam- ics using Eq. ͑27͒ with the analysis of a wavepacket within a localized window using Eq. ͑30͒. Applicability of Eq. ͑30͒ for C(t) clearly depends on the spacial extent of the features of ␺(t) compared to the win- dow function: They have to be wide for a polynomial to be accurate. Therefore, for an auto-correlation function we can- not use a single window function; we perform a local expan- sion around each trajectory as outlined in Sec. II D. The width matrix of the window function is a diagonal matrix with the elements ͕16a11 ,16a22͖. The two correlation func- FIG. 5. Photodissociation cross section of ICN: ͑a͒ Absolute value of the tions are shown on Fig. 5 along with the corresponding spec- auto-correlation function ͑real part on the inset͒; ͑b͒ corresponding spectra. ͑ ͒ tra and the quantum result. One can see that two approximate LQF results are shown with solid line when using Eq. 27 andwithadash when using Eq. ͑30͒. Quantum results are marked with circles. correlation functions are slightly different from each other. This can be understood as follows: During approximate dy- namics accuracy of LQF decreases with time as the wave- tion gives a propagation scheme which is linear with respect packet looses its Gaussian shape and the error in the resulting to the number of trajectories. Approximating the log- wave function is not uniform in space. From Fig. 5͑a͒ we see derivative of the density eliminates numerical instabilities that Eq. ͑27͒ describes more accurately the first recurrence, associated with low-density regions. LQF is exact for a since it comes from half the propagation time. However, its Gaussian wavepacket in a locally quadratic potential, and in agreement deteriorates exaggerating the next peak as ␺(t) addition to the energy conservation satisfies a criterion for becomes more delocalized for longer times. Equation ͑30͒ the average quantum force to be zero. LQF can be viewed as has contributions only from a region of space with nonzero a simple model of quantum dynamics. Its implementation is initial density and does not have this feature. The corre- computationally cheap and approaches pure classical trajec- sponding spectra are shown on Fig. 5͑b͒. Both spectra agree tory propagation. quite well with the quantum-mechanical result, with results Our numerical applications to the collinear hydrogen ex- obtained from Eq. ͑30͒ being slightly more accurate. change reaction and to the photodissociation of ICN suggest that LQF can accurately describe dominant quantum effects IV. CONCLUSIONS in semiclassical systems. Future multidimensional applica- We have described a general approach to construction of tions will show if LQF is a viable model for molecular dy- the energy-conserving approximations to the quantum poten- namics. More sophisticated trial functions are also of great tial, in which it is determined from the optimal fit of the interest, since they will provide a systematic way of going non-classical component of the momentum operator, re- from classical to exact quantum dynamics. Development of a stricted by a trial function. As its simplest implementation criterion on the accuracy of semiclassical methods is highly we used a linear trial function for the nonclassical momen- desirable and it is a subject of our research. tum in many dimensions, that produces linearized quantum ACKNOWLEDGMENTS force ͑LQF͒ and can be related to an approximation of den- sity in terms of a single correlated Gaussian. The equations Acknowledgment is made to the Donors of the American for optimal parameters are solved analytically in terms of the Chemical Society Petroleum Research Fund, for support of first and second moments of the trajectory distribution. For- this research. This work was also supported by the NSF/ mulation in terms of summation over the trajectory distribu- EPSCOR under Grant No. EPS-0296165.

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 1190 J. Chem. Phys., Vol. 120, No. 3, 15 January 2004 S. Garashchuk and V. A. Rassolov

1 J. C. Light and T. Carrington, Jr., Adv. Chem. Phys. 114,263͑2000͒. 17 R. E. Wyatt and E. R. Bittner, J. Chem. Phys. 113, 8898 ͑2000͒. 2 M. Herman and E. Kluk, Chem. Phys. 91,27͑1984͒. 18 D. Nerukh and J. H. Frederick, Chem. Phys. Lett. 332, 145 ͑2000͒. 3 K. G. Kay, J. Chem. Phys. 100,4377͑1994͒. 19 B. K. Dey, A. Askar, and H. Rabitz, J. Chem. Phys. 109, 8770 ͑1998͒. 4 W. H. Miller, J. Phys. Chem. A 105, 2942 ͑2001͒. 20 A. Donoso and C. C. Martens, Phys. Rev. Lett. 87, 223202 ͑2001͒. 5 M. L. Brewer, J. S. Hulme, and D. E. Manolopoulos, J. Chem. Phys. 106, 21 E. R. Bittner, J. Chem. Phys. 115,6309͑2002͒. ͑ ͒ 4832 1997 . 22 R. E. Wyatt, C. L. Lopreore, and G. Parlant, J. Chem. Phys. 114,5113 6 ͑ ͒ S. Garashchuk and D. J. Tannor, Chem. Phys. Lett. 262, 477 1996 . ͑2001͒. 7 ͑ ͒ H. B. W. X. Sun and W. H. Miller, J. Chem. Phys. 109,7064 1998 . 23 R. E. Wyatt and K. Na, Phys. Rev. E 65, 016702 ͑2002͒. 8 ͑ ͒ O. Kuhn and N. Makri, J. Phys. Chem. A 103, 9487 1999 . 24 P. R. Holland, The Quantum Theory of Motion ͑Cambridge University 9 S. M. Anderson, J. Ka, P. M. Felker, and D. Neuhauser, Chem. Phys. Lett. Press, Cambridge, 1993͒. 328,516͑2000͒. 25 A. S. Sanz, F. Borondo, and S. Miret-Artes, J. Phys. 14, 6109 ͑2002͒. 10 V. Guallar, V. S. Batista, and W. H. Miller, J. Chem. Phys. 110, 9922 26 S. Garashchuk and V. A. Rassolov, Chem. Phys. Lett. 364,562͑2002͒. ͑1999͒. 27 ͑ ͒ 11 V. Guallar, B. F. Gherman, and W. H. Miller, J. Am. Chem. Soc. 124, 3377 S. Garashchuk and V. A. Rassolov, J. Chem. Phys. 118, 2482 2003 . 28 ͑ ͒ ͑2002͒. S. Garashchuk and V. A. Rassolov, Chem. Phys. Lett. 376,358 2003 . 29 ͑ ͒ 12 V. Guallar, V. S. Batista, and W. H. Miller, J. Chem. Phys. 113, 9510 M. D. Feit, J. A. Fleck, and A. Steiger, J. Comput. Phys. 47,412 1982 . 30 ͑2000͒. R. C. Brown and E. J. Heller, J. Chem. Phys. 75, 186 ͑1981͒. 31 13 K. G. Kay, J. Chem. Phys. 101,2250͑1994͒. R. D. Coalson and M. Karplus, J. Chem. Phys. 93, 3919 ͑1990͒. 14 A. Walton and D. Manolopoulos, Chem. Phys. Lett. 244, 448 ͑1995͒. 32 J. A. Beswick and J. Jortner, Chem. Phys. 24,1͑1977͒. 15 D. Bohm, Phys. Rev. 85, 167 ͑1952͒. 33 Single superscript indexes are used to label dimensions; subscript indexes 16 C. L. Lopreore and R. E. Wyatt, Phys. Rev. Lett. 82, 5190 ͑1999͒. are used to label trajectories.

Downloaded 15 Mar 2011 to 129.252.71.114. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions