Rate constants from the reaction path Hamiltonian. I. Reactive flux simulations for dynamically correct rates

Similar documents
Semi-Classical Theory for Non-separable Systems :

Exploring the energy landscape

Chapter 8. Chemical Dynamics

Quantum mechanical transition state theory and tunneling corrections

Direct ab initio dynamics studies of N H 2^NH H reaction

Molecular Dynamics and Accelerated Molecular Dynamics

Figure 1: Transition State, Saddle Point, Reaction Pathway

A growing string method for determining transition states: Comparison to the nudged elastic band and string methods

ARTICLES. Normal-mode analysis without the Hessian: A driven molecular-dynamics approach

Energy Barriers and Rates - Transition State Theory for Physicists

Path Optimization with Application to Tunneling

D. De Fazio, T. V. Tscherbul 2, S. Cavalli 3, and V. Aquilanti 3

Mass-Related Dynamical Barriers in Triatomic Reactions

Vibrational Spectroscopy & Intramolecular Vibrational Redistribution (IVR)

A Theoretical Study of Oxidation of Phenoxy and Benzyl Radicals by HO 2

Quantum Dissipation: A Primer

Semiclassical initial value representation for the Boltzmann operator in thermal rate constants

(1 \ C_typical tunneling energy at room temperature.

Introduction to Geometry Optimization. Computational Chemistry lab 2009

Introduction to Theories of Chemical Reactions. Graduate Course Seminar Beate Flemmig FHI

Eran Rabani, S. A. Egorov, a) and B. J. Berne Department of Chemistry, Columbia University, 3000 Broadway, New York, New York 10027

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

IV. Classical Molecular Dynamics

An Introduction to Quantum Chemistry and Potential Energy Surfaces. Benjamin G. Levine

P. W. Atkins and R. S. Friedman. Molecular Quantum Mechanics THIRD EDITION

Foundations of Chemical Kinetics. Lecture 19: Unimolecular reactions in the gas phase: RRKM theory

Selected Topics in Mathematical Physics Prof. Balakrishnan Department of Physics Indian Institute of Technology, Madras

Modeling the Non-Equilibrium Behavior of Chemically Reactive Atomistic Level Systems Using Steepest-Entropy-Ascent Quantum Thermodynamics

Semiclassical molecular dynamics simulations of excited state double-proton transfer in 7-azaindole dimers

Appendix C - Persistence length 183. Consider an ideal chain with N segments each of length a, such that the contour length L c is

Advanced Physical Chemistry CHAPTER 18 ELEMENTARY CHEMICAL KINETICS

Rigid bodies - general theory

Rotations and vibrations of polyatomic molecules

Electronic structure theory: Fundamentals to frontiers. 1. Hartree-Fock theory

CHAPTER 17 RAW Quantum transition state theory G. Mills a, G. K. Schenter b, D. E. Makarov b;c and H. Jonsson a; a Department of Chemistry, Box

Ab initio molecular dynamics and nuclear quantum effects

Tools for QM studies of large systems

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

Dynamics of Lennard-Jones clusters: A characterization of the activation-relaxation technique

Semiclassical molecular dynamics simulations of intramolecular proton transfer in photoexcited 2-2 -hydroxyphenyl oxazole

Effective theory of quadratic degeneracies

Kinetic Pathways of Ion Pair Dissociation in Water

Foundations of Chemical Kinetics. Lecture 12: Transition-state theory: The thermodynamic formalism

22 Path Optimisation Methods

Tutorial on rate constants and reorganization energies

Chem 442 Review for Exam 2. Exact separation of the Hamiltonian of a hydrogenic atom into center-of-mass (3D) and relative (3D) components.

5.62 Physical Chemistry II Spring 2008

Density Functional Theory. Martin Lüders Daresbury Laboratory

Ayan Chattopadhyay Mainak Mustafi 3 rd yr Undergraduates Integrated MSc Chemistry IIT Kharagpur

# ( 1) j=1. 1 Computer Experiment 5: Computational Thermochemistry. 1.1 Background:

Example questions for Molecular modelling (Level 4) Dr. Adrian Mulholland

CHEMICAL REACTION DYNAMICS BEYOND THE BORN-OPPENHEIMER APPROXIMATION

Chemical Science EDGE ARTICLE

Molecular body vs. molecular skeleton Ideal surface 1, 29

Coherent state semiclassical initial value representation for the Boltzmann operator in thermal correlation functions

OH CH 3 D H 2 O CH 2 D and HDO CH 3, R2

Summary of free theory: one particle state: vacuum state is annihilated by all a s: then, one particle state has normalization:

d 3 r d 3 vf( r, v) = N (2) = CV C = n where n N/V is the total number of molecules per unit volume. Hence e βmv2 /2 d 3 rd 3 v (5)

Reaction class transition state theory: Hydrogen abstraction reactions by hydrogen atoms as test cases

Molecular energy levels

Rotation in liquid 4 He: Lessons from a highly simplified model

Wavepacket Correlation Function Approach for Nonadiabatic Reactions: Quasi-Jahn-Teller Model

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

Submitted to The Journal of Physical Chemistry

Communication: An accurate full fifteen dimensional permutationally invariant. Jun Li, 1,* and Hua Guo 2,* China. New Mexico 87131, United States

TECHNIQUES TO LOCATE A TRANSITION STATE

QUANTUM MECHANICAL METHODS FOR ENZYME KINETICS

Handout 10. Applications to Solids

M02M.1 Particle in a Cone

Quantum and Classical Evolution of Chemical Reaction Wave-Front by Fast Marching Method: A New Perspective of Reaction Dynamics

Charge and Energy Transfer Dynamits in Molecular Systems

Time-Dependent Transition State Theory to Determine Dividing Surfaces and Reaction Rates in Multidimensional Systems

Mixed quantum-classical dynamics. Maurizio Persico. Università di Pisa Dipartimento di Chimica e Chimica Industriale

Although we rarely see a shock wave in everyday life,

Dynamic force matching: Construction of dynamic coarse-grained models with realistic short time dynamics and accurate long time dynamics

Time-Dependent Statistical Mechanics 5. The classical atomic fluid, classical mechanics, and classical equilibrium statistical mechanics

Molecular propagation through crossings and avoided crossings of electron energy levels

Vibrationally Mediated Bond Selective Dissociative Chemisorption of HOD on Cu(111) Supporting Information

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer

Rate constants, timescales, and free energy barriers

Multi-Configuration Molecular Mechanics Based on Combined Quantum Mechanical and Molecular Mechanical Calculations

Lecture 12. Electron Transport in Molecular Wires Possible Mechanisms

Theory of electron transfer. Winterschool for Theoretical Chemistry and Spectroscopy Han-sur-Lesse, Belgium, December 2011

T6.2 Molecular Mechanics

Express the transition state equilibrium constant in terms of the partition functions of the transition state and the

Brownian Motion and Langevin Equations

Ab initio classical trajectories on the Born Oppenheimer surface: Hessian-based integrators using fifth-order polynomial and rational function fits

Seth B. Harkins and Jonas C. Peters

Detailed Balance in Ehrenfest Mixed Quantum-Classical Dynamics

Kinetics of nitric oxide desorption from carbonaceous surfaces

Thermochemistry in Gaussian

Dynamics of Nucleation in the Ising Model

Equivalence between Symmetric and Antisymmetric Stretching Modes of NH 3 in

PHYSICAL REVIEW LETTERS

3. RATE LAW AND STOICHIOMETRY

Transition Theory Abbreviated Derivation [ A - B - C] # E o. Reaction Coordinate. [ ] # æ Æ

Project: Vibration of Diatomic Molecules

Stereodynamics of the O( 3 P) with H 2 (D 2 ) (ν = 0, j = 0) reaction

Neural networks for the approximation of time-dependent dividing surfaces in Transition State Theory

Phonons and lattice dynamics

Transcription:

JOURNAL OF CHEMICAL PHYSICS VOLUME 121, NUMBER 10 8 SEPTEMBER 2004 ARTICLES Rate constants from the reaction path Hamiltonian. I. Reactive flux simulations for dynamically correct rates Baron Peters and Alexis T. Bell Department of Chemical Engineering, University of California, Berkeley, California 94720 Arup Chakraborty a) Department of Chemical Engineering, University of California, Berkeley, California 94720; Department of Chemistry, University of California, Berkeley, California 94720; Materials Sciences Division, Lawrence Berkeley National Laboratory, University of California, Berkeley, California 94720; and Physical Biosciences Division, Lawrence Berkeley National Laboratory, University of California, Berkeley, California 94720 Received 19 April 2004; accepted 9 June 2004 As ab initio electronic structure calculations become more accurate, inherent sources of error in classical transition state theory such as barrier recrossing and tunneling may become major sources of error in calculating rate constants. This paper introduces a general method for diabatically constructing the transverse eigensystem of a reaction path Hamiltonian in systems with many degenerate transverse frequencies. The diabatically constructed reaction path Hamiltonian yields smoothly varying coupling constants that, in turn, facilitate reactive flux calculations. As an example we compute the dynamically corrected rate constant for the chair to boat interconversion of cyclohexane, a system with 48 degrees of freedom and a number of degenerate frequencies. The transmission coefficients obtained from the reactive flux simulations agree with previous results that have been calculated using an empirical potential. Furthermore, the calculated rate constants agree with experimental values. Comparison to variational transition state theory shows that, despite finding the true bottleneck along the reaction pathway, variational transition state theory only accounts for half of the rate constant reduction due to recrossing trajectories. 2004 American Institute of Physics. DOI: 10.1063/1.1778161 I. INTRODUCTION There are two major challenges in calculating accurate, ab initio rate constants for polyatomic systems. The first is determining the potential energy surface PES on which the reaction occurs. A large number of electronic structure theories are available for calculating the Born-Oppenheimer potential energy as a function of the nuclear coordinates. 1 4 The first and second derivatives of the potential energy with respect to the nuclear coordinates are also available from many electronic structure theories. 2,4 Continuing improvements in electronic structure theory promise increasingly accurate energies and applications to larger systems in the future. 2 6 The second challenge is to use the relevant features of the PES to calculate a rate constant. The widely successful solution to this second challenge has been transition state theory. 7 11 In the harmonic approximation, transition state theory TST requires only the energy and frequencies at three stationary points along the reaction pathway, i.e., in the reactant and product basins and at the saddle point between these basins. 9 The success of harmonic TST at predicting rate constants in truly remarkable because it relies on such a sparse description of the PES. There are, however, many systems that are not well described by harmonic TST. The failure of TST can result because of classical recrossing which decreases the rate or because of quantum tunneling which increases the rate. 12,13 Recrossing and tunneling corrections are often neglected in calculating rate constants because these factors are usually small compared to the error caused by an inaccurate estimate of the PES, which is magnified exponentially in the rate constant calculation. However, with recent fundamental improvements in electronic structure theory, activation barriers for many reactions can now be calculated to within a few kcal/mol of those measured experimentally. 4 The new accuracy in state-of-the-art electronic structure calculations translates to TST rate constants that are often within a factor of 10 of the limiting accuracy of harmonic TST. With such accurate energies, neglect of barrier recrossing and tunneling can be the dominant sources of error in rate constant calculations. The effects of tunneling and recrossing can be quantified using an ab initio reaction path Hamiltonian RPH. 14 While such calculations have been performed, the illustrations of this approach have been limited to systems with very few atoms 15,16 or to a few chosen degrees of freedom from larger systems. 17 This paper employs an improved method for coma Author to whom correspondence should be addressed. Electronic mail: arup@uclink.berkeley.edu 0021-9606/2004/121(10)/4453/8/$22.00 4453 2004 American Institute of Physics

4454 J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Peters, Bell, and Chakraborty puting RPH coupling constants to enable efficient dynamic simulations of reactive systems with many degrees of freedom. The coupling constants that result from the new method are smooth functions of the reaction coordinate even in large systems with many degenerate transverse vibrational frequencies. This paper uses the new method for constructing an ab initio RPH to perform reactive flux simulations. The dynamic rate constants are compared to TST rate constants and to variational TST VTST rate constants. In the literature it has been noted that VTST minimizes the effects of dynamics by identifying the true bottleneck along the reaction coordinate. 18,19 However, VTST is a transition state theory, and thus it may not fully account for recrossing. The chair-boat interconversion of cyclohexane provides an interesting example because previous studies 20 22 using an empirical potential energy surface 23 have shown evidence of ballistic trajectories that recross the dividing surface by traversing the product basin and then bouncing back. These studies resulted in transmission coefficients significantly smaller than unity. Cyclohexane has 48 internal degrees of freedom and a number of degenerate vibrational frequencies, and thus provides a challenging test of our procedure for constructing a reaction path Hamiltonian. Additionally, there are a number of experimental measurements of cyclohexane isomerization kinetics. 24 26 II. DIABATIC CONSTRUCTION OF A REACTION PATH HAMILTONIAN For a reactive molecular system, the configurations and potential energy profile along the minimum energy reaction pathway MEP, along with the vibrational frequencies and eigenvectors in directions perpendicular to the MEP, can be used to construct an atomistic Hamiltonian description of the reaction pathway region on the potential energy surface. Miller, Handy, and Adams developed a RPH that includes coupling amongst the reaction coordinate, transverse or bath vibrational modes, and rotational degrees of freedom. 14 This paper treats angular momentum and internal degrees of freedom as separable contributions to the energy. Thus we use the zero angular momentum version of the RPH. 14 The potential energy in the RPH is a harmonic valley surrounding the steepest descent reaction path a(s), a curve in the 3N-dimensional space of mass weighted Cartesian coordinates that is commonly referred to as the MEP. The MEP is typically parametrized by its arclength with the convention that the saddle point is a 0. There are F 3N 6 internal degrees of freedom, F-1 of which describe position in the hyperplane perpendicular to the steepest descent path. The Fth degree of freedom, the arclength along the steepest descent path, is the reaction coordinate s. A coordinate system for directions normal to the path is established by diagonalizing the Hessian matrix at a(s) after projecting out the path tangent vector a (s). 14,27 Let K(s) be the projected Hessian at a point a(s). Diagonalization of K(s) provides an orthonormal eigenvector basis for the F-1 directions transverse to the path. The eigenvectors of K(s) are the columns of L(s) interpreted as tranverse vibrational modes. The associated FIG. 1. The top figures depict two transverse eigenvectors along the MEP black. The lower figures depict the corresponding transverse frequencies as functions of the reaction coordinate. Black and gray are used to show the correspondence between eigenvectors and frequencies along the reaction pathway. The two ordering systems differ where there is a frequency crossing. eigenvalues are the squared transverse angular frequencies 2 (s): 14,27 2 s L s T K s L s. 1 The curvature of the steepest descent pathway in the F-dimensional subspace results in curvature coupling coefficients, b s L s T a s, 2 where a prime denotes a derivative with respect to the arclength parameter s. 14,27 The curvature coupling causes the transfer of energy between the reaction coordinate and the transverse bath modes. The transverse eigenvectors in L(s) twist about the steepest descent path leading to Coriolis coupling coefficients that describe the transfer of energy amongst the bath modes, 14,27 B s L s T L s. 3 Let q measure the displacement from the path along each transverse eigenvector in L, and let the conjugate momenta to (q,s) be(p,p s ). Then the RPH for a system with zero total angular momentum is 14 H p s q T B s p 2 2 1 q T b s 1 2 2 p2 V 0 s 1 2 qt 2 s q, 4 where V 0 (s) V a(s) is the potential energy along the MEP. In practical applications, the RPH is computed from a collection of discrete points along the path a(s). 27 The diagonalization of K(s) at a point s i on the MEP yields an unordered transverse eigensystem that must be matched to the transverse eigensystem at the neighboring points on the MEP, i.e., to the eigensystems of K(s i 1 ) and K(s i 1 ). A diabatic construction where the transverse eigenvectors are matched to neighboring sets of transverse eigenvectors so as to maximize overlap is preferable to an adiabatic construction where the eigensystem at each point on the path is ordered according to ascending transverse frequency. 27 The diabatic construction avoids unphysical spikes in B(s) where transverse frequencies cross. The result is smoothly varying

J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Reaction path Hamiltonian. I 4455 FIG. 2. A diabatically ordered set of eigenvectors L 1 and L 2 and a degenerate pair of eigenvectors S 1 and S 2 at the next discretized point on the MEP, the axis passing through both planes. S 1 and S 2 are rotated by the angle in the plane to give the vector L 1 and the dotted vector. By considering changes in sign shown by reversal of the dotted vector to obtain L 2 ) the rotation angles required for an optimal matching are between /4 and /4. These rotations and sign changes minimize spurious Coriolis twisting among degenerate transverse modes. coupling coefficients, B(s) and b(s), that facilitate dynamics calculations. Figure 1 illustrates a reaction pathway with crossing transverse frequencies and the difference between a diabatic and adiabatic matching of eigenmodes. In practice, the matching is more complicated than the situation depicted in Fig. 1. The eigenvectors are determined numerically at best to within a minus sign. In the case of degenerate frequencies, rotations among all degenerate pairs of eigenvectors must also be considered. These considerations suggest the following general scheme for constructing the function L(s) from a collection of adiabatically ordered eigenvector matricies S(s k ) where k goes from one to the number of points on the discretized path a(s). Our strategy is to start with an adiabatically ordered eigensystem at the transition state L(0). We then move along the reaction pathway in both directions diabatically matching the next unordered eigensystem S to the previously diabatically ordered L. Each degenerate pair of eigenvectors in S requires optimization with respect to a mixing angle during the matching of eigenvectors in S and L. Degenerate pairs of eigenvectors are those corresponding to eigenvalues that differ by less than the accuracy of the electronic structure frequencies. Figure 2 depicts the matching process for a single pair of degenerate eigenvectors in S to the eigenvectors in L. A general procedure for matching eigenvectors of S to those of L uses the matrix A 1,..., num pairs ij i L j. 5 deg pairs k 1 R k k S Here R k ( k ) is a rotation matrix that rotates the kth degenerate pair of eigenvectors by an angle k. For a given set of angles the Hungarian algorithm 28,29 can be used on A to find the matching with the maximum total absolute overlap. This allows a number of schemes for maximizing the /2-periodic sum of matched overlap elements in the theta variables. The result is a one-to-one correspondence between eigenvectors of S and those of L with optimized mixing angles in all degenerate subspaces. The signs of the optimized eigenvectors are undetermined because the matrix A was defined using the absolute values of the overlaps. The undetermined signs are easily rectified by choosing the sign that makes the overlap positive. After diabatically ordering the transverse eigenmodes, the coupling elements can be computed using difference approximations at the discrete reaction coordinate values, s k where k goes from one to the number of points on the discretized path. Now all s-dependent functions that appear in the RPH are known at discrete values of the reaction coordinate. The discrete values are used to spline continuous versions of the s-dependent RPH functions: B(s), b(s), 2 (s), and V 0 (s). III. VARIATIONAL TRANSITION STATE THEORY In conventional TST, the reactant and product basins are separated by a hyperplane dividing surface that includes the saddle point and is perpendicular to the unstable eigenvector at the saddle point. VTST improves upon TST by optimizing the dividing surface location so that it minimizes the TST rate constant. 30,31 This optimization is equivalent to placing the dividing surface at the position along the reaction coordinate s that maximizes the free energy barrier. If k TST (s,t) is the flux through a hyperplane perpendicular to the reaction pathway at s, then the VTST rate constant is k VTST T min k TST s,t, s where k TST (s,t) is defined as k TST s,t T k BT h 6 exp G s G R k B T, 7 where G(s) is the Gibbs free energy for states on a dividing surface at a point s along the reaction pathway and G R is the Gibbs free energy for the entire reactant basin 3,9 G R V min k BT 2 ln 2k 3 BT det I min F k 1 k,min 2 2 k B T ln 1 exp k,min k B T. Here V min, I min, and k,min are the potential energy minimum, the inertia tensor at the minimum, and the kth frequency at the minimum. The factors of and (T) represent the symmetry factor and a tunneling correction. Tunneling corrections are discussed in part II of this paper, but for the cyclohexane reaction the small imaginary frequency allows use of the parabolic barrier tunneling correction. 32,33 From the reaction path Hamiltonian, the free energy from the dividing surface partition function at s is 3,34 8

4456 J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Peters, Bell, and Chakraborty G s V 0 s k BT 2 ln 2k 3 BT det I s F 1 k s 2 2 k B T ln 1 exp k s k B T. k 1 9 The term depending on the inertia tensor I(s) accounts for classical, rigid rotation. Here we have assumed that the difference between the Gibbs free energy G(s) and the Helmholtz free energy is negligible, i.e., that the molar volume is constant along the reaction pathway. Note that this assumption is not suitable for bimolecular reactions. The optimal dividing surface location, i.e., the bottleneck, may differ from the saddle point location if the pathway width changes rapidly near the transition state. The reduced rate constant at the shifted dividing surface approximately accounts for those trajectories that have enough energy to cross the barrier, but are deflected by narrowing valley walls before reaching the VTST dividing surface. IV. REACTIVE FLUX SIMULATIONS The reactive flux correlation function 35,36 has been used to quantify dynamical effects on rate constants in classical molecular dynamics simulations. 17,20 22 These studies are usually restricted to systems for which empirical models are available and valid for the whole reaction pathway. However, a recent study simulated barrier recrossing dynamics by integrating the motion of selected degrees of freedom from an ab initio RPH. 17 Here an improved construction of the RPH makes it possible to include all degrees of freedom in a reactive flux simulation. The reactive flux correlation function is defined as 35 k DYN t 1 h s ṡ 0 s 0 h s t, 10 where (s) is Dirac s delta function, and h(s) is a step function that indicates a product species by returning 1 if s 0 and 0 otherwise. The function k DYN (t) decays to zero on the reaction time scale. However, on timescales much shorter than the reaction timescale, the rate constant decays from the transition state theory rate, k TST 1 h s ṡ 0 s 0 h s 0, 11 to an apparent plateau. 35,37 The plateau value is the true, dynamically corrected rate constant. Molecular dynamics algorithms for computing the reactive flux correlation function 20,21,36,38 are easily adapted for use with the reaction path Hamiltonian. An ensemble of trajectories that cross the dividing surface are obtained by sampling initial positions and momenta from the Boltzmann distribution on the dividing surface (s 0). Each initial condition is then propagated forward and backward in time using Hamilton s equations from the RPH. In practice the reactive flux correlation function is normalized by its short time limit, so that the plateau value determines the ratio of flux that is reactive to the total flux crossing the dividing surface at time t 0. FIG. 3. Free energy at three temperatures and the energy along the MEP, V 0 (s), as functions of the reaction coordinate. The configurations shown hydrogen atoms omitted are the chair minimum, the half chair saddle point, and the twisted boat minimum. The dots in the upper left inset plot show the VTST correction that results from the leftward shift in the optimal dividing surface at each temperature. The free energy curves vanish for a portion of the boat basin where the ring-twisting transverse frequency is imaginary. The imaginary transverse frequency indicates that in the ringtwisting direction the PES looks like a ridge rather than a valley along the MEP. A. Example: Chair to boat interconversion of cyclohexane The saddle point on the PES was located using a Cerjan- Miller transition state search 39 started from the chair configuration at the HF/6-31G level. The energies and geometries of the stationary states along the MEP are shown in the Appendix. The steepest descent path was computed using the local quadratic approximation 27 at the HF/6-31G level with a stepsize of 0.25a 0 amu 1/2, where a 0 denotes a bohr radius. All electronic structure calculations were done using Q-Chem2.02. 40 The coupling coefficients were computed with the eigenvector matching algorithm of this paper. At the ends of the pathway, the RPH was smoothly joined to a harmonic half well matching the transverse frequencies at the minima and the second derivative along the path tangent at each of the minima. To obtain an accurate activation energy, the stationary points were reoptimized at the B3LYP/6-31G level. The forward and backward barrier heights were scaled to match the B3LYP/6-31G values by rescaling the function V 0 (s). 41,42 A VTST calculation for the chair-boat interconversion of cyclohexane was performed using the ab initio RPH. Figure 3 shows V 0 (s) and G(s) for several temperatures. All energies have been shifted so that the zero of energy is the chair minimum. The stationary point configurations of cyclohexane in the chair, transition state, and twist minimum are shown below the reaction coordinate axis. The inset shows the ratio k TST (s,t)/k TST (0,T) exp G(s) G(0) /k B T at each of the nonzero temperatures. The minimum value of this ratio is the variational correction to the standard transition state theory rate constant. As temperature increases, the optimal dividing surface shifts along the reaction coordinate towards the chair basin.

J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Reaction path Hamiltonian. I 4457 FIG. 4. The ring twisting frequency decreases rapidly near the saddle point. The free energy curves in Fig. 3 vanish because the ring twisting frequency becomes imaginary for an interval on the boat side of the saddle point. A band of C-H stretching frequencies above 3000 cm 1 has been omitted from the plot. The inset plot contains three frequency crossings to show that the diabatic construction of the RPH successfully tracks eigenvectors along the reaction pathway. This suggests that the pathway widens with increasing reaction coordinate near the saddle point. The transverse frequencies shown in Fig. 4 reveal that this widening of the pathway is directed along the ring twisting mode. The ring twisting frequency decreases rapidly near the saddle point indicating that the pathway widens in the direction of the transverse ring twisting mode. On the twist-boat side of the reaction pathway, the frequency of the twisting mode actually becomes imaginary. Where the transverse twisting frequency is imaginary, the MEP looks like a path down a ridge rather than a valley. The ridge is not a problem for determining the variational bottleneck along the reaction pathway as seen in Fig. 3. However, for the dynamics calculations, the imaginary transverse frequency does pose a problem. To prevent trajectories from running off to infinity along the imaginary twisting mode, its frequency was reset to 10 cm 1 at each point where the ring twisting frequency is imaginary. The dynamics calculations also require the RPH curvature and Coriolis coupling constants b and B. These are shown in Figs. 5 a and 5 b. The coupling elements are only slightly affected by filtering, 43 which indicates that they are smooth functions on the length-scale of the MEP stepsize. To obtain decorrelated initial conditions for the reactive flux simulations, a thousand Monte Carlo steps were taken between each trajectory. The autocorrelation functions for the sampling of initial conditions are shown in Fig. 6. An adaptive version of the fourth-order Runge-Kutta algorithm was used to solve the equations of motion. The time step was adjusted based on q(t) T b s(t) and the difference between k 1 and k 2 evaluations in the Runge-Kutta algorithm. 43,44 One-twelfth of the fastest transverse vibration period was used as a maximum timestep. All trajectories conserved energy to within 0.01k B T. Figure 7 shows the reaction coordinate versus time for three trajectories that were all initially crossing the barrier in the chair to boat direction. The trajectories are color coded according to the difference between the number of forward and reverse FIG. 5. a shows the absolute value of curvature coupling elements and the norm of the curvature vector along the MEP. Figure 5 b shows the norm of individual rows of the Coriolis coupling matrix. The coupling elements are forced to zero at the ends of the MEP so that the RPH smoothly joins to a harmonic end cap. The dotted curves show the effect on the norms of smoothing the individual curvature and Coriolis coupling elements using the formula x k (x k 1 2x k x k 1 )/4 where k indicates the discretized point along the MEP. The smoothing procedure has little effect on the coupling elements indicating that they are relatively smooth functions on the lengthscale of the MEP stepsize. The units of s are bohr amu 1/2 and the units of the coupling elements are in inverse units of s. barrier crossings for 1600 fs of forward and backward history. The reactive flux correlation functions are shown in Fig. 8. The drop in k(t) at 200 fs corresponds indicates a large FIG. 6. Autocorrelation functions from Monte Carlo sampling of each dividing surface degree of freedom. 0.95 10 6 Monte Carlo steps were used to generate 950 trajectories 10 per degree of freedom for each temperature. Here x n x n x where x is a degree of freedom and n is the number of Monte Carlo steps.

4458 J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Peters, Bell, and Chakraborty TABLE I. Transmission coefficients from VTST and dynamic corrections, k VTST /k TST and k DYN /k TST, respectively. The ratio k DYN /k VTST is 0.5 for all conditions. The dynamic corrections were taken as the value of the reactive flux autocorrelation function at 1000 fs. 173 K 223 K 273 K FIG. 7. Plot of reaction coordinate as a function of time for three trajectories initially (t 0) crossing the dividing surface from chair to boat. The reaction coordinate values corresponding to boat and chair minima are indicated by labeled horizontal lines, s 5.7a 0 amu 1/2 and to s 7.8a 0 amu 1/2, respectively. The black trajectory commits to the boat basin in forward time and originates from the chair basin in backward time. The gray trajectory is a non-reactive temporary crossing event. The trajectory shown by the dotted line is actually a reaction from boat to chair. k VTST /k TST C 6 H 12 0.81 0.78 0.74 C 6 HD 11 0.81 0.78 0.76 k DYN /k TST C 6 H 12 0.47 0.41 0.37 C 6 HD 11 0.42 0.42 0.35 k DYN /k VTST C 6 H 12 0.58 0.53 0.50 C 6 HD 11 0.52 0.54 0.46 number of trajectories recross on this timescale. The transmission coefficient slightly decreases with increasing temperature from 0.47 at 173 K to 0.37 at 273 K. Table I compares the VTST rate constants to the dynamically correct rate constants. The VTST rate is approximately twice as large as the true dynamically corrected rate constant for this system. The dynamic corrections in Table I are very similar to those obtained from reactive flux simulations by Kuharski et al. 20 using the Pickett-Strauss potential energy surface for cyclohexane. Although the transmission coefficients from these studies are similar, the barrier height used by Kuharski et al. was scaled to match experimental estimates. In this work the potential energy surface and dynamic rate constants were obtained without adjustable parameters except for those in the B3LYP density functional. Figure 9 shows Arrhenius plots of the TST, VTST, and dynamic rate constants with experimental measurements of the chair to boat isomerization rate. The experimental data are taken from 1 H-NMR studies of cyclohexane interconversion. 25,26 Results are shown for cyclohexane and also for cyclohexane- D 11 where the experiments of Anet and Bourn 26 cover a large temperature range. The agreement of the dynamically corrected rate constant with experiment is excellent for cyclohexane. However, below 200 K where data are available only for the D 11 isotope, the experimental results begin to agree most closely FIG. 8. The reactive flux autocorrelation function for isomerization in a cyclohexane and b cyclohexane-d 11. The autocorrelation functions drop at 200 fs because many trajectories reemerge from the boat basin after 200 fs. For cyclohexane-d 11 half of the trajectories were propagated to 1600 fs to ensure that the plateau had been reached. FIG. 9. Arrhenius plot showing the transition state theory TST, variational transition state theory VTST, and dynamic DYN rate constants. In a the computed rates are compared with experimental measurements of Anet et al. and Allerhand et al. for chair to boat isomerization of cyclohexane, and in b the computed rates are compared to the measurements of Anet et al. for chair to boat isomerization of cyclohexane-d 11.

J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Reaction path Hamiltonian. I 4459 TABLE II. Energies and geometries of the chair minimum, the saddle point, and the twist-boat minimum at the HF/6-31G and B3LYP/6-31G levels. Chair Saddle Twist-boat HF B3LYP HF B3LYP HF B3LYP Energy kcal/mol 0 0 11.92 11.20 6.46 6.11 Zero point corrected energy kcal/mol C 6 H 12 0 0 11.82 11.06 6.44 6.10 C 6 HD 11 0 0 11.74 11.00 5.87 6.07 Bond lengths Å 1 2 1.535 1.543 1.553 1.558 1.545 1.553 2 3 1.535 1.543 1.559 1.568 1.533 1.540 3 4 1.535 1.543 1.553 1.565 1.545 1.554 4 5 1.535 1.543 1.531 1.541 1.545 1.552 5 6 1.535 1.543 1.528 1.536 1.533 1.540 6 1 1.535 1.543 1.531 1.537 1.545 1.554 Bond angles deg 1 2 3 111.4 111.4 118.2 117.6 111.1 111.0 2 3 4 111.4 111.4 118.2 118.6 111.1 111.0 3 4 5 111.4 111.4 114.6 115.3 112.2 112.0 4 5 6 111.4 111.4 109.7 110.6 111.1 111.0 5 6 1 111.4 111.4 109.7 109 111.1 111.0 6 1 2 111.4 111.4 114.6 113.8 112.2 112.0 Dihedral angles deg 1 2 3 4 54.9 54.9 13.4 13.3 63.7 64.1 2 3 4 5 54.9 54.9 7.3 0.6 30.7 30.2 3 4 5 6 54.9 54.9 48.0 40.9 30.7 31.5 4 5 6 1 54.9 54.9 69.5 68.6 63.7 64.2 5 6 1 2 54.9 54.9 48.0 54.2 30.7 30.3 6 1 2 3 54.9 54.9 7.3 13.9 30.7 31.4 with the TST rates. The low-temperature discrepancy may be the result of an inadequate treatment of tunneling. Nevertheless, the dynamic rate constant is within a factor of 3 from the experimental rate over five orders of magnitude in the rate constant. This agreement suggests that the importance of barrier recrossing can be quantified using a reactive-flux simulation with an ab initio reaction path Hamiltonian even for large, weakly coupled systems that exhibit correlated recrossings on the picosecond timescale. V. CONCLUSIONS This study presentes a general diabatic matching scheme for constructing the transverse eigensystem of a reaction path Hamiltonian. The matching procedure extends the applicability of reaction path Hamiltonian studies to large systems with many degenerate transverse frequencies. The RPH coupling constants that result from the matching procedure exhibit smooth reaction coordinate dependence that facilitates reactive flux simulations. When used with ab initio methods for obtaining potential energy surfaces, reactive flux simulations based on the reaction path Hamiltonian can be applied to a variety of chemical systems. The cyclohexane example demonstrates that the reactive flux formalism can be applied to an ab initio reaction path Hamiltonian for a system with 54 degrees of freedom and several degenerate transverse vibrational frequencies. This study, based entirely on ab initio calculations, found similar transmission coefficients as obtained in reactive flux simulations using empirical models for cyclohexane. The reactive flux simulations also agree with previous simulation results that the trajectories tend to recross in a quasiperiodic fashion indicative of a weakly coupled unimolecular system. These quasiperiodic type barrier recrossings are a purely dynamic property of the system, and cannot be described by transition state theory or variational transition state theory. Additionally, the rates obtained are in agreement with experiment. ACKNOWLEDGMENTS The authors would like to thank Professor Martin Head- Gordon and his lab for providing Q-Chem 2.02. 40 We are grateful to Wan-Zhen Liang for valuable assistance with Q-Chem 2.02. The authors acknowledge useful suggestions from Professor William H. Miller, Professor Herb Strauss, Dr. Mark Brewer, and Michele Ceotto. The authors acknowledge British Petroleum for support. One of the authors B.P. acknowledges NSF for fellowship support. APPENDIX: STATIONARY POINT GEOMETRIES ALONG MEP FOR CYCLOHEXANE INTERCONVERSION The geometry along the MEP affects the coupling coefficients which in turn affect the reactive flux trajectories. The MEP was computed at the HF/6-31G level and the potential energy profile, V 0 (s), was rescaled using energies from re-

4460 J. Chem. Phys., Vol. 121, No. 10, 8 September 2004 Peters, Bell, and Chakraborty FIG. 10. The saddle point optimized at the B3LYP/6-31G level with carbon atoms labeled according to the numbering system used in Table II. optimized stationary points at the B3LYP/6-31G level. Table II shows the energies and geometries of the chair minimum, transition state, and twist-boat minimum at each of the two levels. The energies are relative to the energy of the chair minimum. The B3LYP/6-31G transition state is shown in Fig. 10 with the numbering system used in Table II. Hydrogens have been omitted for clarity. 1 A. Szabo and N. S. Ostlund, Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory Dover Publications, Mineola, NY, 1996. 2 M. Head-Gordon, J. Phys. Chem. 100, 13213 1996. 3 F. Jensen, Introduction to Computational Chemistry Wiley, Chichester, 1999. 4 T. Helgaker, P. Jorgensen, and J. Olsen, Molecular Electronic Structure Theory Wiley, Sussex, England, 2000. 5 S. Goedecker, Rev. Mod. Phys. 71, 1085 1999. 6 T. Vreven and K. Morokuma, J. Comput. Chem. 21, 1419 2000. 7 H. Eyring, J. Chem. Phys. 3, 107 1935. 8 M. G. Evans and M. Polanyi, Trans. Faraday Soc. 31, 875 1935. 9 P. Hanggi, P. Talkner, and M. Borkovec, Rev. Mod. Phys. 62, 251 1990. 10 D. G. Truhlar, W. L. Hase, and J. T. Hynes, J. Phys. Chem. 87, 2664 1983. 11 D. G. Truhlar, B. C. Garrett, and S. J. Klippenstein, J. Phys. Chem. 100, 12771 1996. 12 J. Keck, Discuss. Faraday Soc. 33, 173 1962. 13 V. A. Benderskii, V. I. Goldanskii, and D. E. Makarov, Phys. Rep. 233, 195 1993. 14 W. H. Miller, N. C. Handy, and J. E. Adams, J. Chem. Phys. 72, 99 1980. 15 K. Yamashita and W. H. Miller, J. Chem. Phys. 82, 5475 1985. 16 H. Wang and W. L. Hase, Chem. Phys. 212, 247 1996. 17 H. Hu, M. N. Kobrak, C. Xu, and S. Hammes-Schiffer, J. Phys. Chem. A 104, 8058 2000. 18 D. G. Truhlar and B. C. Garrett, Acc. Chem. Res. 13, 440 1980. 19 W. T. Duncan, R. L. Bell, and T. N. Truong, J. Comput. Chem. 19, 1039 1998. 20 R. A. Kuharski, D. Chandler, J. A. Montgomery, Jr., F. Rabii, and S. J. Singer, J. Phys. Chem. 92, 3261 1988. 21 M. Wilson and D. Chandler, Chem. Phys. 149, 11 1990. 22 D. Chandler and R. A. Kuharski, Faraday Discuss. Chem. Soc. 85, 329 1988. 23 H. M. Pickett and H. L. Strauss, J. Am. Chem. Soc. 92, 7281 1970. 24 R. K. Harris and N. Sheppard, Proc. Chem. Soc. 419 1961. 25 A. Allerhand, F. Chen, and H. S. Gutowsky, J. Chem. Phys. 42, 3040 1965. 26 F. A. L. Anet and A. J. R. Bourn, J. Am. Chem. Soc. 89, 760 1967. 27 M. Page and J. W. McIver, J. Chem. Phys. 88, 922 1988. 28 J. A. Dossey, A. D. Otto, L. E. Spence, and C. V. Eynden, Discrete Mathematics, 2nd ed. Harper Collins, New York, 1993. 29 C. Berge, Theory of Graphs Dover, Mineola, NY, 2001. 30 J. Keck, Adv. Chem. Phys. 13, 85 1967. 31 D. G. Truhlar and B. C. Garrett, Ann. Rev. Phys. Chem. 35, 159 1984. 32 E. Wigner, Z. Phys. Chem. Abt. B 19, 203 1932. 33 R. P. Bell, Trans. Faraday Soc. 55, 1 1959. 34 S. C. Smith, J. Phys. Chem. A 104, 10489 2000. 35 D. Chandler, J. Chem. Phys. 68, 2959 1978. 36 B. J. Berne, M. Borkovec, and J. E. Straub, J. Phys. Chem. 92, 3711 1988. 37 D. Chandler, Introduction to Modern Statistical Mechanics Oxford University Press, New York, 1987. 38 J. E. Straub and B. J. Berne, J. Chem. Phys. 83, 1138 1985. 39 C. J. Cerjan and W. H. Miller, J. Chem. Phys. 75, 2800 1981. 40 J. Kong et al., J. Comput. Chem. 21, 1532 2000. 41 J. C. Corchado, J. Espinoza-Garcia, W.-R. Hu, I. Rossi, and D. G. Truhlar, J. Phys. Chem. 99, 687 1995. 42 Y.-Y. Chuang and D. G. Truhlar, J. Phys. Chem. A 101, 3808 1997. 43 R. W. Hamming, Numerical Methods for Scientists and Engineers, 2nd ed. Dover, Mineola, NY, 1973. 44 J. D. Faires and R. L. Burden, Numerical Methods PWS-Kent, Boston, 1993.