<<

THE JOURNAL OF CHEMICAL PHYSICS 123, 084106 ͑2005͒

Ab initio Ehrenfest dynamics ͒ Xiaosong Lia Department of Chemistry, University of Washington, Seattle, Washington 98195 ͒ John C. Tullyb Department of Chemistry, Yale University, New Haven, Connecticut 06520 H. Bernhard Schlegel Department of Chemistry, Wayne State University, Detroit, Michigan 48202 Michael J. Frisch Gaussian, Inc., Wallingford, Connecticut 06492 ͑Received 21 April 2005; accepted 5 July 2005; published online 30 August 2005͒

We present an ab initio direct Ehrenfest dynamics scheme using a three time-step integrator. The three different time steps are implemented with nuclear velocity Verlet, nuclear-position-coupled midpoint Fock integrator, and time-dependent Hartree-Fock with a modified midpoint and unitary transformation algorithm. The computational cost of the ab initio direct Ehrenfest dynamics presented here is found to be only a factor of 2–4 larger than that of Born-Oppenheimer ͑BO͒ dynamics. As an example, we compute the vibration of the NaCl molecule and the intramolecular + torsional motion of H2CvNH2 by Ehrenfest dynamics compared with BO dynamics. For the vibration of NaCl with an initial kinetic energy of 1.16 eV, Ehrenfest dynamics converges to BO + dynamics with the same vibrational frequency. The intramolecular rotation of H2CvNH2 produces significant electronic excitation in the Ehrenfest trajectory. The amount of nonadiabaticity, suggested by the amplitude of the coherent progression of the excited and ground electronic states, is observed to be directly related to the strength of the electron-nuclear coupling. Such nonadiabaticity is seen to have a significant effect on the dynamics compared with the adiabatic approximation. © 2005 American Institute of Physics. ͓DOI: 10.1063/1.2008258͔

I. INTRODUCTION dynamics algorithm using surface hopping has been imple- mented using complete active-space self-consistent-field Classical trajectory calculations1 are often able to pro- ͑CASSCF͒ wave functions.7 When the system consists of vide greater insight into the dynamics of reactions than do many atoms, this approach, however, becomes rather de- stationary analysis of molecular properties. The Born- manding computationally because an explicit computation of Oppenheimer ͑BO͒ and extended Lagrangian ͑EL͒ adiabatic time-dependent excited-state wave functions is re- trajectories2–4are founded on the assumption that a single quired in surface hopping. electronic potential surface governs the dynamics. Such adia- The exact treatment of electronic nonadiabaticity can be batic approaches are widely used in investigations of reac- obtained in principle by solving the time-dependent tions on ground-state surfaces. A major limitation of adia- Schrödinger equation ͑TDSE͒ simultaneously for both elec- batic trajectories is that they are not applicable to reactions trons and nuclei. However, explicit numerical integration of involving nonadiabatic electronic processes, i.e., multiple the full TDSE is computationally prohibitive, except for very potential-energy surfaces. Proper incorporation of the elec- small systems. An approximation to the expansion of TDSE tronic response is crucial for describing a host of dynamical over all the excited states is the reduction to a single Slater processes, including laser-induced chemistry, dynamics at determinant, as exemplified by the time-dependent Hartree- metal or semiconductor surfaces, and electron transfer in mo- Fock ͑TDHF͒8–10 and time-dependent density-functional lecular, biological, interfacial, or electrochemical systems. theory ͑TDDFT͒.11 The TDHF or TDDFT avoids explicit The two most widely used approaches to account for nona- computations of excited states and represents the wave func- diabatic effects are the surface-hopping method with its tion as a coherent superposition state. As a result, all chemi- many variants,5,6 and the Ehrenfest method implemented cal properties such as potential, force, and electron density here. The surface-hopping approach extends the Born- are calculated as mean expectation values. The propagation Oppenheimer framework to the nonadiabatic regime by al- of classical trajectories for molecules represented by such a lowing stochastic electronic transitions subject to a time- and mean potential surface is often called Ehrenfest dynamics. momenta-dependent hopping probability. A direct ab initio Ehrenfest dynamics have been computed numerically in sev- eral applications.5,8,12,13 ͒ The integration of the Ehrenfest dynamics by a straight- a Author to whom correspondence should be addressed. Electronic mail: [email protected] forward single time-step method is not efficient because the ͒ b Electronic mail: [email protected] integration step size must be on the time scale of the elec-

0021-9606/2005/123͑8͒/084106/7/$22.50123, 084106-1 © 2005 American Institute of Physics

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-2 Li et al. J. Chem. Phys. 123, 084106 ͑2005͒ tronic motion, usually more than an order of magnitude ͑11͒ A,B are atomic indices. shorter than that of the nuclear motion. Micha and co- ͑12͒ ␮,␯ are indices. workers have developed the relax-and-drive procedure with ͑13͒ k,k+1 are time indices. different time scales to propagate the coupled electronic and nuclear motions.13–15 The relax step corresponds to an inte- In an orthonormal basis, the TDHF equation for the den- gration of the electronic TDHF equations at a fixed nuclear sity can be written as Pץ ,position. The drive step i = FP − PF. ͑2͒ tץ P = U · ⌬Ј · U†, ͑1͒⌬ ⌬ ͕͓͐ ͑ ͒ ͑ ͔͒ ͑ ͒ †͖ where Ј= F tЈ −F tk ,U·P tk ·U dtЈ, comes from the In general, the basis functions are not orthonormal, hence the effect of the motion of the nuclei on the integration of elec- overlap matrix is not the identity. This basis can always be tronic degrees of freedom. It accounts for the change of the orthonormalized by means of Löwdin or Cholesky transfor- electron density, resulting from the change of the Fock op- mation methods. The and the Fock matrix are erator, and indirectly couples the electronic degrees of free- transformed from the AO basis ͑PЈ and FЈ͒ into an orthonor- dom with nuclear motion through the Fock operator. mal basis ͑P and F͒ by a V: ab intio In this paper, we introduce an efficient direct P = V · PЈ · VT and F = V−T · FЈ · V−1. ͑3͒ Ehrenfest dynamics algorithm within the Hartree-Fock ap- proximation. The energy, gradient, and molecular properties In the Löwdin orthonormalization method, V=S1/2;inthe are computed “on the fly.” A three-time-step integrator with Cholesky method, the upper triangular V is obtained by the nuclear velocity Verlet, nuclear-position-coupled midpoint decomposition S=VTV. Fock integrator, and time-dependent Hartree-Fock with a When an atom-centered basis set is used to represent the modified midpoint and unitary transformation algorithm is density, additional terms arise because of the time depen- presented. dence of the position of the basis functions.3,13,15 These spu- rious coupling terms can be eliminated through the use of ͉␰͘ ͑ ͉͒␹͘ 3,13,15 II. METHODOLOGY traveling atom functions, =Tm r,t . The atom translation factor is given by We use the following notations and index conventions throughout this paper: i · m ␯2t ͑ ͒ ͭ e ͫ␯ ͑ ͒ ͵ ͬͮ ͑ ͒ Tm r,t = exp m t · r − dt · . 4 ͑1͒ x denotes atomic Cartesian coordinates. q 2 ͑ ͒ 2 p denotes atomic momenta. When the nuclear velocities are small, the atom translation ͑ ͒ ͑ 3 g denotes the energy gradient first derivative with re- factors will be near unity. As a first approximation, they are ͒ spect to nuclear coordinates . omitted in the present study. ͑4͒ ␹ denotes the atomic orbital ͑AO͒ basis ͑real values in this paper͒. A. Time-dependent Hartree-Fock electronic dynamics ͑5͒ ␾ denotes the . 10 ͑6͒ S denotes the overlap matrix, S␮␯ =͗␹␮ ͉␹␯͘. In a recent paper, we introduced the integration of the ͑7͒ PЈ and P denote the complex density matrices in the time-dependent Hartree-Fock theory using a modified mid- AO basis and orthonormal basis, respectively. point and unitary transformation TDHF ͑MMUT-TDHF͒ al- ͑8͒ hЈ͑x͒ and h͑x͒ denote the one-electron matrices calcu- gorithm to study the electron optical response in intense laser lated at atomic coordinates x, fields. A comparison with a direct numerical integration of TDSE indicates that TDHF is capable of realistically simu- 1 2 ZA .hЈ ͑x͒ ␹␮ͯ ٌ ͚ ͯ␹␯ lating some aspects of nonadiabatic electronic dynamics ␮ ␯ = ͳ − − ʹ, , ͉ ͉ 2 A r − rA In the MMUT-TDHF algorithm, the Fock matrix at time t is expressed in its eigenspace, in the AO basis and orthonormal basis, respectively. k †͑ ͒ ͑ ͒ ͑ ͒ ␧͑ ͒ ͑ ͒ ͑9͒ GЈ͑x,PЈ͒ and G͑x,PЈ͒ denote the two-electron matrices C tk · F tk · C tk = tk . 5 Ј calculated with density P at atomic coordinates x, ͑ ͒ A unitary transformation matrix U tk , which is written in 1 ͑ ͒ ␧͑ ͒ Ј␣ ͑ Ј͒ Ј␣ͭͫ␹ ͑ ͒␹ ͑ ͒ͯ ͯ␹ ͑ ͒␹ ͑ ͒ͬ terms of the eigenvectors C tk and eigenvalues tk of the G x,P = ͚ P ␮ 1 ␯ 1 ␭ 2 ␴ 2 ␮,␯ ␭␴ Fock matrix, is used to propagate the density matrix from ␭,␴ r12 time tk−1 to tk+1. 1 ͫ␹ ͑ ͒␹ ͑ ͒ͯ ͯ␹ ͑ ͒␹ ͑ ͒ͬͮ ͑ ͒ ͑ ͒ †͑ ͒ ͑ ͒ − ␮ 1 ␴ 1 ␭ 2 ␯ 2 Pk+1 = U tk · P tk−1 · U tk , 6 r12 1 U͑t ͒ = exp͓i ·2⌬t · F͑t ͔͒ Ј␤ͫ␹ ͑ ͒␹ ͑ ͒ͯ ͯ␹ ͑ ͒␹ ͑ ͒ͬ k e k + ͚ P␭␴ ␮ 1 ␯ 1 ␭ 2 ␴ 2 , r ͑ ͒ ͓ ⌬ ␧͑ ͔͒ †͑ ͒ ͑ ͒ ␭,␴ 12 = C tk · exp i ·2 te · tk · C tk . 7 in the AO basis and orthonormal basis, respectively. The MMUT-TDHF method takes into account linear changes ͑10͒ FЈ and F denote the complex Fock matrices, F␮Ј ␯ in the density during the time step by computing the Fock =h␮Ј ␯ +G␮Ј ␯͑x,PЈ͒, in the AO basis and orthonormal ba- matrix at the midpoint of the step. The time step for integrat- ⌬ sis, respectively. ing the MMUT-TDHF equations will be denoted as te.

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-3 Ab initio Ehrenfest dynamics J. Chem. Phys. 123, 084106 ͑2005͒

Equation ͑6͒ is exact only for a constant Fock matrix. Even if p ͑t ͒ ͑ ͒ ͑ ͒ A k+1/2 ⌬ ͑ ͒ the nuclei do not move, the Fock matrix is still time depen- xA tk+1 = xA tk + · tN, 11 MA dent since the electron density P is time dependent. Because ⌬ of this, the time step te has to be small enough so that the p ͑t ͒ = p ͑t ͒ − 1 g ͑t ͒ · ⌬t . ͑12͒ MMUT-TDHF step is accurate. However, the integrals used A k+1 A k+1/2 2 A k+1 N to construct the Fock matrix change very little on this time The mean force comes from the calculation of the first ⌬ scale and can be kept constant for several te time steps. derivative of the energy with respect to nuclear coordinates, using the TDHF-propagated wave function. Since ͓F,P͔ 0, the derivative is different than for Born-Oppenheimer B. Nuclear-position-coupled midpoint Fock dynamics.4 integrator GЈץ v dhЈ 1ץ Eץ = NN +Trͫ PЈ + ͯ ͯ PЈͬ ץ ץ ץ The Fock matrix xA xA dxA 2 xA PЈ FЈ = hЈ͑x͒ + GЈ͑x,PЈ͒͑8͒ dV dV −TrͫFЈV−1 PЈ + PЈ V−1FЈͬ. ͑13͒ is a function of the nuclear coordinates and electron density. dx dx The MMUT-TDHF steps ͓Eqs. ͑5͒–͑7͔͒ require a reconstruc- A A tion of the Fock matrix from updates of the nuclear positions The derivatives of the transformation matrix V must be com- and electron density. Considering that the change of the puted explicitly. In the case of the Löwdin orthonormaliza- nuclear coordinates is much slower than the change of the tion, electronic wave function, it is reasonable to assume station- dV 1 dSЈ ary nuclei for several MMUT-TDHF iterations before recal- = ͚ s ͩsT s ͪsT, ͑14͒ i ␴1/2 ␴1/2 i dx j j culating the integrals with the updated nuclear coordinates. dxA i,j i + j A ⌬ ⌬ We introduce a time step tNe=m te. The Fock matrix for ␴ where si and i are, respectively, the ith eigenvector and the MMUT-TDHF step is built using the integrals calculated eigenvalue of the overlap matrix SЈ; in the case of the ⌬ Ј ⌬ at the midpoint of the tNe time step, t + tNe/2, and the Cholesky transformation,4 instantaneous electron density matrix dV dVT Ј͑ ͒ Ј͓ ͑ Ј ⌬ ͔͒ Ј͓ ͑ Ј ⌬ ͒ Ј͑ Ј͔͒ ͫ −1ͬ ͫ −T ͬ F t = h x t + tNe/2 + G x t + tNe/2 ,P t . V = V dxA ␮,␯ dxA ␯,␮ ͑9͒ Ј −T dS −1 This nuclear-position-coupled midpoint Fock algorithm can = ͩV V ͪ when ␮ Ͻ ␯, dx ␮ ␯ be further understood in analogy to the nuclear velocity Ver- A , let ͑Sec. II C͒ if we assign the Fock operator and electron 1 dSЈ = ͩV−T V−1ͪ when ␮ = ␯, density the roles of nuclear momentum and position. In the 2 dxA ␮,␯ nuclear velocity Verlet method, the momentum at half time is ␮ Ͼ ␯ ͑ ͒ ͑ Ј͒ ͑ Ј ͒ = 0 when . 15 used to move nuclei from x tk to x tk+1 , while in the mid- point Fock algorithm the Fock matrix constructed with Note that the gradient is always real due to the fact that both nuclear position at half time propagates the electron density F and P are Hermitian and the basis functions are real. ͑ Ј͒ ͑ Ј ͒ from P tk to P tk+1 . The time step for two consecutive up- dates of nuclear coordinates for constructing the Fock matrix and electron-nuclear coupling will be denoted as ⌬t . The Ne D. The overall Ehrenfest integration scheme and change of nuclear position affects electronic degrees of free- accuracy dom at a much shorter time scale than do changes in the ⌬ ⌬ ⌬ nuclear gradient. It will be computationally efficient to em- The three time steps, tN , tNe, and te are associated bed several updates of nuclear coordinates for constructing with nuclear velocity Verlet ͑Sec. II C͒, nuclear-position- the Fock matrix and electron-nuclear coupling between two coupled midpoint Fock integrator ͑Sec. II B͒ and MMUT- consecutive calculations of the gradient, mainly because TDHF ͑Sec. II A͒ integrators, respectively. The idea is to computing the derivative of the two-electron term propagate the nuclei using the velocity Verlet with steps of ⌬ ץ ͒ ͑ ץ GЈ x,P / xA is about four times as expensive as computing tN assuming that the change in the gradient is small during ͑ ͒ ⌬ ⌬ ⌬ ⌬ just GЈ P . We introduce a time step tN =n tNe for the in- the step. The tN step is divided into n smaller steps tNe. terval between the recalculation of the energy gradient. The various integrals are recomputed for each step so that they can be used in the construction of the Fock matrix. Since the density changes more rapidly than do the integrals, C. Classical nuclear dynamics and energy derivative ⌬ ⌬ the tNe step is divided into m smaller steps of te. The x , can density is propagated using the MMUT-TDHF method usingץ/Eץ−= ¨The classical equations of motion, M x A A A ⌬ be solved numerically with a number of standard methods. In steps of te. As illustrated in Fig. 1, several midpoint Fock this paper, velocity Verlet16 is used to integrate the nuclear steps are embedded in one nuclear velocity Verlet step. To motion moving on a mean potential of a superposition state, drive the electron density, several MMUT-TDHF steps are carried out within one nuclear-position-coupled midpoint ͑ ͒ ͑ ͒ 1 ͑ ͒ ⌬ ͑ ͒ pA tk+1/2 = pA tk − 2 gA tk · tN, 10 Fock step.

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-4 Li et al. J. Chem. Phys. 123, 084106 ͑2005͒

The overall accuracy of the three-time-step Ehrenfest dynamics lies in the error of each individual integrator and their intrinsic connections. In Ehrenfest dynamics, nuclei are treated as classical particles and nuclear operators do not appear in the total Hamiltonian. Therefore, the integration of nuclear trajectory is an independent problem of solving the Newton’s equations. A single velocity Verlet step for classi- cal dynamics has been shown to be accurate up to the third order ͑see Ref. 18 for details͒. Nuclear terms appear in the electron Hamiltonian as instantaneous attractive potentials, resulting in the general time-dependent Hartree-Fock equa- tion for the electron wave function,

␾͑x,t͒ץ i = Fˆ ␾͑x,t͒ tץ

and

␾͑ ⌬ ͒ −i⌬tFˆ ␾͑ ͒ ͑ ͒ x,t + te = e x,t . 18

Equation ͑18͒ is a general problem of factorizing exponential operators. The midpoint Fock is equivalent to the widely used second-order three-split-operator symmetric decompo- sition of the time-dependent Schrödinger equation,19,20 where the Fock operator is split into nuclear-position-dependent and nuclear-position-independent parts. Such decomposition is second order for a single step and has been shown to approach a third-order accuracy with increasing number of ⌬ ⌬ →ϱ 19 splitting or iterations, tN =n tNe and n . In this paper, FIG. 1. Three-time-scale integration scheme for Ehrenfest dynamics. a second operator splitting is used to decompose the nuclear- position-independent Fock operator into Fˆ and ⌬Fˆ , where Fˆ It is instructive to examine the computational cost of the ⌬ ⌬ ⌬ ˆ is a constant within te and updated after te with F re- midpoint Fock integration scheme described above. The ex- sulting from the change of electron density. The modified pensive steps in Ehrenfest dynamics consist of computations midpoint TDHF is applied to integrate the second level of of the energy first derivative with respect to the nuclear po- splitting. The MMUT-TDHF is a second-order method as ͑ ͒ ͑ ͒ sition g Tg , the MMUT-TDHF step TTDHF , the two- well and approaches a third-order accuracy when m→ϱ, ͑ ͒ ͑ ͒ electron integrals T2e , the one-electron integrals T1e , and where ⌬t =m⌬t . ͑ ͒ Ne e the overlap matrix S TS . Note that TTDHF includes the cost Figure 2 illustrates numerical analyses of energy errors of forming and diagonalizing the Fock matrix. For small- to for the midpoint Fock and MMUT-TDHF integrators, respec- medium-sized systems, the CPU times are usually in the or- tively. We choose the vibrational dynamics of NaCl, starting Ͼ Ϸ Ͼ Ϸ der Tg TTDHF T2e T1e TS. The cost of Ehrenfest dy- with a converged wave function and an initial kinetic energy namics using only one time step for M integration steps of of 4.6 eV, integrated up to one velocity Verlet step for this ⌬ te can be estimated as numerical analysis. The error for the midpoint Fock algo- rithm is an accumulative error, in reference to the BO energy ͑ ͒ ͑ ͒ TEhrenfest = M TS + T1e + T2e + TTDHF + Tg . 16 at the end of one velocity Verlet step, computed at the end of a nuclear step ⌬t =0.5 fs with a constant ⌬t =0.0001 fs for By comparison, the cost of the midpoint Fock algorithm is N e integrating the TDHF. Similarly, the error for the MMUT- approximately TDHF is an accumulative energy error in a single nuclear ⌬ step ⌬t =0.5 fs, ⌬t =0.25 fs. Both integrators exhibit te N Ne T = M · ͩ · ͑T + T + T ͒ + T O͑⌬t2–3͒ growth of errors with respect to the step size. At the mid-Fock-Ehrenfest ⌬t S 1e 2e TDHF Ne step-size limit, the MMUT-TDHF still introduces about ⌬t 1.59ϫ10−4 a.u. energy error ͓Fig. 2͑b͔͒. This additional error + e · T ͪ. ͑17͒ ⌬ g tN can be corrected by using multiple midpoint Fock steps within one nuclear step. Figure 2͑a͒ shows that the midpoint The cost of the nuclear-position-dependent electron integra- Fock method corrects the energy error, approaching the com- ⌬ ⌬ tion is reduced by te / tNe, and the evaluation of the gradi- plete error correction at the step-size limit. It is clear that the ⌬ ⌬ ent is reduced by te / tN. In anticipation of applications to use of several midpoint Fock steps is able to reduce the over- larger systems, two-electron integrals are computed in a all energy error without introducing too much additional direct17 fashion in this paper. computational cost.

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-5 Ab initio Ehrenfest dynamics J. Chem. Phys. 123, 084106 ͑2005͒

FIG. 4. Comparison of total-energy conservation for various sets of time steps using the Ehrenfest dynamics.

parent from Fig. 3 that the vibrational frequencies by the Ehrenfest approach are almost the same as the BO trajectory with less than 0.1% difference. It also indicates that the vi- bration of NaCl with an initial kinetic energy of 1.16 eV is nearly adiabatic. For the ab initio Ehrenfest dynamics, the midpoint Fock method is computationally very efficient. The self-consistent convergence of the wave function at each step ͑ ͒ ͑⌬ ⌬ FIG. 2. Accuracy analyses for a midpoint Fock tN =0.5 fs, te in the BO trajectory is replaced with iterations to propagate ͒ ͑ ͒ ͑⌬ =0.0001 fs , and b modified midpoint TDHF integrators tN P according to the time-dependent Hartree-Fock theory. For =0.5 fs, ⌬t =0.25 fs͒. Ne the system considered here, integrating one direct Ehrenfest trajectory up to 300 fs with ⌬t =0.5, ⌬t =0.05, and ⌬t III. BENCHMARK AND DISCUSSION N Ne e =0.005 fs in the “direct” fashion takes only twice as many To illustrate the implementation and properties of the total iterations as converging the self-consistent field, and midpoint Fock Ehrenfest algorithm discussed above, we therefore twice of the CPU time of one BO trajectory with ⌬ have considered two examples: the vibration of NaCl and the tN =0.5 fs. + internal rotation of H2CvNH2. Electron integrals are com- Figure 4 compares the total-energy conservation for vari- ͑ ͒ ⌬ ͑ ͒ puted using the development version of the GAUSSIAN ous sets of time steps: a BO trajectory with tN =0.5 fs; b 21 ⌬ ⌬ ⌬ ͑ ͒ package. Natural population analysis ͑NPA͒ is used to ana- Ehrenfest with tN =0.5, tNe=0.05, and te =0.005 fs; c ⌬ ⌬ ⌬ ͑ ͒ lyze atomic properties on the fly. All the timing data pre- Ehrenfest with tN =0.5, tNe=0.1, and te =0.005 fs; d ⌬ ⌬ ⌬ sented here were collected on an Intel Pentium-4 3.2-GHz Ehrenfest with tN =0.5, tNe=0.05, and te =0.01 fs, and ͑ ͒ ⌬ ⌬ ⌬ single-processor workstation. e Ehrenfest with tN =0.25, tNe=0.05, and te =0.005 fs. The total energy by the Ehrenfest approach is conserved to A. The vibration of NaCl 0.023 and 0.012 kcal/mol for ͑b͒ and ͑e͒, respectively, com- ⌬ pared with 0.013 kcal/mol for the BO trajectory with tN Figure 3 illustrates the vibration of NaCl computed with =0.5 fs ͑a͒. For an energy conservation that is comparable to ͑ ͒ the direct Born-Oppenheimer BO and the midpoint Fock the BO trajectory, the Ehrenfest trajectory as ͑e͒ in Fig. 4 is Ehrenfest approaches. Calculations were carried out at the about four times as expensive as the BO trajectory. If the HF or TDHF level of theory using the 3-21G all-electron ⌬ ⌬ midpoint Fock step tNe or the MMUT-TDHF step te is too basis. The dynamics with midpoint Fock integrator are tested big as in the ͑c͒ and ͑d͒ cases, a systematic error is readily using several sets of time steps. Trajectories were started at noticeable even though the total energy is still conserved ͑ ͒ the NaCl equilibrium geometry RNa–Cl=2.4210 Å with an very well over a short interval. When an appropriate set of initial vibrational kinetic energy of 1.16 eV. It is readily ap- time steps is used as in the ͑b͒ and ͑e͒ case, the time evolu- tion of the energy conservation by the Ehrenfest dynamics is comparable to the Born-Oppenheimer dynamics, indicating an excellent time reversibility.

B + B. The rotation of H2C NH2 A 180° internal rotation about a covalent double bond involves the breaking and reformation of the relatively rigid ␲ bond. During the rotation, excited states often participate in the dynamics as a result of the strong electron-nuclear coupling. The rotational barrier of H C NH+ is formed by FIG. 3. Vibration of NaCl computed using the Born-Oppenheimer approach 2 v 2 ͑ the avoided crossing of two rotational potential curves. In and the Ehrenfest dynamics with different integration time steps Rmin ͒ =1.97 Å and Rmax =3.42 Å . many cases, such rotational potential curves may cross

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-6 Li et al. J. Chem. Phys. 123, 084106 ͑2005͒

netic energies. The natural charges of the NH2 group are plotted as a function of the torsional angle. When excited states participate in the dynamics, nuclear gradients often change more drastically than for the adiabatic case. In order to conserve the energy better, nuclear time steps have to be smaller than those without much nonadiabaticity as in Sec. III A. The total energy is conserved to better than 0.08, 0.09, and 0.15 kcal/mol in Fig. 6͑a͒,6͑b͒, and 6͑b͒, respectively, ⌬ ⌬ ⌬ with tN =0.1, tNe=0.01, and te =0.001 fs. The BO trajec- tories propagate strictly on the S0 state as required by the adiabatic BO approximation. In the avoided crossing region, the strong electron-nuclear coupling can result in an elec- tronic nonadiabatic transition so that molecules can be found in the excited state. Such a phenomenon cannot be simulated by the adiabatic BO molecular dynamics. By contrast, the FIG. 5. Adiabatic potential-energy curves of the ground and first excited Ehrenfest dynamics implicitly incorporates the electron- + states of H2CvNH2 computed using linear-response TDHF. nuclear coupling, and therefore is able to account for the nonadiabaticity in the simulation. Near the avoided crossing region, the Ehrenfest trajectories begin to deviate from the through a conical intersection. Figure 5 shows the adiabatic adiabatic behavior, apparently due to the electron-nuclear- potential-energy curves of the ground ͑S ͒ and first excited 0 coupling-induced electronic transition. The resulting coher- states ͑S ͒ from a rigid rotation; i.e., the torsional angle is the 1 ent superposition of excited and ground states leads to the only geometric variable. The ␲→␲* excited state is calcu- oscillation of charge. The strength of the nonadiabatic behav- lated using linear response TDHF.22 The ground-state ͑S ͒ 0 ior, as indicated by the amplitude of the coherent oscillation potential curve corresponds to the inhomogeneous breakage or deviation from the BO dynamics, depends systematically of the ␲ bond, leading to the H C+ −NH configuration at the 2 2 on the initial nuclear momentum, with higher nuclear kinetic top of the barrier. The potential-energy surface of the first energy resulting in a stronger nonadiabatic behavior. A more excited state ͑S ͒ comes from the homogenous bond break- 1 detailed investigation of the electronic nonadiabatic behavior age forming a H C·−·N+H diradical at the torsional angle of 2 2 will be discussed in a subsequent paper. 90°. The smallest energy gap between two curves is about 1.6 eV. + IV. CONCLUSION Figure 6 illustrates the dynamics of the H2CvNH2 in- ternal rotation about the double bond starting from the planar Ehrenfest dynamics propagates classical nuclei on a equilibrium geometry and with several different torsional ki- mean potential surface corresponding to an electronic super- position state. We have developed a three-time-scale algo- rithm with nuclear velocity Verlet, nuclear-position-coupled midpoint Fock integrator, and time-dependent Hartree-Fock using modified midpoint and unitary transformation for ab initio direct Ehrenfest dynamics. In particular, the coupled electronic and nuclear motions are propagated by the nuclear-position-coupled midpoint Fock method. The Ehrenfest method presented in this paper is compu- tationally practical and able to account for nonadiabatic elec- tronic excitations within a superposition state framework. The computational cost of Ehrenfest dynamics is only 2–4 times of that for the BO dynamics. The energy conservation and time reversibility are well maintained in the midpoint Fock Ehrenfest dynamics. The algorithm introduced here can be extended to the Ehrenfest dynamics within the time- dependent density-functional theory ͑TDDFT͒.

ACKNOWLEDGMENTS This work was supported by grants from the National Science Foundation ͑CHE0314208 and CHE0131157͒.

1 D. L. Bunker, Methods in Computational Physics ͑Academic Press, New York, 1971͒, Vol. 10, p. 287; L. M. Raff and D. L. Thompson, Theory of Chemical Reaction Dynamics ͑CRC, Boca Raton, 1985͒; Advances in + ͑ FIG. 6. Dynamics of H2CvNH2 internal rotation about the double bond Classical Trajectory Methods, edited by W. L. Hase JAI, Greenwich, with different initial torsional kinetic energies. 1992͒, Vol. 1; D. L. Thompson, in Encyclopedia of Computational Chem-

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 084106-7 Ab initio Ehrenfest dynamics J. Chem. Phys. 123, 084106 ͑2005͒

istry, edited by P. v. R. Schleyer, N. L. Allinger, T. Clark, J. Gasteiger, P. 8 K. C. Kulander, Phys. Rev. A 36, 2726 ͑1987͒; K. C. Kulander, K. R. S. A. Kollman, H. F. Schaefer III, and P. R. Schreiner ͑Wiley, Chichester, Devi, and S. E. Koonin, ibid. 25, 2968 ͑1982͒. New York, 1998͒,p.3059. 9 G. H. Chen and S. Mukamel, J. Chem. Phys. 103, 9355 ͑1995͒. 2 K. Bolton, W. L. Hase, and G. H. Peslherbe, in Modern Methods for 10 X. Li, S. M. Smith, A. N. Markevitch, D. A. Romanov, R. J. Levis, and Multidimensional Dynamics Computation in Chemistry, edited by D. L. H. B. Schlegel, Phys. Chem. Chem. Phys. 7, 233 ͑2005͒. Thompson ͑World Scientific, Singapore, 1998͒; R. Car and M. Parrinello, 11 E. Runge and E. K. U. Gross, Phys. Rev. Lett. 52,997͑1984͒;B.X.Xu Phys. Rev. Lett. 55, 2471 ͑1985͒; M. E. Tuckerman, P. J. Ungar, T. von and A. K. Rajagopal, Phys. Rev. A 31, 2682 ͑1985͒; A. K. Dhara and S. Rosenvinge, and M. L. Klein, J. Phys. Chem. 100, 12878 ͑1996͒;M.C. K. Ghosh, ibid. 35, 442 ͑1987͒; V. Chernyak and S. Mukamel, ibid. 52, Payne, M. P. Teter, D. C. Allan, T. A. Arias, and J. D. Joannopoulos, Rev. 3601 ͑1995͒; J. Chem. Phys. 112, 3572 ͑2000͒. Mod. Phys. 64, 1045 ͑1992͒; D. K. Remler and P. A. Madden, Mol. 12 S. I. Sawada, A. Nitzan, and H. Metiu, Phys. Rev. B 32,851͑1985͒;Z. Phys. 70,921͑1990͒; D. Marx and J. Hutter, edited by J. Grotendorst Kirson, R. B. Gerber, A. Nitzan, and M. A. Ratner, Surf. Sci. 137,527 ͑John von Neumann Institute for Computing, Julich, 2000͒, Vol. 1, p. ͑1984͒; D. A. Micha, J. Chem. Phys. 78, 7138 ͑1983͒; A. D. McLachlan, 301; J. M. Millam, V. Bakken, W. Chen, W. L. Hase, and H. B. Schlegel, Mol. Phys. 8,39͑1964͒; M. J. Field, J. Chem. Phys. 96, 4583 ͑1992͒;K. J. Chem. Phys. 111,3800͑1999͒. Runge and D. A. Micha, Phys. Rev. A 53, 1388 ͑1996͒;K.Runge,D.A. 3 E. Deumens, A. Diz, R. Longo, and Y. Ohrn, Rev. Mod. Phys. 66, 917 Micha, and E. Q. Feng, Int. J. Quantum Chem. Suppl., 24,781͑1990͒. ͑1994͒; J. B. Delos, ibid. 53, 287 ͑1981͒. 13 D. A. Micha, J. Phys. Chem. A 103, 7562 ͑1999͒; D. A. Micha and K. 4 H. B. Schlegel, J. M. Millam, S. S. Iyengar, G. A. Voth, A. D. Daniels, G. Runge, Phys. Rev. A 50, 322 ͑1994͒. E. Scuseria, and M. J. Frisch, J. Chem. Phys. 114, 9758 ͑2001͒. 14 D. A. Micha, Advances In Quantum Chemistry ͑Academic Press, New 5 P. V. Parandekar and J. C. Tully, J. Chem. Phys. 122 094102 ͑2005͒. York, 1999͒, Vol. 35, p. 317. 6 B. Space and D. F. Coker, J. Chem. Phys. 94, 1976 ͑1991͒; G. Parlant 15 E. Q. Feng, D. A. Micha, and K. Runge, Int. J. Quantum Chem. 40,545 and E. A. Gislason, ibid. 91, 4416 ͑1989͒; N. C. Blais, D. G. Truhlar, and ͑1991͒. C. A. Mead, J. Chem. Phys. 89,6204͑1988͒; C. W. Eaker, ibid. 87, 16 L. Verlet, Phys. Rev. 159,98͑1967͒. 4532 ͑1987͒; R. E. Cline and P. G. Wolynes, ibid. 86, 3836 ͑1987͒;J.R. 17 S. Saebo and J. Almlof, Chem. Phys. Lett. 154,83͑1989͒; J. Almlof, K. Stine and J. T. Muckerman, J. Chem. Phys. 91, 459 ͑1987͒; N. C. Blais Faegri, and K. Korsell, J. Comput. Chem. 3, 385 ͑1982͒. and D. G. Truhlar, J. Chem. Phys. 79, 1334 ͑1983͒; J. R. Stine and J. T. 18 A. K. Mazur, J. Comput. Phys. 136, 354 ͑1997͒. Muckerman, ibid. 68,185͑1978͒; 65, 3975 ͑1976͒; W. H. Miller and T. 19 Q. Sheng, IMA J. Numer. Anal. 9, 199 ͑1989͒; S. Z. Burstein and A. A. F. George, ibid. 56, 5637 ͑1972͒; J. C. Tully and R. K. Preston, ibid. 55, Mirin, J. Comput. Phys. 5, 547 ͑1970͒; A. D. Bandrauk and H. Shen, 562 ͑1971͒; J. C. Tully, ibid. 93,1061͑1990͒; J. C. Tully, in Dynamics Chem. Phys. Lett. 176, 428 ͑1991͒; A. D. Bandrauk and H. Shen, J. of Molecular Collisions, edited by W. H. Miller ͑Plenum, New York, Chem. Phys. 99, 1185 ͑1993͒. 1976͒, Vol. 2, p. 217. 20 K. C. Kulander, Phys. Rev. A 38, 778 ͑1988͒; M. Suzuki, J. Math. Phys. 7 A. Sanchez-Galvez, P. Hunt, M. A. Robb, M. Olivucci, T. Vreven, and H. 26, 601 ͑1985͒. 21 B. Schlegel, J. Am. Chem. Soc. 122, 2911 ͑2000͒; T. Vreven, F. Bernardi, M. J. Frisch, G. W. Trucks, H. B. Schlegel et al., GAUSSIAN, Development M. Garavelli, M. Olivucci, M. A. Robb, and H. B. Schlegel, ibid. 119, Version ͑Wallingford, CT, 2004͒. 12687 ͑1997͒; G. A. Worth, P. Hunt, and M. A. Robb, J. Phys. Chem. A 22 R. E. Stratmann, G. E. Scuseria, and M. J. Frisch, J. Chem. Phys. 109, 107,621͑2003͒. 8218 ͑1998͒.

Downloaded 02 Sep 2005 to 141.217.26.92. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp