Sharafeddin 1997

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

<— —<

Mixing Quantum-Classical Molecular


Dynamics Methods Applied to
Intramolecular Proton Transfer
in Acetylacetone

OMAR A. SHARAFEDDIN,1 KONRAD HINSEN,1, 2 TUCKER


CARRINGTON, JR.,1 BENOIT ˆ ROUX 1, 2
1
´
Departement de chimie, Universite´ de Montreal,
´ C.P. 6128, succ. Centre-ville H3C 3J7, Canada
2
Groupe de Recherche en Transport Membranaire (GRTM), Departement
´ de physique, Universite´ de
´
Montreal, C.P. 6128, succ. Centre-ville H3C 3J7, Canada

Received 23 January 1997; accepted 29 April 1997

ABSTRACT: We present results of mixed quantum-classical molecular


dynamics simulations of the intramolecular proton transfer in acetylacetone.
Simulations are performed starting from the reactant and transition state
configurations with initial velocities at each configuration chosen from an
ensemble at 300 K. The proton motion is treated quantum mechanically and the
remaining degrees of freedom are treated classically. Two mixed quantum-
classical molecular dynamics methods are implemented. In the first, a
quantum-classical time-dependent self-consistent field method ŽQCrTDSCF.,
¨
the time-dependent Schrodinger equation for the proton is solved using the split
operator approach and a plane-wave basis. In the second, a mixed quantum-
classical adiabatic method ŽQCrA., the instantaneous ground state wave
¨
function is calculated by solving the time-independent Schrodinger equation for
the configurations of the classical particles by propagating in imaginary time
using the split operator approach and the same plane-wave basis. A comparison
of the two approaches with classical trajectories is presented. The QCrTDSCF
and QCrA results are very similar for trajectories started from the reactant
configuration. The two methods, however, yield somewhat different results
when the trajectories are started from the transition state configuration. The
proton wave function of the QCrA method adjusts instantaneously to the
position of the classical particles, whereas the motion of the QCrTDSCF
wavepacket more faithfully represents the true proton dynamics. Q 1997 John
Wiley & Sons, Inc. J Comput Chem 18: 1760]1772, 1997
Correspondence to: B. Roux
´ .
Contractrgrant sponsor: FCAR ŽQuebec
Contractrgrant sponsor: NSERC ŽCanada.

Journal of Computational Chemistry, Vol. 18, No. 14, 1760]1772 (1997)


Q 1997 John Wiley & Sons, Inc. CCC 0192-8651 / 97 / 141760-13
QUANTUM-CLASSICAL MD METHODS

Keywords: computer simulations; wavepacket; zero-point vibration; activation


energy; reaction coordinate; empirical valence bond; Fourier transform

amenable to quantum-computational methods;


Introduction however, this necessitates imposing an approxi-
mate, and necessarily somewhat unrealistic, form
on the potential. To avoid imposing an approxi-

D ue to their practical importance and special


properties proton transfer reactions have
been studied intensively.1 ] 3 Symmetric intra-
mate form on the potential, and nonetheless in-
corporate quantum effects, one can either use
Feynman path integral simulations, or mixed
molecular proton transfer reactions of the type quantum-classical methods. In recent years, dis-
ŽA]H—A. l ŽA—H]A. are of particular interest. cretized Feynman path integral simulations have
It is much simpler to consider a proton transfer allowed the calculation of statistical equilibrium
process that occurs in a single molecule because properties for a variety of large complex quantum-
one does not have to deal with the complications molecular systems.14, 15 For example, the path inte-
associated with the diffusion of multiple donors gral centroid quantum transition state theory sug-
and acceptors in the liquid phase. From a theoreti- gested by Gillan16 and by Voth et al.17, 18 can be
cal point of view, proton transfer processes are used to characterize the activation free energy for
proton transfer.19 ] 23 Warshel proposed related ap-
especially interesting and challenging because
proximations that may also be useful.24, 25 How-
quantum effects, such as zero point motion and
ever, despite recent progress, this powerful com-
tunneling, are expected to be important due to the
putational approach is not easily extended to cal-
light mass of the hydrogen nucleus. As quantum-
culate time-dependent dynamical properties.26 A
mechanical calculations are difficult, many theoret-
reasonable approach to study dynamical proper-
ical studies of proton transfer reactions have been
ties of a proton transfer process is to combine
based on idealized models.4 ] 6 In a simplified view,
quantum and classical mechanics to correctly in-
proton transfer is often pictured as a one-dimen- corporate the important quantum character of the
sional process taking place in a double-well poten- dynamics of the proton while describing the heav-
tial. Sometimes, important insights can be gleaned ier atoms classically. In this study, we therefore
even from such one-dimensional treatments.7, 8 An apply and compare mixed quantum-classical com-
effective barrier height, for example, can be ex- putational schemes using a realistic potential to
tracted from the splitting of the lowest two energy study an intramolecular proton transfer process.
levels. Nevertheless, theoretical studies based on We study the intramolecular proton transfer of
such simplified low dimensional models do not the enol form of acetylacetone, whose structure is
allow direct comparisons with experimental data. given in Figure 1. The proton Ha can be bound to
Furthermore, because the motion of the proton is either of the two oxygen atoms and the molecule
often strongly coupled to the A—A distance, treat- can undergo interconversion between these forms.
ments which neglect this coupling are qualitatively
in error.9 It is therefore important to take account
of the multidimensional character of the proton-
transfer problem.
Unfortunately, it is extremely difficult to treat
multidimensional systems quantum mechanically.
For a general potential, the cost of the most effi-
cient quantum-mechanical computational methods
increases exponentially with the number of de-
grees of freedom. It is possible to treat six degrees
of freedom accurately, but it is very hard to handle
larger systems.10 ] 13 If the form of the potential
¨
is specifically chosen to facilitate solving the Schro-
dinger equation Žtime-dependent or time-inde- FIGURE 1. The reactant state geometry of
pendent., proton transfer problems become more acetylacetone.

JOURNAL OF COMPUTATIONAL CHEMISTRY 1761


SHARAFEDDIN ET AL.

The potential function is constructed on the basis tempts to incorporate the influence of nonadia-
of the empirical valence bond ŽEVB. approach of badicity in an average way by generating a single
Warshel.24, 27, 28 The fact that the reactant and prod- mean trajectory. In such cases, the QCrTDSCF
uct configurations are mirror images of each other results should be more accurate because the QCrA
simplifies the construction of a model potential equations may be obtained from the QCrTDSCF
significantly. A more popular molecule for studies equations in the adiabatic limit. The approxima-
of intramolecular proton transfer is malonalde- tion may break down in some cases, however. For
hyde,9, 29 ] 33 which differs from acetylacetone by example, if the proton wavepacket splits and sepa-
the absence of the two methyl groups. Malonalde- rates to two distant regions of space, the QCr
hyde is simpler, but it is not stable in common TDSCF may yield unphysical results, because the
solvents, whereas acetylacetone is stable in a wide classical nuclei are feeling average forces that do
range of polar and nonpolar solvents. The acety- not correspond to any well-defined state. Thus, the
lacetone molecule has 15 atoms and a total of 45 QCr TDSCF method corresponds to an approxi-
degrees of freedom. mation of unknown validity for long propagation
In this article we examine the dynamics of the times. Stochastic algorithms have been proposed
central proton of acetylacetone using two mixed and used to incorporate nonadiabatic effects by
quantum-classical ŽQC. dynamical methods, a
allowing sudden transitions to occur between dif-
quantum-classical time-dependent self-consistent-
ferent adiabatic states Ž‘‘surface hopping’’..47 ] 50
field ŽQCrTDSCF. 34 ] 43 method and a quantum-
Nevertheless, for short times, the QCr TDSCF
classical adiabatic method ŽQCrA. 44 ] 46 . To use the
method remains a useful approximation to exam-
QCrTDSCF method one solves a time-dependent
ine the dynamics of rapid events such as the bar-
¨
Schrodinger equation for the 3 degrees of freedom
rier crossing at the transition state during a proton
of the proton with a potential obtained from the
transfer process. In addition, a comparison of
full potential by evaluating it at a classical config-
QCrA and QCr TDSCF can be used as a diagnosis
uration describing the position of all the other
to examine the importance of nonadiabatic effects.
nuclei. The trajectory of the other nuclei in the
molecule is determined from classical mechanics In particular, such a comparison in the context of a
using the expectation value of the forces obtained complex molecule described with a realistic poten-
by averaging over the wavepacket associated with tial has never been done.
the proton. The system is propagated by integrat- Even though they are conceptually simple, the
ing simultaneously the coupled quantum and clas- QCrA and QCrTDSCF mixed methods have been
sical equations with discrete time-steps. In the applied mostly to simplified and highly idealized
classical limit, Newton’s equations of motion are proton transfer model systems Žfor QCrTDSCF,
recovered by the QC-TDSCF equation. QCrTDSCF see Refs. 40]43; for QCrA, see Refs. 44]46.. It is a
can be equivalently formulated in terms of the purpose of the present article to examine the mul-
pure-state density matrix evolution ŽDME. ap- tidimensional dynamics of a proton in the context
proach.42 One obtains the QCrA method if, in- of a complex molecule described with a realistic
stead of solving a time-dependent Schrodinger ¨ potential. Mixed quantum-classical simulations are
equation for the proton, one assumes that the costly, but will be used more and more frequently
proton wave function at time t is the instanta- as accurate potentials become available. Often,
neous ground state eigenfunction of the proton such methods are developed and tested on simple
Hamiltonian at the classical configuration, RŽ t .. model systems Že.g., the hydrogen motion is re-
Both mixed dynamical methods are approxima- stricted to one dimension. and only a small num-
tions to a fully quantum propagation. The QCrA ber of excited states are included Že.g., two to
method gives results which are reliable in a well- four..40 ] 46, 48 ] 50 It is therefore worthwhile explor-
defined limit. If this limit is realized for the dy- ing the relative merits of the two methods for a
namics of a given system, then the mixed-state realistic potential with a discrete grid representa-
propagation underlying the QCrTDSCF method tion for the wave function that is able to describe a
should yield results that agree with those from very large number of excited states. In the next
QCrA. However, when there is strong nonad- section, we present the details of the methods and
iabatic mixing between the quantum states the the potential energy surface used in our calcula-
QCrA method is non adequate. QCrTDSCF at- tion.

1762 VOL. 18, NO. 14


QUANTUM-CLASSICAL MD METHODS

where m is the mass of the proton and V Žr, RŽ t .. is


Theory and Method the potential energy. The position of the ith nu-
cleus obeys the classical equation of motion:
MIXED QUANTUM-CLASSICAL
PROPAGATION METHODS ¨ iŽt. s U
Mi R Hd rc 3 Ž r, t . Fi Ž r, R Ž t .. c Ž r, t . Ž 3 .
In this section, we present the mixed classical-
quantum methods we use. As several groups have where Mi is the mass of the nucleus and Fi Žr, RŽ t ..
used both the QCrTDSCF and the QCrA meth- s y­ V Žr, R.r­ R i , evaluated with the classical
ods, we merely state the equations and briefly nuclei at their position RŽ t ..
discuss the underlying assumptions.34 ] 41, 44 ] 46 We have also implemented a mixed quantum-
The fundamental approximation inherent in any classical method based on an adiabatic approxima-
TDSCF method is one of separability of the time- tion ŽQCrA..44 ] 46 One can imagine at least two
dependent wave function. To derive the QCr ways of deriving the QCrA equations. One way is
TDSCF equations one begins by assuming separa- to start from the QCrTDSCF and make an adia-
bility of the degrees of freedom to be treated batic approximation to solve the time-dependent
classically and the degrees of freedom to be treated ¨
Schrodinger equation w eq. Ž1.x . For this purpose,
quantum mechanically. In our case, the quantum one considers the expansion of c Žr, t .:
degrees of freedom are r ' Ž x, y, z ., the three
X X
Cartesian components of the proton position, and
Ý c n Ž t . fn Žr; R Ž t .. eyi H « Ž t . d t r "
t
c Ž r, t . s 0 n Ž4.
the classical degrees of freedom are R ' Ž X 1 , Y1 , n
Z1 , X 2 , Y2 , Z2 , . . . ., the Cartesian coordinates of the
other atoms in the system. It is assumed that in terms of the instantaneous eigenfunctions
c Žr, R, t ., the time-dependent wave function of the fnŽr, RŽ t .. of the proton Hamiltonian Hp :
entire system, can be approximated by the product
c Žr, t . x ŽR, t .. Substituting this product ansatz into Hp fn Ž r, R Ž t .. s « n Ž R Ž t .. fn Ž r, R Ž t .. Ž5.
¨
the Schrodinger equation yields two time-depen-
dent self-consistent Schrodinger ¨ equations for the If the system is adiabatic so that:
two parts of the total system, each of which con-
tains an effective potential determined by an aver- 1
² fn < ­ t V < fm : g w « n Ž R Ž t .. y « m Ž R Ž t ..x 2 Ž 6 .
age over the time-dependent wave function of the "
other equation. The mixed quantum-classical ver-
sion of the TDSCF can be obtained by assuming then Žsee Ref. 51. ­ t c nŽ t . ; 0, and if the wavepacket
that the wave function x ŽR, t . is localized about is f 0 Žr; RŽ0.. at time t s 0 then c nŽ t . f c nŽ t s 0. s
the expectation value of R. If x ŽR, t . were per- dn0 . According to the adiabatic approximation the
fectly localized, the expectation values ²R i Ž t .: proton wave function adapts instantaneously to
would evolve according to classical equations of the movement of the classical atoms, and is always
motion. Henceforth, we shall not distinguish be- equal to the ground state of the proton potential
tween the expectation value of the position of the determined by the position of the classical atoms
ith classical nucleus ²R i Ž t .: and R i Ž t ., the corre- at time t. In this case, the classical equation for the
sponding classical position. motion of the heavier atoms becomes:
According to the QCrTDSCF prescription, the
proton wave function, c Žr, t ., obeys the time-de- ¨ iŽt. s U
pendent Schrodinger¨ equation34 ] 36 :
Mi R H dr f 0 r;
Ž R Ž t .. Fi Ž r, R Ž t .. f 0 Ž r, R Ž t ..

­« 0 Ž R Ž t ..
i" ­ t c Ž r, t . s Hp c Ž r, t . Ž1. sy Ž7.
­ Ri

for the proton Hamiltonian Hp : where the last equality follows from the Hell-
mann]Feynman theorem.
"2 Another way to derive the QCrA equations,
Hp s y = 2 q V Ž r, R Ž t .. Ž2. which offers a different perspective and does not
2m
directly invoke the TDSCF approximation, is to

JOURNAL OF COMPUTATIONAL CHEMISTRY 1763


SHARAFEDDIN ET AL.

¨
represent the time-dependent Schrodinger equa- ¨
Schrodinger w eq. Ž1.x and c Žr, t . is a function of
tion for the entire system: time, whereas the quantum part of the QCrA
method implicates a time-independent Schrodi- ¨
w Tc q Hp x C Ž r, R, t . s i" ­ tC Ž r, R, t . Ž8. nger equation w eq, Ž5.x and the f 0 Žr; RŽ t .. depends
parametrically on time. In both cases, the forces on
where Tc is the kinetic energy operator of the the classical particles are computed from an aver-
classical particles, in the fnŽr; R. basis; age over the proton wave function w using c Žr, t . in
QCrTDSCF and f 0 Žr; RŽ t . in QCrAx . According
Ý ² fn Žr; R. w Tc q Hp x fm Ž r; R .: to eq. Ž6., if the initial quantum wavepacket for the
m
QCrTDSCF propagation is chosen to be the ground
=² fm Ž r; R . ¬ C Ž r, R, t .: state f 0 Žr; RŽ0.., the solution of the time-inde-
s i" ­ t² fn Ž r; R . C Ž r, R, t .: Ž9. ¨
pendent Schrodinger equation w eq. Ž5.x , then the
QCrTDSCF and QCrA methods should give very
If one neglects nonadiabatic coupling, ² fm Žr; similar results as long as the energy difference
R.< Tc < fnŽr; R.:, this becomes Žfor the ground state.: between the ground and the lowest excited states
of the proton is large enough to prevent significant
w Tc q « 0 Ž R .x u 0 Ž R, t . s i" ­ t u 0 Ž R, t . Ž 10 . population of the latter. The QCrTDSCF allows
the proton to occupy any combination of eigen-
where: state fnŽr; RŽ t .., whereas the QCrA restricts the
proton wave function to f 0 Žr; RŽ t ... The QCr
u 0 Ž R, t . s ² f 0 Ž r; R . ¬ C Ž r, R, t .: Ž 11. TDSCF method incorporates nonadiabatic effects
not present in the QCrA method. Nonadiabatic
In effect, the R nuclei move on an effective poten- coupling effects could also be included in a mixed
tial, « 0 ŽR., corresponding to the configuration-de- quantum-classical method by allowing transitions
pendent proton ground state energy. In the classi- between the different adiabatic eigenstates fn us-
cal limit, eq. Ž10. is replaced by eq. Ž7.. ing a surface-hopping model.48 ] 50 In Tully’s
Clearly, one obtains the same QCrA equations ‘‘fewest switch surface hopping algorithm,’’ 48, 49
using either the first or the second route. In the
the full coherence of the wave function c Žr, t . is
first derivation, one begins from the QCrTDSCF
retained, as in eq. Ž1., while the classical nuclei
result Ži.e., assuming separability of the classical
move according to forces determined from the nth
and quantum degrees of freedom., takes the classi-
adiabatic state fnŽr; RŽ t ... A Monte Carlo process
cal limit, and finally makes an adiabatic approxi-
is used to introduce abrupt transitions between the
mation. In the second derivation, one begins from
various adiabatic states and the dynamics has a
¨
the time-dependent Schrodinger equation for the
stochastic character. In QCrTDSCF, the classical
entire system, makes an adiabatic approximation,
and finally makes a classical approximation. In nuclei move via eq. Ž3. according to the average
particular, the product ansatz is never explicitly forces and the result is a single mean trajectory.
invoked in the derivation. The first derivation
shows that the QCrTDSCF and the QCrA meth- POTENTIAL ENERGY SURFACE
ods should give very similar results if the adia-
batic approximation is good. If one did not wish to To study the proton transfer of acetylacetone we
approximate the motion of the heavier nuclei with require an accurate potential energy surface which
classical mechanics, one could still derive an adia- allows the proton to move from one oxygen atom
batic approximation by following the second to the other. On the other hand, because the poten-
derivation. This would not be possible with the tial is evaluated many times, it should be simple.
first derivation, because it is built upon the The most rigorous potential energy surfaces are
QCrTDSCF equations, which include the classical obtained directly from ab initio calculations, but it
approximation. is computationally prohibitive to use ab initio po-
Although the QCrTDSCF and the QCrA ap- tentials in mixed quantum-classical molecular dy-
proximations are different, the implementation of namics, which requires the repeated evaluation of
the two mixed quantum-classical methods is, in the potential energy V Žr, R. and forces Fi Žr, R. at a
practice, very similar. The quantum part of the large number of points in r space to permit evalua-
QCrTDSCF method implicates a time-dependent tion of the integrals in eqs. Ž3. and Ž7.. To devise a

1764 VOL. 18, NO. 14


QUANTUM-CLASSICAL MD METHODS

practical model, an empirical potential energy sur- deformation from the reactant geometry; and Ž5.
face was constructed and its parameters were fit- the potential energy surface for the transferring
ted to reproduce key results of ab initio calcula- proton with all other atoms fixed in the reactant or
tions. The model is described in detail in Ref. 52. in the transition state geometries.
The model uses the empirical valence bond Details about the construction of this potential
ŽEVB. approach to Warshel 24, 27, 28 to combine a energy model and a list of the values of all param-
potential energy function for the reactant, V1 , and eters can be found in Ref. 52.
a potential energy function for the product, V2 ,
into a total potential energy surface that allows a CALCULATIONS
transition between these two geometries. The EVB
potential is defined as the lower eigenvalue of an To implement the QCrTDSCF method we solve
effective ‘‘Hamiltonian’’ of a two-state system eq. Ž1. by applying the split operator technique
whose diagonal elements are V1 and V2 and whose w 55x :
off-diagonal element e is constant:
tq
c Ž r, t q tq . s exp yiV Ž r, R Ž t ..
V1 q V2 2"
V Ž V1 , V2 , e . s tq "
2
= exp qi =2
1r2 2m
V1 q V2 2
tq
y ž / y V1V2 q e 2
= exp yiV Ž r, R Ž t .. c Ž r, t .
2 2"
Ž 12. Ž 13.

The EVB potential V is a multidimensional dou- where tq is the Žquantum. time step and V Žr,
ble-well potential with reactant and product wells RŽ t .. is the potential, which is a function of r for a
which are similar Žbut not identical. to the wells of given classical configuration RŽ t . at time t. The
V1 and V2 and a barrier between the wells whose wavepacket is represented on a three-dimensional
height is controlled by the coupling parameter e , Cartesian grid. The potential energy operator is
the value of which must be found as part of a applied directly in real space. The kinetic energy
fitting procedure. The reactant and product poten- operator is applied by transforming to a basis of
tials V1 and V2 consist mostly of standard bond, products of plane waves using a three-dimensional
angle, dihedral, and nonbonded potentials; we use fast Fourier transform Ž3D-FFT..10, 56
the CHARMM potential,53 in which the bond and To implement the QCrA method, we calculate
angle potentials are harmonic, the dihedral poten- f 0 Žr; RŽ t .. using a relaxation procedure by propa-
tials consist of a single cosine term, and the non- gating the Schrodinger¨ equation in imaginary
bonded interactions contain both a Coulomb and a time.57 Again, the propagation is done with the
Lennard]Jones contribution. To reproduce the sig- split operator technique: at each classical configu-
nificant anharmonicity observed in the ab initio ration RŽ t ., we repeatedly use eq. Ž13. with an
potential energy surface, the O—H bonds were imaginary time step, tq ª itq , to relax to the
described by a Lennard-Jones type potential. All ground state, maintaining the normalization of the
other bonds were represented by harmonic func- wave function.
tions. For both methods, the classical degrees of free-
The parameters of the empirical model were dom were propagated with a classical time step tc
fitted to key features of an ab initio potential using a variant of the Verlet algorithm58 :
energy surface with points calculated at the HFr
4-31GU and HFr6-31GU levels of theory using R i Ž t q tc . s 2R i Ž t . y R i Ž t y tc .
GAUSSIAN-90.54 The features used as fitting crite-
tc2 4
ria are: Ž1. the reactant and product geometries; Ž2. q Fi Ž t . q O Ž tc . Ž 14.
the energy barrier between the reactant Žor prod- Mi
uct. geometry and transition state geometry; Ž3.
the electrostatic potential surface around the as implemented in the molecular dynamics pack-
molecule; Ž4. the energy as a function of torsional age CHARMM.53 This algorithm was also used for

JOURNAL OF COMPUTATIONAL CHEMISTRY 1765


SHARAFEDDIN ET AL.

the purely classical simulations that we performed est proton energy well for the fixed configuration
for comparison. of the classical nuclei Žreactant or transition state.
The QCrTDSCF simulations were started with and then using the imaginary time relaxation
an initial wavepacket c Žr, t s 0. and an initial method 57 to relax to the ground state. When the
classical configuration RŽ t s 0.. The QCrA simu- classical atoms are fixed at their transition state
lations were started with an initial classical config- configuration positions the proton potential has a
uration RŽ t s 0.. The simulations were performed single well and, therefore, the initial wave function
by executing the following steps: is localized. The initial velocities of the classical
atoms were taken from a Boltzmann distribution
(i) Evaluate the potential V Žr, RŽ t .. and the at a temperature of 300 K and chosen such that the
center-of-mass velocity of the total system was
forces Fi Žr, RŽ t .. at all points r for the clas-
sical configuration RŽ t .. zero. The simulations were performed for a total
time of 0.5 ps. A time step of 1 fs was used in the
(ii) Calculate the average forces ² Fi : on the numerical integration of the classical equations.
classical particles using either c Žr, t . ŽQCr Because the wavepacket propagation requires a
TDSCF. or f 0 Žr; RŽ t .. ŽQCrA.. smaller time step than the classical trajectory cal-
(iii) Generate a classical dynamical step with culation, we used a smaller time step in eq. Ž13.
eq. Ž14. to compute the next classical con- than in eq. Ž14.: tc s 12tq . That is, for each classical
figuration RŽ t q tc . using the average integration time step, we performed 12 quantum
forces ² Fi :. time steps. The same quantum time step was used
(iv) Using the potential V Žr, RŽ t .., compute the to propagate the QCrTDSCF time-dependent
next wave function, c Žr, t q tc ., with the ¨
Schrodinger equation for the proton and to solve
split operator technique with real-time ¨
the time-dependent Schrodinger equation for the
propagation ŽQCrTDSCF. or the next eigenstates of the proton.
ground state f 0 Žr; RŽ t q tc . by relaxation The wave functions were represented on a grid
using the split operator technique with extending from y1.5 A ˚ to q1.5 A ˚ with the center
imaginary time propagation ŽQCrA.. at the minimum of the proton potential corre-
sponding to the initial classical configurations. The
(v) With the new classical configuration RŽ t q
number of grid points was increased until conver-
tc . and the new c Žr, t q tc . ŽQCr TDSCF.,
gence was obtained with 32 points along each
or the new f 0 Žr; RŽ t q tc . ŽQCr A. at time
proton degree of freedom. There were a total of
t q tc , repeat the above steps.
32 = 32 = 32 s 32,768 plane wave basis functions
or grid points. To verify that the density of grid
Two initial configurations of the classical atoms points was high enough, we increased the size of
were considered, the reactant and transition state ˚ keeping the same number
the grid from 3 to 3.5 A,
configurations. Classical, QCrTDSCF, and QCrA of grid points, and found that the average bond
trajectories were generated for both initial configu- lengths Žsee Figs. 2 and 3. did not change by more
rations for a total of six trajectories. The reactant than 2% and that the average kinetic energy Žsee
configuration is the absolute minimum of the po- Fig. 5. did not change by more than 10%. Because
tential energy surface for this molecule; the transi- increasing the size of the grid from 3 to 3.5 A ˚
tion state configuration is the saddle point be- caused a drastic change in the density of points
tween the reactant and product configuration. The our results are converged to within less than 2%
geometries are given in Ref. 52. To directly com- Žbond lengths. and 10% Žkinetic energy.. To facili-
pare the simulations started at the reaction or tate the propagation we used a potential ceiling at
transition configurations, exactly the same initial V s 100 kcalrmol above the energy of the reactant
positions and velocities were used for the classical geometry, which is the absolute minimum of the
atoms for both the QCrTDSCF and the QCrA, potential energy surface. The value of the ceiling
and the QCrTDSCF initial proton wavepacket was must be chosen high enough to both compute the
chosen as the ground state wave function corre- ground state of the proton and to propagate the
sponding to the configuration of the classical nu- QCrTDSCF wavepacket accurately. Increasing its
clei. The ground state was obtained by starting value did not change our results. We have also
with a Gaussian wavepacket centered in the deep- confirmed that our quantum and classical time

1766 VOL. 18, NO. 14


QUANTUM-CLASSICAL MD METHODS

FIGURE 2. Geometric data as a function of time, taken from the trajectories starting from the reactant configuration.
The first row of panels shows the O a —O b distance, the second row the O a —H a and O b —H a distances, and the third
row the O a —C a and O b —C b distances. The columns of panels are, from left to right, the classical, adiabatic, and
QCrTDSCF results. Time is indicated in ps.

steps are small enough to assure correct propaga- Most of this time was spent evaluating the energy
tion; that is, decreasing them does not change the and forces at each of the 32,768 grid points needed
trajectories significantly. for the proton wave function.
To calculate the proton ground state wave func-
tion with the imaginary time relaxation method,
we used a convergence criterion of <² V :nq 1 y Results and Discussion
² V :n < F 1.0 = 10y1 0 hartrees, where ² V :n is the
average potential for the nth iteration of the relax- We first compare the trajectories obtained from
ation process. We found that it was critical to the different methods. Figures 2 and 3 show sev-
impose such a strict convergence criterion to ob- eral bond distances obtained from the classical,
tain conservation of energy throughout the adia- QCrTDSCF and QCrA calculations starting from
batic simulations. Using this criterion, the number the reactant and transition geometries, respec-
of relaxation iterations required to converge to the tively. For the QCrTDSCF and QCrA simulations,
ground state, per geometry of the classical atoms, the O—H distance is calculated as ² rO — H Ž t .:,
was as large as 400. To calculate the ground state averaging over the proton wave function. For the
wave function at time t q tc we began the itera- trajectories starting from the reactant configuration
tion with the ground state wave function at time t. ŽFig. 2., the classical, QCrTDSCF, and QCrA cal-
Difficulties in realizing conservation of energy in culations all give very similar results. It appears,
QCrA propagation have been discussed previ- however, that at least for the Oa —H a , O b —H a ,
ously in the context of the dynamics of the sol- and Oa —O b bonds, the classical calculation yields
vated electron.59 bond lengths that oscillate more rapidly than their
The CPU times, on a 40-MHz R3000 SGI ma- QCrTDSCF counterparts, which in turn oscillate
chine, were about 50 and 140 hours for each more quickly than the corresponding QCrA bond
QCrTDSCF and QCrA simulation, respectively. lengths. This probably indicates that energy flow is

JOURNAL OF COMPUTATIONAL CHEMISTRY 1767


SHARAFEDDIN ET AL.

FIGURE 3. Same as Figure 2 except for the trajectories starting from the transition state configuration. Time is
indicated in ps.

less constrained in the classical than in the of the QCr TDSCF wavepacket at time t q t de-
QCrTDSCF simulation and less constrained in the pends not only on the potential at time t q t but
QCrTDSCF simulation than in the adiabatic simu- also on the wavepacket at time t, which results in
lation. a certain inertia in the dynamics of the proton
For the transition state trajectory the Oa —H a , wavepacket. The evolution of the QCrA and the
O b —H a , and Oa —O b bond length from QCr QCr TDSCF wavepackets for the trajectory initi-
TDSCF and QCrA are noticeably different. The ated at the transition state is shown in Figure 4. It
average bond lengths computed with the QCr A can be seen that the adiabatic ground state is
method change more abruptly than their QCr always localized near the deepest minimum,
TDSCF counterparts because of the nature of the whereas the position of the QCrTDSCF wavepack-
time dependence of the proton wave function. For et can stray from the minimum Žthis is most clearly
both the QCr A and the QCr TDSCF trajectory, visible at 120 fs.. This is a consequence of the
the proton wave function is chosen to be the QCr TDSCF wavepacket’s dependence on its pre-
ground state at t s 0. Because the proton’s poten- vious position.
tial has a single well when the classical atoms The two mixed quantum-classical methods yield
assume their positions at the transition state con- different configurational properties during the tra-
figuration, the initial wave function is localized. jectories Žproton position, bond lengths, etc.. due
During the trajectory, the QCr A wave function to the influence of inertial effects. In addition,
adapts instantaneously to the configuration of the there are also important differences in the dynami-
classical atoms. The wave function remains local- cal properties. For example, as shown in Figure 5,
ized as we propagate and simply falls into a well the average proton kinetic energy is much higher
because its position at time t q t depends only on for the QCrTDSCF simulation than for the adia-
the values of the potential at time t q t and not on batic simulation. This is consistent with the obser-
its position Žor the potential. at time t. The evolu- vation that the values of the wave vector k, which
tion of the QCr A proton wave function has thus contribute to the QCrTDSCF wavepacket, are
no memory of its past, which corresponds to a larger than the values that contribute to the proton
total absence of inertial dynamical effects. This is ground state wave function and that the QCr
not true of the QCr TDSCF trajectory. The position TDSCF wave function has more spatial oscillations

1768 VOL. 18, NO. 14


QUANTUM-CLASSICAL MD METHODS

Žsee Fig. 4.. Due to the neglect of inertial effects, energy. On the other hand, there is almost no
the proton does not need to dissipate kinetic en- noticeable difference in the two types of propaga-
ergy as it falls into the energy well during the tion for the trajectory initiated in the reactant con-
QCrA trajectory. In contrast, inertial effects during figuration because the QCrA and the QCr TDSCF
the QCrTDSCF trajectory result in a higher kinetic wave functions behave very similarly in that case.
The reason is that the potential felt by the proton is
always a single well potential and does not change
much with time. Nevertheless, it should be em-
phasized that the ultimate fate of the classical,
QCr A, and QCr TDSCF trajectories, initiated at
the transition state, is the same.
For the transition state trajectories, the QCrA
method appears to be clearly inadequate. During
the QCrA trajectories the average proton position
varies too rapidly because the proton wave func-
tion reacts instantaneously to the configuration of
the classical atoms. In effect, the rapid time varia-
tion of the proton potential energy surface near the
transition state causes a breakdown of the adia-
batic approximation expressed by eq. Ž6.. As is
clear from the first derivation of the QCrA equa-
tions presented above, the QCrA method cannot
be better, and in fact must be worse than the
QCrTDSCF results, because one must make an
additional approximation to obtain the QCrA
equations from the QCrTDSCF equations. Al-
though we do not claim to have implemented
either method in the most efficient manner possi-
ble, we note that our QCrA calculations are much
more costly than the QCrTDSCF calculations. The
QCrA calculation is more costly, because, to con-
serve energy during the propagation, we needed
to converge the instantaneous ground states very
accurately.
An important question remains: Should one ex-
pect the QCrTDSCF approximation to be accurate
for our proton transfer problem? It is clear that the
QCrTDSCF approximation will be good if the
FIGURE 4. A one-dimensional cut through the proton effect of the coupling between the r and the R
potential energy surface (solid curve), the SCF proton degrees of freedom is well represented by average
density (dashed curve), and the ground state proton potentials. In general, the approximation will be
density corresponding to the potential energy surface poor if the r degree of freedom wavepacket bifur-
(dotted curve) for several configurations along the cates. Although our proton wavepacket is gener-
QCrTDSCF trajectory starting from the transition state. ally localized about a single point, sometimes the
The time interval between consecutive snapshots is 20 proton wavepacket does briefly separate into two
fs. The first snapshot represents the starting parts. Such a bifurcation is shown in Figure 6.
configuration; the single proton density profile shown When bifurcations occur, a trajectory determined
represents the ground state. The line along which the
from a mean force might not represent the true
values are given is parallel to the O a —O b axis and
passes through the mean value of the proton position.
dynamics. However, even if the proton wavepacket
The zero point on this line is defined by the projection of does separate into two parts, the QCrTDSCF ap-
the midpoint between O a and O b onto this line. The proximation should be good if either: Ži. the posi-
proton densities have been scaled arbitrarily (but tion of the classical nuclei does not depend sensi-
identically), and their baselines indicate the energy of the tively on the proton wavepacket Ži.e., the coupling
corresponding wavepacket. (See also Fig. 6.) is very weak.; or Žii. the motion of the proton is

JOURNAL OF COMPUTATIONAL CHEMISTRY 1769


SHARAFEDDIN ET AL.

FIGURE 5. The average kinetic energy, ² T :, for the transition state trajectory as a function of time from the
QCrTDSCF (solid line) and adiabatic (dashed line) simulations.

fast relative to the response time of the classical QCrA. Trajectories were run on a realistic multidi-
nuclei. The sensitivity of the trajectory of the clas- mensional potential energy surface fitted to ab
sical atoms to an occasional and brief bifurcation initio properties of the acetylacetone molecule. The
of duration D t will depend on the magnitude of three quantum degrees of freedom of the proton
the resulting uncertainty in the impulse, D Frms D t, were treated quantum mechanically and the re-
relative to the net impulse ² F : D t caused by the maining 42 degrees of freedom were treated classi-
average force and the net momentum P. Thus, we cally. Although such computational approaches
expect the QCrTDSCF approximation to remain have been used previously by others in various
quite good for simulations at 300 K, despite occa- contexts,39 ] 46 it is the first time, to our knowledge,
sional splitting of the proton wavepacket, because that the QCrTDSCF and QCrA methods have
the proton motion is fast enough that the motion been applied to study the multidimensional dy-
of the classical nuclei is strongly determined by namics of a proton in a complex molecule using a
the average potential irrespective of the bifurca- realistic potential Žof course, proton transfers in
tion. In the tunnelling regime this would not be multidimensional systems have been investigated
true: the motion of the proton would not be fast previously using other approaches; see for exam-
enough to compensate for bifurcation of the ple Refs. 19]25.. For classical initial velocities cho-
wavepacket. Where QCrTDSCF is insufficient one sen from a Boltzmann distribution at 300 K the
might resort to stochastic surface-hopping al- results of classical, QCrTDSCF, and adiabatic cal-
gorithms 48 ] 50 or multiconfiguration TDSCF.60 ] 62 culations were compared. We find that the time-
dependent QCrTDSCF wavepacket, c Žr, t ., and
the adiabatic ground state of the proton, f 0 Žr; RŽ t ..,
Conclusion sometimes behave rather differently. The position
of the adiabatic ground state depends on the posi-
Initial results of a mixed quantum-classical tion of the deepest minimum of the proton poten-
molecular dynamics study of acetylacetone have tial. If the classical atoms move so as to create a
been presented. Two quantum-classical dynamical potential for the proton with one deep minimum,
propagation methods were used, QCrTDSCF and then the proton ground state wave function imme-

1770 VOL. 18, NO. 14


QUANTUM-CLASSICAL MD METHODS

their QCrTDSCF counterparts by making an adia-


batic approximation w see eq. Ž6.x . This obviously
means that the QCrTDSCF results are always more
accurate than the QCrA results. Although both
methods are approximate one can assess the im-
portance of nonadiabatic coupling by comparing
the QCrTDSCF and the QCrA results. Differences
between our QCrTDSCF and our QCrA results
are a measure of the importance of nonadiabatic
coupling. It is important to study the difference
between QCrA and QCrTDSCF for a realistic
potential. We have observed important differences
for a realistic Žmultidimensional. acetylacetone po-
tential. As we implement them, the QCrTDSCF
method is also more computationally efficient than
the QCrA method. Due to the apparent impor-
tance of the nonadiabatic coupling, and the rela-
tive inefficiency of the QCrA method, we there-
fore recommend the QCrTDSCF over the QCrA
approach.
Simulations of the kind we have performed are
too short to actually observe a net proton transfer.
The QCrA and QCrTDSCF trajectories do not
provide an effective sampling of transfer events
for cases with a non-negligible activation barrier.
However, the present calculations are only the first
stage of investigations of the dynamics of proton
transfer in acetylacetone. In the future, we will
generate an ensemble of activated trajectories initi-
ated near the transition state. Such calculations
will enable us to calculate the proton transfer rate
constant and provide invaluable information con-
FIGURE 6. Three-dimensional density plot of the cerning important recrossing events at the transi-
QCrTDSCF wave function superimposed on the tion state. To calculate the rate constant it is obvi-
molecule at 0, 20, and 180 fs along the transition state ously necessary to develop and understand the
trajectory. limitations of methods for solving the dynamics. In
addition, the influence of solvent molecules during
the proton transfer reaction could be investigated
diately moves so that it sits over this minimum. by performing simulations of acetylacetone in po-
The position of the QCrTDSCF wavepacket at lar and nonpolar solvents. Finally, because there
time t q t depends not only on the potential the simulations will be CPU intensive, it will be
proton feels at time t q t , but also on where the worthwhile to explore the possibility of accelerat-
wavepacket was at time t. If the configuration of ing the calculations by optimizing the proton grid,
the classical atoms changes very slowly, the posi- improving the propagation algorithm, and reduc-
tion of the proton ground state will move slowly ing the number of times the potential is evaluated.
and it will be more consistent with the QCrTDSCF
wavepacket.
As we discussed, although the quantum-mecha- Acknowledgments
nical adiabatic equations are not obtained from the
quantum-mechanical TDSCF equations by intro- B. Roux is a MRC research fellow. This work
ducing an adiabatic approximation, the working was supported by grants from FCAR ŽQuebec. and
equations of the QCrA method are obtained from NSERC ŽCanada..

JOURNAL OF COMPUTATIONAL CHEMISTRY 1771


SHARAFEDDIN ET AL.

34. R. B. Gerber and M. A. Ratner, Adv. Chem. Phys., 70, 97


Ž1988..
References
35. R. B. Gerber and R. Alimi, Israel J. Chem., 31, 383 Ž1991..
1. S. Scheiner, Acc. Chem. Res., 18, 174 Ž1985.. 36. R. Alimi, R. B. Gerber, A. D. Hammeric, R. Kosloff, and
M. A. Ratner, J. Chem. Phys., 93, 6484 Ž1990..
2. R. P. Bell, The Proton in Chemistry, Chapman & Hall, Lon-
don, 1973. 37. G. D. Billing, Comp. Phys. Rep., 12, 383 Ž1990..
3. D. Borgis, Electron and Proton Transfer Processes in Chemistry 38. G. D. Billing, Chem. Phys., 70, 223 Ž1982..
and Biochemistry. Elsevier, New York, 1991. 39. K. Haug and H. Metiu, J. Chem. Phys., 99, 6253 Ž1993..
4. D. Borgis and J. T. Hynes, Chem. Phys., 170, 315 Ž1993.. 40. P. Bala, B. Lesyng, and J. A. McCammon, Chem. Phys., 180,
´ and R. Silbey, J. Chem. Phys., 94, 4809 Ž1991..
5. A. Suaez 271 Ž1994..
6. J. Lobaugh and G. A. Voth, J. Chem. Phys., 100, 3039 Ž1993.. 41. P. Bala, B. Lesyng, and J. A. McCammon, Chem. Phys. Lett.,
7. H.-H. Limbach and J. Hennig, J. Chem. Phys., 71, 3120 219, 259 Ž1994..
Ž1979.. 42. H. J. C. Berendsen and J. Mavri, J. Chem. Phys., 97, 13464
8. Y. Tomioka, M. Ito, and N. Mikami, J. Phys. Chem., 87, 4401 Ž1993..
Ž1983.. 43. J. Mavri, H. J. C. Berendsen, and W. F. van Gunsteren, J.
9. T. Carrington, Jr. and W. H. Miller, J. Chem. Phys., 84, 4364 Chem. Phys., 97, 13469 Ž1993..
Ž1986..
44. D. Laria, G. Ciccotti, M. Ferrario, and R. Kapral, J. Chem.
10. R. Kosloff, J. Phys. Chem., 92, 2087 Ž1988.. Phys., 97, 378 Ž1992..
11. S. Carter and N. C. Handy, Comp. Phys. Rep., 5, 115 Ž1986.. 45. D. Borgis, G. Tarjus, and H. Azzouz, J. Chem. Phys., 96,
12. J. Tennyson, S. Miller, and J. R. Henerson, Methods in 3188 Ž1992..
Computational Chemistry, Vol. 4, Plenum Press, New York, 46. D. Borgis, G. Tarjus, and H. Azzouz, J. Chem. Phys., 97,
1992. 1390 Ž1992..
13. M. J. Bramley and T. Carrington, J. Chem. Phys., 99, 8519 47. A. Warshel and J.-K. Hwang, J. Chem. Phys., 84, 4938 Ž1986..
Ž1993..
48. S. Hammes-Schiffer and J. C. Tully, J. Chem. Phys., 101,
14. R. P. Feynman and A. R. Hibbs, Quantum Mechanics and 4657 Ž1994..
Path Integrals, McGraw-Hill, New York, 1965.
49. S. Hammes-Schiffer and J. C. Tully, J. Chem. Phys., 103,
15. D. Chandler, In Liquides, Cristallisation et Transition Vitreuse,
8528 Ž1995..
Les Houches, Session LI, D. Levesque, J.-P. Hansen, and J.
Zinn-Justin, Eds., Elsevier, New York, 1991. 50. S. Consta and R. Kapral, J. Chem. Phys., 104, 4581 Ž1996..
16. M. J. Gillan, J. Phys. C, 20, 3621 Ž1987.. 51. D. ter Haar, Problems in Quantum Mechanics, Pion, London,
1975.
17. G. A. Voth, D. Chandler, and W. H. Miller, J. Chem. Phys.,
91, 7749 Ž1989.. 52. K. Hinsen and B. Roux, J. Comput. Chem., 18, 368 Ž1997..
18. G. A. Voth, J. Phys. Chem., 97, 8365 Ž1993.. 53. B. R. Brooks, R. E. Bruccoleri, B. D. Olafson, S. Swami-
19. J.-K. Hwang and A. Warshel, J. Phys. Chem., 97, 10053 nathan, D. J. States, and M. Karplus, J. Comput. Chem., 4,
Ž1993.. 187 Ž1983..
20. H. Azzouz and D. Borgis, J. Chem. Phys., 98, 7361 Ž1993.. 54. M. J. Frisch, M. Head-Gordon, G. W. Trucks, J. B. Foresman,
H. B. Schlegel, K. Raghavachari, M. Robb, J. S. Binkley, C.
21. D. Laria, G. Ciccotti, M. Ferrario, and R. Kapral, Chem.
Gonzalez, D. J. Defrees, D. J. Fox, R. A. Whiteside, R.
Phys., 180, 181 Ž1994..
Seeger, C. F. Melius, J. Baker, R. L. Martin, L. R. Kahn,
22. S. Consta and R. Kapral, J. Chem. Phys., 101, 10908 Ž1994.. J. J. P. Stewart, S. Topiol, and J. A. Pople, GAUSSIAN,
23. K. Hinsen and B. Roux, J. Chem. Phys., 106, 3567 Ž1997.. Gaussian, Inc., Pittsburgh, PA 1990.
24. A. Warshel, J. Phys. Chem., 86, 2218 Ž1981.. 55. M. D. Feit, J. A. Fleck, Jr., and A. Steiger, J. Comput. Phys.,
25. A. Warshel and Z. T. Chu, J. Chem. Phys., 93, 4003 Ž1990.. 47, 412 Ž1982..
26. J. Cao and G. A. Voth, J. Chem. Phys., 101, 6157 Ž1994.. 56. W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P.
27. A. Warshel and R. M. Weiss, J. Am. Chem. Soc., 102, 6218 Flannery, Numerical Recipe in C, Cambridge University
Ž1980.. Press, Cambridge, UK, 1992.
28. Y.-T. Chang and W. H. Miller, J. Phys. Chem., 94, 5884 57. R. Kosloff and H. Tal-Ezer, Chem. Phys. Lett., 127, 223
Ž1990.. Ž1986..
29. S. Takada and H. Nakamura, J. Chem. Phys., 100, 98 Ž1994.. 58. M. P. Allen and D. J. Tildesley, Computer Simulations of
Liquids, Clarendon Press, Oxford, 1987.
30. Y. Guo, T. D. Sewell, and D. L. Thompson, Chem. Phys.
Lett., 224, 470 Ž1994.. 59. F. Webster, P. J. Rossky, and R. A. Friesner, Comput. Phys.
31. Z. Smedarchina, W. Siebrand, and M. Z. Zgierski, J. Chem. Commun., 63, 494 Ž1991..
Phys., 103, 5326 Ž1995.. 60. N. Makri and W. H. Miller, J. Chem. Phys., 87, 5781 Ž1987..
¨ J. Chem. Phys., 91,
32. N. Shida, P. F. Barbara, and J. E. Almlof, 61. H. D. Meyer, U. Manthe, and L. S. Cederbaum, Chem. Phys.
4061 Ž1989.. Lett., 165, 73 Ž1990..
´ J. Chem.
33. E. Bosch, M. Moreno, J. M. Lluch, and J. Bertran, 62. U. Manthe, H. D. Meyer, and L. S. Cederbaum, Chem. Phys.
Phys., 93, 5865 Ž1990.. Lett., 97, 9062 Ž1992..

1772 VOL. 18, NO. 14

You might also like