You are on page 1of 10

Simulation of Slow Reaction with

Quantum Character: Neutral Hydrolysis


of Carboxylic Ester

MARC F. LENSINK,1, 2 JANEZ MAVRI, 2, 3


HERMAN J. C. BERENDSEN 2
1
´ de Calcul Atomique et Moleculaire
Centre Europeen ´ (CECAM), Ecole Normale Superieure
´ de Lyon,
´
46 Allee d’Italie, 69364 Lyon cedex 07, France
2
BIOSON Research Institute, Department of Biophysical Chemistry, the University of Groningen,
Nijenborgh 4, 9747 AG Groningen, The Netherlands
3
National Institute of Chemistry, P.O.B. 30, Hajdrihova 19, 61115, Ljubljana, Slovenia

Received 6 July 1998; accepted 27 January 1999

ABSTRACT: By computer simulation, using both quantum and classical


dynamics, we determined the rate constant and the kinetic isotope effect of the
rate-determining step in the neutral hydrolysis of p-methoxyphenyl
dichloroacetate in aqueous solution. This step involves a proton transfer
concerted with the formation of a C—O bond. A method of biased sampling
was used; the Gibbs free energy of the biased configuration from which proton
transfer is likely to occur was determined by a combination of semiempirical
quantum calculations and thermodynamic integration. The proton dynamics
was modeled with the quantum-dynamical density matrix evolution method
that includes nonadiabatic pathways. The proton dynamics is driven by a
fluctuating proton potential that was derived from a classical molecular
dynamics simulation of the system including solvent. The calculated rate
constant of 3 = 10y2 sy1 agrees within the error of the calculation with the
experimentally observed value of 2.78 = 10y3 . The calculated pseudo-first-order
kinetic isotope effect of 3.9 is in good agreement with the experimentally
observed value of 3.2. The results show the feasibility of computational
approaches to slow reactions in complex environments, where proton transfer
with an essential quantum-dynamical nature is the rate-limiting step. 䊚 1999
John Wiley & Sons, Inc. J Comput Chem 20: 886᎐895, 1999

Keywords: nonadiabatic quantum dynamics; molecular dynamics; proton


transfer; ester hydrolysis

Correspondence to: H. J. C. Berendsen; e-mail: berendsen


@chem.rug.nl

Journal of Computational Chemistry, Vol. 20, No. 8, 886᎐895 (1999)


䊚 1999 John Wiley & Sons, Inc. CCC 0192-8651 / 99 / 080886-10
SIMULATION OF A SLOW QUANTUM REACTION

Introduction

E ster or amide bond cleavage by hydrolysis is


a process that occurs in many enzymatic
reactions. The total reaction consists of a cascade of
events in which proton transfers over hydrogen
bonds are important subprocesses and in many
cases even form the rate-determining steps.1 It is a
challenge to present-day computational techniques
to simulate such a reaction, combining the simula-
tion of a slow, activated event with the simulation
of a proton transfer with essential quantum-dy-
namical nature. In this article we describe the
calculation of the proton transfer rate in the neu-
tral hydrolysis of an ester in water that proceeds
on a time scale of several minutes.
Normal ester bond cleavage by hydrolysis can
occur rapidly in acidic or basic solution, but it is
unmeasurably slow in neutral solution for simple
esters like ethyl acetate. However, neutral hydro-
lysis can proceed in minutes in appropriately sub-
stituted esters Že.g., containing an electrophilic
group such as dichloromethyl. combined with the
presence of a strong leaving group. We chose to
study the neutral ester bond hydrolysis of p-
methoxyphenyl dichloroacetate in water, a well-
documented reaction that has a pseudo-first-order
rate constant k of Ž2.78 " 0.06. = 10y3 sy1 at 298 K
and a primary deuterium kinetic isotope effect
ŽKIE. of 3.24.2 It has been proposed that the reac- FIGURE 1. Structure of p-methoxyphenyl
tive complex Žsee Fig. 1. contains two water dichloroacetate with two water molecules corresponding
molecules and that the rate-limiting step in the to the (a) reactant state and (b) product state. AM1 in
neutral hydrolysis reaction is proton transfer be- vacuo optimizations were performed under the constraint
tween these water molecules concerted with the of a linear water᎐water hydrogen bond; the oxygen atom
of the nearest water molecules was constrained to be
formation of a tetrahedral intermediate.2
above the carbonyl carbon atom, perpendicular to the
In this article we compute the rate constant of
plane defined by atoms COO constituting the ester bond.
the concerted reaction as follows: The O—H bond lengths of the hydronium ion were
constrained to their AM1 calculated value of 1.004 A.˚
1. we investigate the Born᎐Oppenheimer hy-
The RCO , ROO , and ROH distances changed from 3.605
persurface of the reactive complex by ˚ 3.062 to 2.540 A,
to 1.442 A, ˚ and 0.963 to 1.536 A,˚
semiempirical quantum mechanics ŽQM.; respectively.
2. we choose proper reaction coordinates and
select the reactant configuration Žactivated method of density matrix evolution ŽDME.,4
state. from which proton transfer is probable; which uses time-dependent proton potentials
3. we compute the probability of that activated derived from MD; and
state in an aqueous environment by thermo- 5. we combine the results to obtain the initial
dynamic integration ŽTI. 3 based on re- overall reaction rate.
strained molecular dynamics simulations
ŽMD.; For reasons of clarity, we present theory and com-
4. we calculate the initial rate constant of the putational results together for each of the steps
proton and deuteron transfer reaction in the mentioned above. Computational details are given
activated state by the quantum-dynamical in the following section.

JOURNAL OF COMPUTATIONAL CHEMISTRY 887


LENSINK, MAVRI, AND BERENDSEN

Because of the many orders of magnitude in- The system consisted of a periodic box with 510
volved in the different multiplicative contributions water molecules that was coupled to a tempera-
to k, we prefer to express the contributions in ture bath at 298 K and pressure bath at 1 bar with
powers of 10 as logŽ krsy1 .. The experimental value coupling constants of 50 fs.24 Hydrogen atoms of
of log k is y2.56 " 0.01. the ester were treated explicitly and had small
To our knowledge only a few studies of reactive repulsive Lennard᎐Jones potentials on them. Bonds
systems in the condensed state have been pub- involving hydrogen atoms were constrained using
lished that combine classical dynamical simula- the SHAKE algorithm.25 A twin-range cutoff of 8
tions with nonadiabatic quantum dynamics of and 10 A ˚ was employed, and the pair list was
nuclear coordinates. Most studies concern model updated every 10 steps. A time step of 1 fs was
systems or appropriately chosen systems with used for the integration of the equations of motion
small activation barriers.5, 6 New approaches in- using the leap-frog Verlet algorithm. Partial atomic
clude surface hopping,7 real-time path integral charges for the solute were determined by fitting
methods8 and centroid dynamics,9 and a nonadia- them to the electrostatic potential in a solvent
batic quantum transition state method to simulate reaction field ŽAM1-SM1. 26 and to the calculated
infrequent events.10 The effects of the environment dipole moment, using the Bessler᎐Merz᎐Kollman
on the Gibbs free energy of transition states11 and procedure.27 For the path from reactant Ž R . config-
on proton potentials12 and dynamics13 in enzymes uration to biased configuration Ž) state in later
were considered. Static solvent effects were in- section. linear interpolation was used between
cluded in semiempirical molecular orbital ŽMO. these two states, and between the ) state and the
calculations on basic carboxylic ester bond clea- product configuration Ž P . charges were deter-
vage14 using a solvent reaction field.15 Possible
mined by cubic spline interpolation using two
proton transfer paths in enzyme carbonic an-
additional intermediate points.
hydrase were investigated using ab initio tech-
During TI the system was moved from the )
niques.16 DME was applied to proton transfer in
configuration Ž ␭ s 1. to the R configuration Ž ␭ s
hydrogen malonate.17, 18 Proton transfer from wa-
0. for six fixed values of ␭. Each run consisted of a
ter to histidine in phospholipase A 2 was studied
25-ps equilibration starting from the previous run
by wave-packet evolution of the proton, embed-
Ž50 ps for the initial run., followed by a 50-ps
ded in an empirical valence bond MD simulation.19
simulation. The values of dGrd ␭ were integrated
Except for ref. 10, none of these studies combined
with the extended trapezoidal rule to obtain ⌬G.
quantum-dynamical proton transfer with an acti-
vated reactive step. Errors were estimated by dividing each run into
The essential quantum-dynamical feature of two halves and considering the variance of the
proton transfer is that nonadiabatic processes Ži.e., distribution of all 64 combinations of the 12
transitions between different protonic quantum derivatives. Using only the first halves of the six
levels. may take place because the level separation runs or only the second halves gave the same
is neither large enough compared to thermal ener- result within the statistical error, indicating that
gies to justify an adiabatic nor small enough to sufficient equilibration had taken place.
justify a classical treatment. Such transitions are DME was carried out for a period of 12 ps,
driven by fluctuating interactions with the dy- starting at time origins that were each 100 fs after
namic environment and must be treated by proper the previous one and using a time step of 0.1 fs.
nonadiabatic quantum-dynamical methods. The ¨
The time-dependent Schrodinger equation was
electrons can still be adequately treated in the solved using 750 different time origins. Thus, a
Born᎐Oppenheimer limit. total of 87 ps of the proton potentials obtained
from the MD simulation Žsee also later section.
was in fact used. In each DME solution the density
Computational Details matrix was set initially to exclusive population of
the ground state Žcorresponding to the reactant
For semiempirical QM calculations the pro- state. with off-diagonal elements set to zero, im-
grams MOPAC and MOBOSOL20 were used. Ab plying unknown phase information. Five simple
initio calculations were carried out with the pro- Gaussian basis functions were used to describe the
gram Gaussian.21 proton. These were optimized to have their first
For classical MD simulations the GROMOS three energy levels correspond with the true en-
package 22 was used with the SPC water model.23 ergy levels of the proton in the average proton

888 VOL. 20, NO. 8


SIMULATION OF A SLOW QUANTUM REACTION

potential. This level of description is adequate to The heat of formation for the reactant and prod-
include tunneling and nonadiabicity.18 uct configurations was calculated for several val-
ues of these two coordinates and was found to be
46.9 kcalrmol higher for the relaxed product state
Neutral Hydrolysis Reaction than for the relaxed reactant state. This in vacuo
difference was smallest for a shortened CO bond
We first give a short qualitative description of and appeared to depend much more strongly on
the succession of events in the neutral hydrolysis R CO than on R OO . The potential wells of the re-
of an ester in aqueous solution in order to derive laxed reactant and product states were approxi-
the necessary computational steps. We divide our mated by fitting to two harmonic oscillators: the
system ŽFig. 1a, reactant state; b, product state. force constants for R CO and R OO were, respec-
into three subsystems: solvent Žwater., solute Žes- tively, 67.2 and 9.0 kcal moly1 A
˚ y2 in the relaxed
ter complexed with two water molecules., and reactant state and 973 and 101 kcal moly1 A ˚ y2 in
the proton. The proton transfer, together with the the relaxed product state. The larger force con-
simultaneous electron distribution rearrangement, stants in the product state indicate the covalent
in the rate-limiting step can only take place when character of the CO bond and the strong ionic
the heavy-atom configuration Žwith the proton in bond to the hydronium ion.
the reactant state. happens to be favorable for The semiempirical data were checked by ab
these concerted reactions Ži.e., when the nuclear initio calculations at the Hartree᎐Fock ŽHF.r6-
geometry is close to that of the product state, the 31G* level and using density functional theory
tetrahedral intermediate depicted in Fig. 1b.. In ŽDFT. for selected AM1-generated configurations.
this improbable nuclear configuration the fluctuat- Depending on the level of theory, energy differ-
ing potential felt by the proton, which is mainly ences between product and initial state were higher
due to the electric field of water molecules, causes than the AM1 results by 5 ŽDFT. to 20 ŽHF.
the proton to transfer to the product side of the kcalrmol, depending on the chosen method and
hydrogen bond. The tetrahedral intermediate, basis set. This is partly due to the fact that the
which may become further stabilized by solvent configurations were not relaxed in the ab initio
reorganization, subsequently dissociates irre- calculations but is mostly due to the intrinsic in-
versibly into products, picking up a proton in the compatibility between ab initio methods and meth-
process, while the hydronium ion diffuses away. ods that are parametrized to reproduce experimen-
tal results. Mainly because of the computational
cost of the ab initio methods, we chose to use the
Born᎐Oppenheimer Surface of semiempirical values for further analysis and for
Reactant and Product construction of the time-dependent proton poten-
tials Žsee later section..
Due to its size, the solute does not easily allow
for repeated high-level ab initio calculations. It was
therefore treated with the AM1 semiempirical MO
method, which was proven to be successful for
Biased State for Proton Transfer
obtaining the energetics of organic reactions with and Its Probability
accuracies in the 1᎐2 kcalrmol range.28
The geometries of the reactant and product con- It was found that the QM energies were rela-
figurations were both optimized in vacuo Žsee Fig. tively insensitive to R OO , suggesting that only the
1.. It was found that, apart from the proton posi- constraint of a shortened CO bond would suffice.
tion, the only significant changes occurred for the However, it is likely that the probability of proton
carbonyl carbon᎐water oxygen distance R CO and transfer itself depends strongly on R OO . Rather
the distance between the two water oxygens R OO , than performing several quantum-dynamical sim-
which were chosen to represent the biasing coordi- ulations for different R OO , we estimated the pro-
nates. ŽIn fact, there are three reaction coordinates: ton potential at the AM1-SM1 level with a short-
R CO , R OO , and R OH ; we refer to R CO and R OO as ened CO bond for different values of R OO . The
the biasing coordinates, while R OH is the pure results showed a significantly lower energy barrier
reaction coordinate.. for a shortened R OO coordinate, which strongly

JOURNAL OF COMPUTATIONAL CHEMISTRY 889


LENSINK, MAVRI, AND BERENDSEN

U
suggests that the biased form of choice is the one where ⌬Gmf is the potential of mean force at the
where R OO also has its product value. This is reactive Ži.e., product. coordinates and k H is now
indeed the configuration used in this work. the proton transfer rate from the ) state.
Separating the two degrees of freedom r 1 s R CO
and r 2 s R OO as the biasing coordinate plane, the
probability for the system to reside in a surface Potential of Mean Force
element dr1 dr 2 on this plane is proportional to of Biased State
expŽy⌬Gmf Ž r 1 , r 2 .rRT . dr1 dr 2 , where ⌬Gmf Ž r 1 , r 2 .
U
is the potential of mean force and equilibration over The potential of mean force ⌬Gmf at the prod-
all other degrees of freedom is assumed. ⌬Gmf is uct-state coordinates consists of a QM component
arbitrarily chosen to be zero at the bottom of the for the solute and a classical solvent component.
potential well in the reactant state. The probability For the QM component we use the in vacuo energy
of being in the reactant state is proportional to the difference of the reactant state Ž46.9 kcalrmol, see
integral of expŽy⌬Gmf Ž r 1 , r 2 .rRT . over a region of earlier section., diminished with the energy differ-
the biasing coordinate plane that covers all ther- ence computed in the next section between the
mally occupied configurations of the reactant state. proton on the product side and on the reactant
For two harmonic oscillators with force constants side. The in vacuo quantum-mechanical compo-
k 1 and k 2 nent of that energy is 20.1 kcalrmol. Thus, the
U
solute component of ⌬Gmf equals 26.8 kcalrmol.
The solvent part is calculated by TI using MD.
HR expŽy⌬G Ž r . rRT . dr s 2␲ RT Ž k 1 k 2 . y1 r2 ,
mf The method we used consists of defining harmonic
restraining potentials in the biasing coordinate
plane for the reactant Ž R . state and the biased Ž).
where r stands for Ž r 1 , r 2 . and the integral is taken
state: the chosen harmonic force constants were
over the reactant region R.
k 1 s 67.2 and k 2 s 9.0 kcal moly1 A ˚ y2 for R CO
We make two assumptions:
and R OO in the R state and k 1 s kX2 s 3000 kcal
X

moly1 A ˚ y2 for both directions in the ) state. The


1. The proton transfer rate k H Ž r . is proportional former are values fitted to AM1 calculations; the
to the probability density of the Žrelaxed. latter values are arbitrary restraints that confine
product state to be at r in the biasing coordi- the bias state but have no influence on the result.
nate plane and therefore strongly peaks in a A path, characterized by a coupling parameter ␭
region that we will indicate by ). The ratio- between 0 and 1, was defined. Intermediate points
nale for this is that the electron redistribu- were obtained by linear interpolation of biasing
tion, concerted with proton transfer, yields coordinates, force constants, and partial charges.
the product state as the end product of this The equilibrated ensemble average at constant
single reactive step. temperature and pressure for ⭸ Vr⭸␭ Ž V is the
2. The potential of mean force ⌬Gmf is constant potential energy. at any ␭ yields the derivative of
within the ) region of the biasing coordinate the Gibbs free energy dGrd ␭ at that point.
plane. Although not correct, the error intro- The result for the Gibbs free energy difference
duced by a slope of ⌬Gmf is negligible com- ⌬G* s G* y G R is y13.2 " 0.4 kcalrmol. If this
pared to the error in ⌬Gmf itself. result is recast in terms of the potential of mean
U
force for the solvent contribution to ⌬Gmf , using
With these assumptions and approximating the 1 k1 k 2
U
product well as two harmonic oscillators with force ⌬s Gmf s ⌬G* y RT ln , Ž2.
2 kX1 kX2
constants kU1 and kU2 , the rate can be expressed as
where kX1 and kX2 are the force constants of the
H# k H Ž r . exp Ž y⌬Gmf Ž r . rRT . dr restraining potential imposed on the ) state, we
ks obtain ⌬s GmfU
s y16.1 " 0.4 kcalrmol. Hence, the
HR exp Ž y⌬Gmf Ž r . rRT . dr
total potential of mean force at the ) state is
U
k1 k 2 ⌬Gmf s 26.8 y 16.1 s 10.7 kcalrmol. We note that
s kH ( kU1 kU2
U
ey⌬G mf r RT , Ž1. the solvent significantly favors the ) state and
thus facilitates the reaction.

890 VOL. 20, NO. 8


SIMULATION OF A SLOW QUANTUM REACTION

points between the two water oxygens. For each of


Biased Proton Reaction Rate these 10 5 configurations the solute contribution to
the potential Vsolute was calculated on the AM1
The next step is to compute the proton quan- level and to this the classical contribution Vsolvent
tum-dynamical reaction rate k H from the biased was added. Cubic spline interpolation was used to
configuration. For this purpose we use a 100-ps complete the potential curves. All solvent᎐solute
dynamic MD run at the ) state. The trajectory was interactions were included, using solute partial
saved every 10 fs, which preserves all information charges at several positions of the proton. In this
available from the MD run. In this classical MD way the complete QM response of the solute to the
run the proton was modeled as a point charge proton position, including polarizability contribu-
fixed in the reactant region, which is an excellent tions, was incorporated.
approximation for interactions with the environ- The time evolution of the classical splitting
ment.29 ⌬ EŽ t ., defined as the energy difference between
Because the proton transfer rate is small on the the reactant and product states,31, 32 is depicted in
time scale of the simulation, we can only derive Figure 2. Although the average proton potential is
the initial rate of the reaction. The proton remains about 10 kcalrmol higher on the product side
in the reactant configuration and no reverse reac- without a clear well and thus would not favor
tion Žor barrier recrossing in classical terms. will proton transfer, Figure 2 shows that on many occa-
occur. This has the advantage that the quantum- sions during the 100-ps simulation a splitting oc-
dynamical part of the calculation can be carried curs that is zero or even negative. It is these points
out a posteriori.30 of the simulation that contribute the most to the
For each of the 10 4 MD frames the proton po- rate constant. The use of biased sampling ensured
tential was determined by taking the solute config- that many of such points did occur during the
uration and putting the proton on 10 different 100-ps simulation time.

FIGURE 2. Time evolution of the splitting ⌬ E ( t ), defined as the energy difference between the structure with the
proton in the product well and the one with the proton in the reactant well.

JOURNAL OF COMPUTATIONAL CHEMISTRY 891


LENSINK, MAVRI, AND BERENDSEN

The normalized probability density w Ž ⌬ E . of Note that Figure 3 represents a single well, be-
the classical energy splitting for the ) state ŽFig. 2. cause in this work only the reactant region is
is shown in Figure 3. The top graph in Figure 3 simulated. The curve of Figure 3 provides the
shows the probability density, the bottom graph probability of zero-energy splitting that could be
the corresponding free energy profile ⌬GŽ ⌬ E . s used in an analytical tunneling rate theory. In this
yk B T log w Ž ⌬ E . q k B T log w Ž² ⌬ E :.. It can be work, however, we include the full fluctuations in
seen that the energy splitting has a Gaussian dis- the prediction of proton transfer rates.
tribution. This is the result of the perturbation The QM parts of the proton potentials, calcu-
coming from many independent sources, because lated in this procedure, gave as an average
the interaction is mainly electrostatic and thus ⌬ EQM Ž t . s y20.1 " 3.0 kcalrmol. This value was
long ranged. It is remarkable that this quadratic used for correction of the QM part of the bias
behavior for ⌬ E, as assumed in Marcus’ theory,33 energy Žprevious section..
is in fact observed for proton transfer.5, 6, 17, 34, 35 With the 100-ps proton potentials in hand, the
¨
time-dependent Schrodinger equation was solved
with the DME method using 750 different starting
points. The product state occupation ⌰ is defined
as

⌰Ž t . s H␰ ⌿U Ž ␰ , t . ⌿ Ž ␰ , t . d ␰ , Ž3.
barrier

where ␰ is the proton coordinate and ␰ barrier is


chosen at the energy barrier for the time-averaged
proton potential. Reported values are averaged
over all time origins. The coarse-grained slope of
the slowly increasing ² ⌰ : ŽFig. 4. can then be
extrapolated 4, 18 to yield a first-order rate constant
k H s 2.4 = 10 7 sy1 . The inaccuracy of this value is
about a factor of 2 and we conclude that
logŽ k H rsy1 . s 7.4 " 0.3.
In order to calculate the KIE k H rk D , the quan-
tum dynamics was recalculated for the deuteron,
using the same procedure as for the proton. A rate
constant k D of 0.62 = 10 7 sy1 was found. We can
thus conclude that the calculated KIE is 2.4r0.62
s 3.9. Within the statistical simulation error this is
in excellent agreement with the experimentally
observed value of 3.2 " 0.1.2

Overall Reaction Rate

FIGURE 3. The top graph shows the normalized Equation Ž1. for the overall rate of the reaction
probability density of the energy splitting w ( ⌬ E ), consists of a proton transfer rate k H from the
calculated from the 100-ps MD simulation. The thick line biased state, multiplied by a probability factor
is a Gaussian approximation to this. Because the p bias for the biased state. Using 36 the values k 1 s
trajectory was saved every 10 fs, this means that the 67.2, k 2 s 9.0, kU1 s 973, kU2 s 101, and the value
graph is constructed from 10.000 samples. The bottom U
of 10.7 kcalrmol for ⌬Gmf , we find p bias s 1.12 =
graph shows the corresponding free energy profile y9
10 or log p bias s y8.95. The error in this value
⌬G ( ⌬ E ) = yk BT log w ( ⌬ E ) + k BT log w (² ⌬ E :) for both
the simulation data and the Gaussian approximation. results largely from the error in the QM terms.
The harmonic approximation has a correlation coefficient Assuming an error of 2 kcalrmol in the potential
of 0.93. The anharmonic behavior at the higher free of mean force, the error in log p bias is 1.4. The
energy points is due to the bad statistics in that area of overall rate becomes 2.7 = 10y2 sy1 or log k s
configurational space. y1.6 " 1.6. This value overestimates the experi-

892 VOL. 20, NO. 8


SIMULATION OF A SLOW QUANTUM REACTION

FIGURE 4. Trajectory-averaged evolution of the product state occupation ⌰( t ) [see eq. (3)]. The ␰ barrier was chosen
˚ Extrapolation of the initial rate of transfer leads to log k bias = 7.4 " 0.3 for the biased ester.
at 1.40 A.

mental rate of log k exp s y2.56 by an order of The occurrence of the ) state can be described
magnitude, but it still agrees within the error of by classical statistical mechanics as in transition
the calculation. state theory ŽTST., but the proton transfer has an
essential quantum character and is driven by a
fluctuating potential due to the environment. Ana-
Discussion and Conclusion lytical theories for this process exist only in the
two-state approximation.4, 31 The Gibbs free energy
We showed that it is possible to compute the of the ) state is considerably stabilized by the
reaction rate of a slow reaction where the rate- orientational polarization of the aqueous environ-
limiting step is solvent-induced proton transfer ment. Applying straightforward TST to the proton
from an activated state, using a combination of transfer Žassuming an AM1-SM1 calculated barrier
semiempirical QM, classical MD, and quantum- of 7.1 kcalrmol and a proton zero point energy of
dynamical simulations. The simulations give in- 1.75 kcalrmol as calculated from the average pro-
sight into the details of the process and serve to ton potential, and including the bias probability
illustrate the inadequacy of standard transition- p bias as computed above., we find a rate constant
state theory for such processes. The overall picture of 0.8 sy1 and a KIE of 2.4. This overestimates the
is that the system ‘‘waits’’ until a fluctuation al- experimental rate by more than two orders of
lows the occurrence of a biased state Ž). from magnitude and underestimates the KIE.
which proton transfer takes place in a nonadiabatic The DME method is quite suitable to simulate
process. This is illustrated in Fig. 5. We note that the initial rate of the reaction where the proton
the " state indicated in Fig. 5 is not relevant for interacting with its classical environment can be
the neutral hydrolysis reaction because it involves approximated as a classical particle and where
an OHy ion as an intermediate, which occurs in backreaction due to the proton quantum dynamics
fact in the base-catalyzed hydrolysis observable at onto its environment can be neglected. The ques-
higher pH.

JOURNAL OF COMPUTATIONAL CHEMISTRY 893


LENSINK, MAVRI, AND BERENDSEN

mixed solvents. Work addressing both these as-


pects Ži.e., the entropic contribution of a water
molecule residing in proximity to the ester or not,
as well as the influence of cosolvents. is in progress.
A disputable point in the present treatment is
the assumption that the ) state has a sufficient
lifetime to allow proton transfer to take place.
However, there is no need for the ) state to persist
over tens of nanoseconds, which is the average
time it takes until a transition occurs. In fact, it
should live a much shorter time for the presumed
preequilibrium between the initial and ) state to
be valid. But the ) state should persist over the
duration of an actual proton transfer, which is less
than 1 ps.19 The treatment could be refined by
incorporating the motion of the biasing coordi-
nates into the quantum dynamics, but then a bias
should be introduced into the initial quantum con-
FIGURE 5. Sketch of the free energy diagram for the ditions to make the transition observable. If this
neutral hydrolysis reaction. The simplified coordinate can be done in a reliable way, the backreaction can
RCO represents both biasing coordinates RCO and ROO; be included as well.10
ROH represents the reaction coordinate along which the While classical MD of complex systems has al-
proton transfer takes place. The concerted reaction ready come of age,37 the embedding of semiempir-
proceeds via the ) state. The " state refers to a reaction ical QM and quantum dynamics in classical MD
pathway via a hydroxyl ion, which we consider of broadens applications to reactive pathways, even
negligible importance for the neutral hydrolysis reaction. on time scales far beyond the brute-force applica-
tion of these methods.

tion of whether it is correct to use Hellmann᎐


Feynman forces from mixed states or whether sur- Acknowledgments
face hopping7 should be introduced does not come
up in this case. The complete proton transfer, which This work was performed at the Centre
occurs only once in several tens of nanoseconds, ´
Europeen ´
de Calcul Atomique et Moleculaire
cannot be directly simulated at present. Therefore, ŽCECAM. in Lyon, France. We wish to thank Gio-
it is also not yet possible to compute the barrier vanni Ciccotti and Stefano Baroni for hospitality at
recrossing ratio Ži.e., the fraction of unsuccessful CECAM. J. M. is grateful for a long-term HFSP
transfers., which is expected to reduce the rate by stipend and to the University of Groningen for the
a factor of about 2.5, 6, 35 hospitality during his stay in Groningen. We are
We made restrictions in the position of the wa- grateful to Jan Engberts, University of Groningen,
ter molecule nearest to the carbonyl, constraining for stimulating initial discussions and arousing
it to a region above the carbon atom, perpendicu- our interest in this reaction.
lar to the ester plane. This restriction was not
further investigated and was not incorporated into
the probability of the ) state. It will lead to a References
negative entropic contribution to the transition
1. Isaacs, N. S. Physical Organic Chemistry; Longman: New
state and an overestimation of the rate constant. If York, 1987.
we estimate this to be a factor of 10 Ža water 2. Ža. Engbersen, J. F. J.; Engberts, J. B. F. N. J Am Chem Soc
molecule must reside in about 1 A ˚2 of surface area 1974, 96, 1231; Žb. Engbersen, J. F. J.; Engberts, J. B. F. N.
of the 10 A ˚ available to it., its inclusion would
2
J Am Chem Soc 1974, 97, 1563; Žc. Holterman, H. A. J.;
lead to perfect but fortuitous agreement with the Engberts, J. B. F. N. J Am Chem Soc 1980, 102, 4256; Žd.
Holterman, H. A. J.; Engberts, J. B. F. N. J Org Chem 1983,
experiment. It is likely, however, that the entropic
48, 4025.
factor related to the probability that a water 3. Ža. Mezei, M.; Beveridge, D. L. Ann NY Acad Sci 1986, 582,
molecule occupies the required position forms the 1; Žb. Van Gunsteren, W. F.; Berendsen, H. J. C. Angew
basis for solvent effects on the reaction rate in Chem Int Ed Engl 1990, 29, 992.

894 VOL. 20, NO. 8


SIMULATION OF A SLOW QUANTUM REACTION

4. Ža. Berendsen, H. J. C.; Mavri, J. J Phys Chem 1993, 97, 20. Ža. Stewart, J. J. P. QCPE Bull 1990, 10, 86; Žb. Jones, J. P.
13464; Žb. Mavri, J.; Berendsen, H. J. C. J Mol Struct 1994, MOBOSOL; Departments of Chemistry and Pharmacology,
322, 1; Žc. Mavri, J.; Berendsen, H. J. C. J Phys Chem 1995, University of Rochester: Rochester, NY.
99, 12711; Žd. Berendsen, H. J. C.; Mavri, J. In Quantum 21. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Gill, P. M. W.;
Mechanical Simulation Methods for Studying Biological Johnson, B. G.; Wong, M. W.; Foresman, J. B.; Robb, M. A.;
Systems; Bicout, D.; Field, M., Eds.; Springer-Verlag: Berlin, Head᎐Gordon, M.; Replogle, E. S.; Gomperts, R.; Andres, J.
1996; p 157. L.; Raghavachari, K.; Binkley, J. S.; Gonzalez, C.; Martin, R.
5. Ža. Azzouz, H.; Borgis, D. J Chem Phys 1993, 98, 7361; Žb. L.; Fox, D. J.; Defrees, D. J.; Baker, J.; Stewart, J. J. P.; Pople,
Staib, A.; Borgis, D.; Hynes, J. T. J Chem Phys 1995, 102, J. A. Gaussian 92rDFT, Revision G.1; Gaussian Inc.: Pitts-
2487. burgh, PA, 1993.
6. Laria, D.; Ciccotti, G.; Ferrario, M.; Kapral, R. J Chem Phys 22. Van Gunsteren, W. F.; Berendsen, H. J. C. GROMOS,
1992, 97, 378. Groningen Molecular Simulation Package; Biomos B.V.:
7. Ža. Tully, J. C. J Chem Phys 1990, 93, 1061; Žb. Coker, D. F. Groningen, The Netherlands, 1987.
In Computer Simulation in Chemical Physics; Allen, M. P.; 23. Berendsen, H. J. C.; Postma, J. P. M.; Van Gunsteren, W. F.;
Tildesley, D. S., Eds.; Kluwer Academic: Boston, 1993; p Hermans, J. In Intermolecular Forces; Pullman, B., Ed.;
315; Žc. Hammes᎐Schiffer, S.; Tully, J. C. J Chem Phys 1994, Reidel: Dordrecht, 1981; p 331.
101, 4657; Žd. Coker, D. F.; Xiao, L. J Chem Phys 1995, 102, 24. Berendsen, H. J. C.; Postma, J. P. M.; DiNola, A.; Haak, J. R.
496. J Chem Phys 1984, 81, 8.
8. Ža. Topaler, M.; Makri, N. Chem Phys Lett 1993, 210, 285; 25. Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. J Comput
Žb. Makarov, D. E.; Makri, N. Phys Rev A 1993, 48, 3626. Phys 1977, 23, 327.
9. Ža. Cao, J.; Voth, G. A. J Chem Phys 1993, 99, 10070; Žb. Cao, 26. Cramer, C. J.; Truhlar, D. G. J Am Chem Soc 1994, 116,
J.; Voth, G. A. J Chem Phys 1994, 100, 5106; Žc. Lobaugh, J.; 3892.
Voth, G. A. J Chem Phys 1996, 104, 2056. 27. Bessler, B. H.; Merz, K. M. Jr.; Kollman, P. A. J Comput
10. Hammes᎐Schiffer, S.; Tully, J. C. J Chem Phys 1995, 103, Chem 1990, 11, 431.
8528. 28. Ža. Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. A.; Stewart, J.
11. Chandrasekhar, J.; Smith, S. F.; Jorgensen, W. L. J Am Chem J. P. J Am Chem Soc 1985, 107, 3902; Žb. Cramer, C. J.;
Soc 1985, 107, 154. Truhlar, D. G. J Comput Aided Mol Design 1992 6, 629.
12. Ža. Van Duijnen, P. T.; Thole, B. T.; Broer, R.; Nieuwpoort, 29. Berendsen, H. J. C.; Mavri, J. Int J Quantum Chem 1996, 57,
W. C. Int J Quantum Chem 1980, 17, 651; Žb. Rullmann, J. 975.
A. C.; Bellido, M. N.; Van Duijnen, P. T. J Mol Biol 1989, 30. This is the ZBR Žzero backreaction. approximation: the force
206, 101. exerted by the quantum subsystem on the classical environ-
13. Ža. Warshel, A. J Phys Chem 1982, 86, 2218; Žb. Warshel, A.; ment is independent of the quantum state. See Xiao, L.;
Chu, Z. T. J Chem Phys 1990, 93, 4003. Coker, D. F. J Chem Phys 1995, 102, 1107.
14. Maraver, J. J.; Marcos, E. S.; Bertran, J. J Chem Soc Perkin 31. Borgis, D.; Hynes, J. T. J Chem Phys 1991, 94, 3619.
Trans 1986, 8, 1323. 32. The proton was considered at R OH distances of about 1.0
15. Tomasi, J.; Persico, M. Chem Rev 1994, 94, 2027. ˚ the exact coordinates depending on the location
and 1.5 A,
16. Lu, D.; Voth, G. A. J Am Chem Soc 1998, 120, 4006. of the reactant and product wells.
17. Mavri, J.; Berendsen, H. J. C.; Van Gunsteren, W. F. J Phys 33. Marcus, R. A.; Sutin, N. Biochim Biophys Acta 1985, 811,
Chem 1993, 97, 13469. 265.
18. Mavri, J.; Berendsen, H. J. C. J Phys Chem 1995, 99, 12711. 34. Ža. Kuharski, R. A.; Bader, J. S.; Chandler, D.; Sprik, M.;
Klein, M. L.; Impey, R. W. J Chem Phys 1988, 89, 3248; Žb.
19. Ža. Bala, P.; Grochowski, P.; Lesyng, B.; McCammon, J. A.
Borgis, D.; Tarjus, G.; Azzouz, H. J Phys Chem 1992, 96,
J Phys Chem 1996, 100, 2535; Žb. Bala, P.; Grochowski, P.;
3188.
Lesyng, B.; McCammon, J. A. In Quantum Mechanical Sim-
ulation Methods for Studying Biological Systems; Bicout, 35. Consta, S.; Kapral, R. J Chem Phys 1996, 104, 4581.
D.; Field, M., Eds.; Springer-Verlag: Berlin, 1996; p 119; Žc. 36. The values of k 1 and k 2 are immaterial, because they cancel
Grochowski, P.; Lesyng, B.; Bala, P.; McCammon, J. A. Int J in eqs. Ž1. and Ž2..
Quantum Chem 1996, 60, 1143. 37. Berendsen, H. J. C. Science 1996, 271, 954.

JOURNAL OF COMPUTATIONAL CHEMISTRY 895

You might also like