1:
2: \documentstyle[aps,overcite,graphics,multicol,array,epic,eepic,amsmath]{revtex}
3:
4: \setlength{\topmargin}{0.01in}
5: %\renewcommand{\baselinestretch}{2.0}
6: \begin{document}
7:
8: \def\M{{\bf M}}
9: \def\A{{\bf A}}
10: \def\B{{\bf B}}
11: \def\F{{\bf F}}
12: \def\X{{\bf X}}
13: \def\H{{\bf H}}
14: \def\R{{\bf R}}
15: \def\S{{\bf S}}
16: \def\U{{\bf U}}
17: \def\u{{\bf u}}
18: \def\t{{\bf t}}
19: \def\x{{\bf x}}
20: \def\ab{{\bf a}}
21: \def\bb{{\bf b}}
22: \def\eb{{\bf e}}
23: \def\fb{{\bf f}}
24: \def\gb{{\bf g}}
25: \def\lb{{\bf l}}
26: \def\yb{{\bf y}}
27: \def\zb{{\bf z}}
28: \def\db{{\bf d}}
29: \def\zero{{\mbox{\bf $0$}}}
30: \def\kk{{\bf k}}
31: \def\vv{{\bf v}}
32: \def\qq{{\bf q}}
33: \def\rr{{\bf r}}
34: \def\pp{{\bf p}}
35: \def\P{{\bf P}}
36: \def\E{{\bf E}}
37: \def\I{{\bf I}}
38: \def\J{{\bf J}}
39: \def\T{{\bf T}}
40: \def\Pt{{\tilde{P}}}
41: \def\Qt{{\tilde{Q}}}
42: \def\Li{{\cal{L}}}
43: \def\gradrn{{\nabla_{{\bf{R}}^{N}}}}
44: \def\gradrj{{\nabla_{{\bf{R}}_{j}}}}
45: \def\gradpn{{\nabla_{{\bf{P}}^{N}}}}
46: \def\gradpj{{\nabla_{{\bf{P}}_{j}}}}
47: \def\odagger{{\cal O}^{\dagger}}
48: \def\btau{\mathbf{\tau}}
49: \def\bphi{\mbox{\boldmath{$\phi$}}}
50: \def\btheta{ \mbox{\boldmath{$\theta$}} }
51: \def\hP{{\hat{P}}}
52: \def\hp{{\hat{p}}}
53: \def\hq{{\hat{q}}}
54: \def\hQ{{\hat{Q}}}
55: \def\hB{{\hat{B}}}
56: \def\hH{{\hat{H}}}
57: \def\Lo{{{\cal L}}}
58: \def\ap{{\alpha^{\prime}}}
59: \def\bp{{\beta^{\prime}}}
60: \def\a{{\alpha}}
61: \def\b{{\beta}}
62: \def\at{{\tilde{\alpha}}}
63: \def\atp{{\tilde{\alpha}^{\prime}}}
64: \def\bt{{\tilde{\beta}}}
65: \def\btp{{\tilde{\beta}^{\prime}}}
66: \def\mut{{\tilde{\mu}}}
67: \def\nut{{\tilde{\nu}}}
68: \def\hmu{{\hat{\mu}}}
69: \def\hnu{{\hat{\nu}}}
70:
71:
72: \def\M{{\bf M}}
73: \def\A{{\bf A}}
74: \def\B{{\bf B}}
75: \def\F{{\bf F}}
76: \def\X{{\bf X}}
77: \def\H{{\bf H}}
78: \def\R{{\bf R}}
79: \def\S{{\bf S}}
80: \def\U{{\bf U}}
81: \def\u{{\bf u}}
82: \def\t{{\bf t}}
83: \def\x{{\bf x}}
84: \def\ab{{\bf a}}
85: \def\bb{{\bf b}}
86: \def\eb{{\bf e}}
87: \def\fb{{\bf f}}
88: \def\gb{{\bf g}}
89: \def\lb{{\bf l}}
90: \def\yb{{\bf y}}
91: \def\zb{{\bf z}}
92: \def\db{{\bf d}}
93: \def\zero{{\mbox{\bf $0$}}}
94: \def\kk{{\bf k}}
95: \def\vv{{\bf v}}
96: \def\qq{{\bf q}}
97: \def\rr{{\bf r}}
98: \def\pp{{\bf p}}
99: \def\P{{\bf P}}
100: \def\E{{\bf E}}
101: \def\I{{\bf I}}
102: \def\J{{\bf J}}
103: \def\T{{\bf T}}
104: \def\Pt{{\tilde{P}}}
105: \def\Qt{{\tilde{Q}}}
106: \def\Li{{\cal{L}}}
107: \def\gradrn{{\nabla_{{\bf{R}}^{N}}}}
108: \def\gradrj{{\nabla_{{\bf{R}}_{j}}}}
109: \def\gradpn{{\nabla_{{\bf{P}}^{N}}}}
110: \def\gradpj{{\nabla_{{\bf{P}}_{j}}}}
111: \def\odagger{{\cal O}^{\dagger}}
112: \def\btau{\mathbf{\tau}}
113: \def\bphi{\mbox{\boldmath{$\phi$}}}
114: \def\btheta{ \mbox{\boldmath{$\theta$}} }
115: \def\hP{{\hat{P}}}
116: \def\hp{{\hat{p}}}
117: \def\hq{{\hat{q}}}
118: \def\hQ{{\hat{Q}}}
119: \def\hB{{\hat{B}}}
120: \def\hH{{\hat{H}}}
121: \def\Lo{{{\cal L}}}
122: \def\ap{{\alpha^{\prime}}}
123: \def\bp{{\beta^{\prime}}}
124: \def\a{{\alpha}}
125: \def\b{{\beta}}
126: \def\at{{\tilde{\alpha}}}
127: \def\atp{{\tilde{\alpha}^{\prime}}}
128: \def\bt{{\tilde{\beta}}}
129: \def\btp{{\tilde{\beta}^{\prime}}}
130: \def\mut{{\tilde{\mu}}}
131: \def\nut{{\tilde{\nu}}}
132: \def\hmu{{\hat{\mu}}}
133: \def\hnu{{\hat{\nu}}}
134:
135:
136:
137:
138: \title{An efficient Monte Carlo method for calculating ab initio
139: transition state theory reaction rates in solution}
140: \author{Radu Iftimie${}^{1,3}$, Dennis Salahub${}^{2}$ and Jeremy Schofield${}^{1}$}
141: \address{1. Chemical Physics Theory Group,
142: Department of Chemistry, University of Toronto,
143: Toronto, Ontario, Canada M5S 3H6}
144: \address{2. University of Calgary,
145: 2500 University Drive N.W.,
146: Calgary, Alberta, Canada T2N 1N4}
147: \address{3. New York University, Department of Chemistry,
148: 100 Washington Sq. E., New York, NY, 10003
149: }
150:
151: \date{\today}
152: \maketitle
153: \begin{abstract}
154: In this article, we propose an efficient method for sampling the
155: relevant state space in condensed phase reactions. In the present method,
156: the reaction is described by solving
157: the electronic Schr\"{o}dinger equation for the solute atoms in the presence
158: of explicit solvent molecules. The sampling algorithm uses a molecular mechanics
159: guiding potential in combination with simulated tempering ideas and allows thorough
160: exploration of the solvent state space in the context of an ab initio
161: calculation even when the dielectric
162: relaxation time of the solvent is long. The method is applied to the study of
163: the double proton transfer reaction that takes place between a molecule of acetic
164: acid and a molecule of methanol in tetrahydrofuran.
165: It is demonstrated that calculations
166: of rates of chemical transformations occurring in solvents of medium
167: polarity can be performed with an increase in the cpu time of factors ranging from
168: 4 to 15 with respect to gas-phase calculations.
169:
170: \end{abstract}
171:
172: \section{Introduction}
173:
174:
175: The concept of reaction mechanism plays a major role in chemistry
176: and represents a synthesis of our understanding of the way in which
177: different topological changes in the bonding structure of a reactant
178: or product are correlated as the reaction proceeds.
179: Recent advances in ultrafast lasers\cite{FlemingPhysToday90,ZewailExperimental},
180: X-ray\cite{PermanScience98} and other spectroscopies as well as in
181: computational chemistry\cite{RosskyNature94,ClaryTheoreticalGasPhase}
182: have made possible the study of most gas-phase and some condensed-phase
183: reactions in molecular detail. However, most experimental investigations
184: of complex reaction mechanisms taking place in liquid environments
185: are still inferred from isotope and solvent (medium) effects on the
186: reaction rate\cite{KlinmanBiophys.,PageWilliams}. Consequently, the
187: interpretation of the experimental results as well as the reaction
188: mechanisms inferred are more controversial than those of gas-phase
189: reactions.
190:
191: Computer studies can be useful as a complement to experimental data
192: in cases where experiments alone cannot provide a definitive picture
193: of the mechanism of the chemical process. It is therefore desirable to
194: develop systematic computational approaches to carefully examine the relation between
195: isotope effects and reaction mechanism in condensed phase systems.
196: However, computational calculations of kinetic isotope effects in
197: condensed-phase reactions can become expensive due to a number of
198: difficulties. Some of the practical challenges involving the calculation
199: of kinetic isotope effects and reaction rates in solution are
200: associated with the fact that accurate descriptions of transformations in which chemical
201: bonds are broken and formed require time-consuming \emph{ab initio}
202: electronic structure methods. Computer time limitations become particularly
203: relevant when investigating ``rare events'' such as
204: chemical reactions, especially when the reactions are accompanied
205: by substantial differences in the structure of the solvent. To further
206: complicate matters, quantum effects such as zero-point vibrations and tunneling
207: effects are important in some chemical processes, such as proton transfer reactions.
208: Another technical problem in the simulations of reactive systems is
209: that the statistical resolution
210: of calculations of the reaction rate depends on how many statistically
211: independent configurations are obtained during the simulation:
212: Simulations in which a large number of successive configurations have
213: similar configurations of the reactive core or of the solvent
214: molecules suffer from large uncertainties in the
215: calculated reaction rates, precluding any definitive interpretation of
216: the reaction mechanism.
217:
218: It is therefore critical to develop methods which sample statistically
219: independent configurations along the reaction path rapidly and correctly.
220: In the case where the reaction path can be characterized by means
221: of a small number of reaction coordinates,
222: accurate, statistically well-resolved calculations of reaction rates
223: can be performed by developing improved methods for computing reaction
224: free energy profiles along these reaction coordinates.
225: A number of techniques for computing free energy profiles along reaction
226: coordinates have been proposed in the literature, including umbrella
227: sampling\cite{UmbrellaSampling}, thermodynamic integration in conjunction
228: with the blue-moon ensemble method\cite{BlueMoonMethod}, projection
229: methods\cite{ProjectionMethods}, variable transformation approaches\cite{VariableTransform}
230: and guiding potentials\cite{Iftimie1,Rothlisberger2000}. The use
231: of molecular mechanics-based guiding potentials was proposed simultaneously
232: and independently by Iftimie et al.\cite{Iftimie1} and Vondele et
233: al.\cite{Rothlisberger2000} and implemented in a Monte Carlo and
234: a molecular dynamics framework, respectively. The basic idea of the
235: method consists of using a ``fast'' \emph{molecular
236: mechanics} potential to guide a computationally intensive \emph{ab
237: initio} simulation.
238:
239: The Monte Carlo version of the ``guiding'' approach
240: in reference {[}\citeonline{Iftimie1}{]} was called the molecular
241: mechanics based importance function method (MMBIF). It was demonstrated
242: that the utilization of a reasonably accurate molecular mechanics
243: potential as an importance function decreases the correlation of an
244: \emph{ab initio} Monte Carlo calculation by two orders of magnitude.
245: The method was illustrated on a gas-phase formic acid-water system
246: in which the activated processes involved breaking and forming hydrogen
247: bonds\cite{Iftimie1}, and was successfully applied to calculate the kinetic isotope
248: effects in a model gas-phase intramolecular proton transfer
249: reaction\cite{Iftimie2, Iftimie3, Iftimie4}.
250:
251: One of the major challenges in \emph{ab initio} simulations of reactions in condensed
252: phase environments is to thoroughly sample configurations of the
253: system when changes in the solvent occur on long time scales. For
254: instance, in molecular dynamics simulations of proton transfer reactions in which the
255: collective behavior of the solvent can strongly influence the dynamics
256: of the reaction, the sampling efficiency can be limited by long
257: solvent dielectric relaxation time. In essence, an independent
258: configuration of the system requires that the equations of motion be
259: propagated for a time which is longer than the dielectric relaxation
260: time. Even simple organic solvents such as tetrahydrofuran, the
261: relevant solvent in this study, have dielectric relaxation times on
262: the order of $4$ ps\cite{relaxationTHF}, so that independent solvent
263: configurations are only obtained after several thousand elementary
264: propagation steps. More structured solvents such as water, with a
265: dielectric relaxation time of roughly $8.3$ ps\cite{relaxationWater},
266: require even longer propagation for proper sampling of solvent configurations.
267: The long time scale of structural rearrangements in solvents pose a
268: serious challenge to \emph{ab initio} calculations even when the solvent is modeled using molecular
269: mechanics since each propagation step in the dynamics involves a
270: time-consuming \emph{ab initio} calculation. Ideally, successive
271: configurations in a simulation involve drastically different solvent
272: and solute configurations. This is only possible using an artificial
273: dynamics to generate the sequence of configurations. One way to generate
274: relatively uncorrelated successive configurations is to apply
275: importance sampling ideas.
276:
277: In this paper, the molecular mechanics-based importance sampling
278: method is adapted to calculate reaction rates of chemical processes in
279: condensed phases where collective motions of the environment can
280: influence the quantitative features of the chemical
281: process and, in some cases, play a critical role in determining the
282: mechanism of a reaction. The Monte Carlo procedure involves separating the task of
283: sampling the configurations of the condensed phase system into two
284: parts. The first part involves an efficient scheme of
285: updating the solvent configuration while the second focuses on the
286: relatively slow \emph{ab initio} calculation of the reactive core. This
287: approach allows extensive sampling of the molecular mechanical solvent
288: without a significant increase in the overall computational work over
289: a gas phase \emph{ab initio} simulation. The method is applied
290: to study the double proton transfer reaction in an acetic
291: acid-methanol complex solvated by tetrahydrofuran.
292:
293: \section{Methodology}
294: \subsection{Motivation for the Model System}
295: The computational study of proton-transfer
296: reactions can be used to understand the conditions for the validity of
297: a well-known conjecture proposed
298: in the physical organic chemistry literature, stating that the breakdown
299: of the rule of geometric mean for kinetic isotope effects, which is a relation\cite{KohenJACS2002} involving
300: ratios of kinetic isotope effects corresponding to different isotopic
301: substitutions at the primary and secondary atoms, is a signature
302: of tunneling for both primary and secondary atoms\cite{KlinmanBiophys.,LimbachBook}.
303: The consequences of applying the rule of geometric mean to interpret
304: experimentally determined isotope effects are of far-reaching importance:
305: The inferred relationship between the rule of geometric mean and reaction
306: mechanism forms the basis for the proposal that \emph{multiple} \emph{intramolecular}
307: proton transfer reactions are likely to proceed via a two-step mechanism,
308: in contrast to \emph{multiple intermolecular} proton transfers which
309: are believed to proceed via a synchronous pathway\cite{LimbachBook}.
310: The same relationship is at the heart of the recent suggestion that
311: tunneling effects have played an important role in the design of the
312: active sites of some proteins\cite{KlinmanBiophys.,KlinmanJACS2001}.
313:
314: Some of the most striking consequences of the rule of geometric
315: mean\cite{LimbachBook,KlinmanBiophys.}
316: appear when studying multiple proton transfer reactions in condensed
317: phases. The study of reactions involving
318: the exchange of a pair of protons between two molecules may provide
319: insight into the dynamics of certain types of enzymatic reactions
320: in which several functional groups in the active center are properly
321: aligned so that concerted catalysis can occur. This type of catalysis
322: mechanism is called \emph{bifunctional catalysis} and is the principal
323: mechanism responsible for the several orders of magnitude increase
324: in the reaction rate in several important biochemical transformations
325: \cite{Dugas}.
326:
327: The double proton transfer reaction between acetic
328: acid and methanol is one of the simplest examples of reactions involving
329: an intermolecular exchange of protons between two molecules and therefore
330: is a good candidate for computationally investigating general aspects
331: of bifunctional catalysis. The chemical processes occurring in a
332: solution of acetic acid-methanol in tetrahydrofuran (THF) have been
333: studied experimentally by Gerritzen and Limbach\cite{Limbach84}. The majority species in the
334: system consist of complexes formed from either a single molecule of acetic acid
335: or a single molecule of methanol hydrogen-bonded with a single molecule of solvent.
336: The minority species in the system consist of linear and
337: cyclic clusters of acetic acid hydrogen-bonded with methanol
338: and solvated by tetrahydrofuran. The double-proton transfer reaction
339: takes place along the hydrogen bonds of the cyclic clusters. When
340: the concentrations of acetic acid and methanol are reduced, the only
341: cyclic cluster formed within which double-proton transfer is experimentally
342: observed is formed from one molecule of acetic acid and a single molecule of
343: methanol\cite{Limbach84} (see Fig. 1).
344:
345: In this work, we will focus only on
346: the intermolecular double proton transfer
347: which takes place in the cyclic cluster formed from one molecule of
348: methanol and a single molecule of acetic acid in a solution of
349: tetrahydrofuran (THF). The quantum nuclear effects due to the small mass
350: of the hydrogen atoms will be neglected here, and we will concentrate only
351: on the sampling issues. Once the sampling method is tested for classical atoms,
352: the method will be developed in a later article to include nuclear quantum
353: effects via centroid transition state theory using sampling ideas similar to those
354: used in our previous studies of malonaldehyde\cite{Iftimie2, Iftimie3}.
355:
356:
357: \subsection{The QM/MM/Implicit solvent approach}
358: Since it is impractical to treat large systems quantum mechanically,
359: one is inevitably faced with the decision of how to combine \emph{ab initio}
360: electronic structure methods with more empirical approaches. One
361: alternative to full \emph{ab initio} calculations consists of using a mixed
362: quantum-mechanical/molecular-mechanical (QM/MM) description of the
363: system. However, in
364: general it is problematic to clearly define the physics of the
365: interface region between the quantum and molecular mechanical subsystems.
366: In the best of circumstances, the separation can be made without
367: ``cutting'' a covalent bond. For reactive processes in which the
368: reaction occurs in a small solute which can be simulated in isolation using \emph{ab initio} methods,
369: the logical separation between the quantum and molecular mechanical
370: regions is at the solute-solvent level. That is, the energies of the solute in which
371: the reaction is occurring are calculated using electronic structure
372: methods for the solute in the presence of the solvent, while the solvent energies are
373: described by molecular mechanical potentials.
374:
375: Even after this separation has been defined, one is faced with another
376: technical issue of how an infinite condensed phase system can be represented in a
377: practical fashion. One common approach utilized in simulations of
378: condensed phase systems is to simulate explicitly a system consisting
379: of a solute surrounded by a small number of solvent molecules
380: periodically replicated in an infinite fashion. Unfortunately, such a
381: periodic replication of the system introduces artifacts. One such artifact
382: is the existence of spurious long-ranged correlations due to the
383: periodicity of the system. In principle, such correlations could have
384: a large impact on the reactive process. The use of periodic boundary
385: conditions also introduces complexity in the electronic structure
386: calculation itself when localized basis sets are used. An alternative
387: approach is to use toroidal boundary conditions in which part of the
388: solvent is explicitly represented while the influence of the solvent
389: from regions far from the solute is implicitly incorporated.
390:
391: In the present study we have utilized a quantum mechanics/molecular
392: mechanics/implicit solvent (QM/MM/continuum) approach, in which the
393: bond breaking and forming processes taking place in the solute are described
394: by including all solute electron-electron, electron-nuclei and nuclei-nuclei
395: interactions in an electronic Hamiltonian. The interactions between
396: the solvent molecules which form the first few solvation shells (in
397: practice, $343$ THF molecules) are
398: described using a molecular mechanics potential. The electrostatic
399: interactions between the solute and the solvent molecules in the first
400: few solvation shells are incorporated into this approach by solving
401: the \emph{ab initio} electronic structure equations for the solute
402: in the presence of the electric field generated by the molecular mechanics
403: solvent charges. The total energy of the system also includes the
404: Lennard-Jones interaction energy between solute and the solvent molecules
405: in the first solvation shells, and the reaction field energy which
406: accounts for the long-range electrostatic interactions between solute
407: and solvent molecules (see Fig. 2).
408:
409: The simulations were performed using the
410: toroidal boundary conditions approach\cite{Neumann}, in which
411: a cutoff distance is used that sets the boundaries of the region within which
412: intermolecular interactions are explicitly counted.
413: Therefore, the only quantum electron-electron,
414: electron-nuclei and nuclei-nuclei interactions which were
415: explicitly counted in the present treatment were
416: those which correspond to atoms separated by less than the cutoff
417: distance, here chosen to be $14$ angstroms.
418:
419: \newcommand{\D}{\displaystyle} \normalsize
420:
421: The effects of the neglected interactions in the toroidal boundary
422: conditions approach have been approximately accounted for by
423: adding reaction field corrections\cite{Neumann} to the electronic
424: energy\cite{Watts}:
425:
426: \begin{equation}
427: \label{ReactionFieldEnergy}
428: E_{RF}\approx -\sum _{i}\frac{\epsilon _{r}-1}{(2\epsilon _{r}+1)R_{c}^{3}}\mu _{i}M_{i},
429: \end{equation}
430: where the sum is over all molecules i inside the primitive cell ,
431: \( \epsilon _{r} \) is the dielectric constant of the solvent, \( R_{c} \)
432: is the radius of the spherical surface, \( \mu _{i} \) is the dipole
433: moment of molecule \( i \), and \( M_{i} \) is the total dipole
434: moment inside the spherical surface surrounding the molecule \( i \).
435:
436: The correct energetics in hydrogen-bonded systems and in systems
437: undergoing proton transfer reactions
438: is difficult to describe even with \emph{ab initio} methods. In particular,
439: DFT studies of weak hydrogen-bonding systems have proved to be particularly
440: difficult and only limited success has been achieved in predicting the geometries and energies
441: for reactant and transition state configurations on
442: the potential energy surface using most exchange-correlation
443: functionals \cite{Barone}. Care should therefore be exercised when choosing a particular
444: \emph{ab initio} method to calculate proton transfer reaction rates.
445: The non-local exchange-correlation schemes developed by Proynov, Vela and Salahub
446: \cite{Proynov} have shown particular promise for the description of hydrogen-bonded
447: systems. Sirois et al. \cite{Sirois} have demonstrated that their kinetic-energy dependent
448: exchange functionals (BLAP and PLAP) performed better than all GGA options (BP86,
449: PP86, PW91), BLYP, or other hybrid methods (B3LYP, B3PW91) on systems involving
450: intra-molecular hydrogen bonds. The predictions
451: for equilibrium and transition state geometries as well as the energetics was
452: in agreement with high-quality post-Hartree-Fock calculations {[}CCSD(T) and
453: G2{]}\cite{Sirois}. Specifically, for the gas-phase cyclic cluster of
454: acetic acid and methanol in Fig. 1, the activation energy using the
455: PLAP exchange correlation functional was found to be approximately
456: $16.4$ kcal/mol, in excellent agreement with the QCISD value of
457: $16.14$ kcal/mol\cite{Smedarchina}.
458:
459: For the simulations described in the present work, the energies of different configurations
460: were carried out using a modified version of the LCGTO-DFT program deMon-KS3.4
461: \cite{Proynov2,deMon} using the PLAP exchange-correlation functional.
462: The DFT electronic structure calculations were carried
463: out as in reference [\citeonline{Sirois}], where the application of
464: DFT electronic structure methods to hydrogen-bonding systems is
465: discussed in detail. A double-\( \zeta \) plus polarization (DZVP) orbital basis set was
466: used for all atoms and the convergence level for the SCF (self-consistent field)
467: energy using the auxiliary fitting basis sets \cite{Sirois} was $0.01$ kcal/mol.
468:
469: \subsection{The molecular mechanics potential describing the interaction between
470: the solvent molecules}
471:
472: In order to implement the QM/MM/continuum approach to compute reaction
473: rates for the double proton transfer reaction in the acetic acid-methanol
474: cluster solvated by tetrahydrofuran, a sufficiently accurate molecular
475: mechanics description of the interaction between the solvent molecules
476: is needed\cite{Iftimie1,Iftimie2}. In this work, the OPLS all atom
477: (OPLS-AA) force field of Jorgensen et al.\cite{Jorgensen} with a modified
478: electrostatic interaction term has been
479: used to describe the interactions between THF solvent molecules. The
480: modifications to the electrostatics were designed to
481: improve the gas-phase distribution of the partial charges in a THF
482: molecule as well as to improve the description of polarization effects
483: in condensed phases. Since the local electrostatic environment as
484: well as long ranged polarization effects can influence the
485: proton transfer process, it is important to properly account for the
486: permanent and induced charges in the solvent. To describe all such
487: electrostatic phenomena, all solvent molecules have been assigned
488: permanent and induced charges.
489: It can be demonstrated\cite{Thesis} using second-order perturbation
490: theory that the electrostatic interaction energy between
491: two polarizable molecules \( A \) and \( B \), each of which carries
492: a set of atomic permanent and induced charges, \( Q^{I}_{\, \textrm{p}} \)
493: and \( Q^{I}_{\, \textrm{in}} \), with \( I=1,\cdots N \)
494: corresponding to the charges on molecule $A$, and \( q^{i}_{\, \textrm{p}} \)
495: and \( q^{i}_{\, \textrm{in}} \) , where \( i=1,\cdots n \) are the
496: site charges on molecule $B$, respectively,
497: can be written to a good approximation as :
498:
499: \begin{equation}
500: \label{TotalInteractionEnergyPermanentPlusInducedText}
501: V(Q^{1}_{\, \textrm{p}},\cdots Q^{N}_{\, \textrm{p}},Q^{1}_{\, \textrm{in}},\cdots
502: Q^{N}_{\, \textrm{in}},q^{1}_{\, \textrm{p}},\cdots q^{N}_{\, \textrm{p}},q^{1}_{\, \textrm{in}},
503: \cdots q^{N}_{\, \textrm{in}})
504: \approx
505: \sum ^{N}_{I=1}\sum ^{n}_{i=1}
506: \frac{Q^{I}_{\, \textrm{p}}q^{i}_{\, \textrm{p}}}{4\pi \epsilon _{0}d_{iI}}+\frac{1}{2}
507: \sum ^{N}_{I=1}\sum ^{n}_{i=1}\left( \frac{Q^{I}_{\, \textrm{p}}q^{i}_{\, \textrm{in}}}{4\pi
508: \epsilon _{0}d_{iI}}+\frac{Q^{I}_{\, \textrm{in}}q^{i}_{\,
509: \textrm{p}}}{4\pi \epsilon _{0}d_{iI}}\right) ,
510: \end{equation}
511: where $d_{iI}$ is the distance between sites $I$ and $i$ on molecules
512: $A$ and $B$.
513: In principle, the induced charge appearing on a given solvent molecule
514: is dependent on its local environment. One simple way of
515: incorporating solvent polarization effects is to assign each solvent
516: molecule the same average induced charge in a ``mean-field''
517: fashion. More sophisticated methods of including polarization effects
518: either assign site polarizabilities or use fluctuating charges
519: distributed at specific locations on each solvent molecule\cite{RickBerne}. Such
520: methods when combined with \emph{ab initio} electronic structure methods
521: either necessitate an iterative solution of the electronic structure
522: and fluctuating charge distributions or involve dynamical methods in
523: an extended Lagrangian system\cite{Roitberg}. Unfortunately, each of these
524: approaches has shortcomings which make them impractical to implement
525: in conjunction with importance sampling Monte-Carlo methods.
526:
527: In principle, the iterative minimization of the Kohn-Sham and
528: fluctuating charge functionals can be implemented within a Monte-Carlo
529: sampling approach at the cost of additional computational effort.
530: However, provided the solvent is not very polarizable, the variation
531: of the induced charges on the solvent molecules from their mean due to
532: the presence of the solute is expected to be small. For this reason,
533: we have utilized fixed induced charges for all solvent molecules.
534: This corresponds to computing ground state energies for the reactive
535: core $E(\rho_s ({\bf{x}}))$ based on a Kohn-Sham functional\cite{HohenbergKohn,KohnSham,Levy}
536: $F[\rho_s ({\bf x})]$ which depends on the ground state electron distribution $\rho_s({\bf x})$
537: of the solute in the presence of the fixed permanent and
538: induced external solvent charges. Note that this functional includes
539: the electrostatic energy of interaction between the quantum solute and
540: the charges on the molecular-mechanical solvent molecules.
541:
542: In general, one complication must be considered when discussing
543: electrostatic interactions in mixed QM/MM systems that arises from the
544: fact that the quantum mechanical electron density can become
545: over-polarized by the molecular mechanical point charges due to the
546: absence of considerations of the Fermi repulsion between quantum and
547: molecular mechanical charges\cite{Rothlisberger}. Such effects
548: are particularly severe when using delocalized basis sets to
549: represent the quantum subsystem, but are less significant when
550: Gaussian or other local basis sets are utilized. In the present work,
551: the full Coulomb interaction potential has been used to describe
552: electrostatic interaction terms between the solute and the solvent
553: charges without any screening
554: modifications at short distances since all DFT calculations for the
555: quantum subsystem use localized basis sets.
556:
557: In our study of the acetic acid-methanol system in a THF solution,
558: the permanent charges have been assigned the value
559: \( Q^{I}_{\, \textrm{p}}=q^{I}_{\, \textrm{p}}=0.887q_{\,
560: \textrm{opls}} \), whereas the induced charges are set to a mean-field
561: value of \( Q^{I}_{\, \textrm{in}}=q^{i}_{\, \textrm{in}}=0.239q_{\textrm{opls}} \)
562: for all indices \( i \) and \( I \), where $q_{\textrm{opls}}$ are the
563: standard charges in the OPLS force field. It can be verified that the electrostatic
564: interaction energy calculated using
565: Eq.~(\ref{TotalInteractionEnergyPermanentPlusInducedText})
566: with this set of permanent and average induced charges is precisely
567: the electrostatic energy calculated using the standard set of fixed
568: OPLS charges. On the other hand, by assigning a mean induced charge
569: to each solvent molecule, the computed values for the gas-phase dipole
570: moment and condensed phase dielectric constant are in better agreement
571: with the corresponding experimental values (see Table 1).
572:
573: It should be emphasized that a less satisfactory means of describing
574: electrostatic interactions between the solute and the solvent would
575: consist of calculating the electrostatic interactions between the gas
576: phase electron distribution with the solvent charges. In such a
577: scenario, the ground state electron distribution for the solute
578: interacts with the solvent via a Coulomb interaction of the form
579: \begin{eqnarray}
580: V= -\frac{1}{4\pi \epsilon _{0}} \sum
581: ^{n}_{i=1}
582: \int \frac{\rho _{0}({\mathbf{x}}) q^{i}}{l_i(x)} d\mathbf{x},
583: \label{bareInteraction}
584: \end{eqnarray}
585: where $\rho _{0}(\mathbf{x})$ is the gas-phase electron distribution
586: and $l_i(x)$ is defined to be the distance between the ith charge
587: $q^i$ in
588: the solvent and the point $x$. This approximation corresponds to
589: calculating the solute energy in the condensed phase by a zeroth order
590: approximation for the electronic distribution of the solute, that is,
591: \begin{equation}
592: \label{InteractionSoluteSolventCorrect}
593: F\left[ \rho _{s}(\mathbf{x})\right] \approx F\left[ \rho
594: _{0}(\mathbf{x})\right] + V.
595: \end{equation}
596: Such a description neglects the fact
597: that the ground state solute electron distribution is influenced by
598: the presence of the solvent charges. The influence of the solvent
599: charges on the ground state energy can be accounted for by
600: incorporating a polarization energy of the solute by the solvent.
601:
602: Although such a crude level of description of the electrostatic
603: interactions may be incomplete, it is useful in developing importance
604: sampling Monte-Carlo schemes based on guiding potentials, such as the
605: molecular mechanics based importance function method (MMBIF) described
606: in the next section.
607: \subsection{\label{SamplingMethod}The sampling methods }
608:
609: The MMBIF method\cite{Iftimie1} consists of utilizing an auxiliary Markov
610: chain with a known asymptotic molecular mechanical distribution to propose trial
611: configurations for an \emph{ab initio} based Monte Carlo simulation.
612: In the method, each trial configuration is obtained as the last state in a classical Markov
613: chain generated from the current configuration in the \emph{ab initio} simulation
614: using various updating schemes. The proposed configurations are then accepted or rejected in the \emph{ab initio}
615: chain according to the usual Metropolis-Hastings algorithm\cite{Iftimie1}.
616: If the previous and new trial configurations in the \emph{ab initio} MC chain are
617: denoted by $\x _{old}$ and $\x _{new}$ respectively, the proposed state
618: is accepted with the probability
619: $\min \{1,\exp (-\triangle \triangle E/k_{B}T)\}$, where
620: $\Delta \Delta E$ is defined to be
621: \begin{equation}
622: \label{acceptance}
623: \Delta \Delta E=(E^{DFT}(\x _{new})-
624: E^{cl}(\x _{new}))-(E^{DFT}(\x _{old})-E^{cl}(\x _{old})),
625: \end{equation}
626: where $E^{DFT}(\x )$ and $E^{cl}(\x )$ are the potential energies
627: of configuration $\x$ calculated by \emph{ab initio}
628: methods (here density functional theory, abbreviated DFT)
629: and the classical potential, respectively, and $k_{B}$ is Boltzmann's constant.
630: It is straightforward to show that
631: this acceptance criterion guarantees that the \emph{ab initio} Markov chain
632: has the correct limiting Boltzmann distribution\cite{Iftimie1}, regardless of the number of
633: classical updates used to generate the proposed configuration.
634:
635: The efficiency of the MMBIF approach relies on constructing a
636: molecular mechanics potential for the entire system which approximates
637: the true interactions of the system at a qualitative level. At first
638: glance, the construction of a molecular mechanics potential for a
639: condensed phase system appears a daunting task given that the electron
640: distribution of the solute changes considerably during the reactive
641: process and is influenced in a complicated fashion by its local
642: environment. However, it is relatively straightforward to construct a
643: molecular mechanics potential based on simple approximate forms of the
644: interaction potentials, such as that in
645: Eq.~(\ref{InteractionSoluteSolventCorrect}). For such forms of the
646: potential, the construction of the potential is reduced to modeling
647: the reaction in gas phase and calculating effective charges on the
648: gas-phase solute which mimic the correct electron distribution. The
649: approximate form for the potential can be corrected for using
650: importance sampling methods. For example, a Monte-Carlo chain of
651: states generated using an approximate expression for the energy can be
652: manipulated by re-weighting the configurations appearing in the chain
653: by an appropriate factor\cite{Iftimie2}. The efficiency of this approach is highly
654: dependent on the quality of the approximation for the true energy of
655: the system. One might anticipate that the crude of level of
656: description of the electrostatic interactions in
657: Eq.~(\ref{InteractionSoluteSolventCorrect}) which neglects any
658: polarization effects of the solute by the solvent molecules would
659: introduce large statistical uncertainties at the re-weighting step.
660: However, the polarization of the solute by the solvent can be
661: approximated by adjusting the charge of the solvent molecules in the
662: expression for the interaction between the solute and the solvent.
663: As will be discussed, the effective charge on the solvent molecules can
664: be designed to approximate solute polarization effects and thereby
665: improve the statistical resolution of the re-weighting procedure.
666:
667: The task of constructing a molecular mechanics potential for the
668: gas-phase proton transfer reaction is facilitated by using bond
669: evolution theory considerations. Following these lines,
670: the molecular mechanics description of the acetic acid-methanol complex
671: in the absence of the solvent was created as suggested in reference {[}\citeonline{Iftimie2}{]}.
672: The total molecular mechanics energy was decomposed into two components.
673: The first component of the total energy was written as a sum of harmonic
674: potentials representing the variation of the potential with bond length,
675: bond angle or bond dihedral displacements from their minimum energy
676: values at a fixed value of the control parameter \( b \), defined by
677: \begin{equation}
678: \label{a_b_1andb_2Definitions}
679: b=d_{\textrm{O}_{7}\textrm{H}_{4}}-d_{\textrm{O}_{7}\textrm{H}_{6}},
680: \end{equation}
681: where the numbering of the atoms is that from Fig. 3.
682:
683: The second component of the total energy was written as a sum of four
684: Morse potentials depending on the four \( \textrm{O}-\textrm{H} \)
685: bond lengths, plus two effective potentials depending on two parameters,
686: \( a_{1} \) and \( a_{2} \) defined as
687: \begin{equation}
688: \label{DefinitionAs}
689: \textrm{ }a_{1}=d_{\textrm{O}_{3}\textrm{O}_{7}}\textrm{ and }a_{2}=d_{\textrm{O}_{5}\textrm{O}_{7}}.
690: \end{equation}
691: These effective potentials account for the flow of electronic charge during
692: the reaction, as well as for the Fermi repulsion between the oxygen
693: atoms \( \textrm{O}_{3} \) and \( \textrm{O}_{7} \) , and between
694: \( \textrm{O}_{5} \) and \( \textrm{O}_{7} \). This second component
695: of the total energy was implemented using the same functional forms
696: as in reference {[}\citeonline{Iftimie2}{]}. As in the case of the
697: malonaldehyde study, no potential which depends explicitly on the
698: angles \( \widehat{\textrm{O}_{3}\textrm{H}_{4}\textrm{O}_{7}} \)
699: or \( \widehat{\textrm{O}_{5}\textrm{H}_{6}\textrm{O}_{7}} \) was
700: utilized, although some dependence on the angle \( \widehat{\textrm{H}_{4}\textrm{O}_{7}\textrm{H}_{6}} \)
701: was explicitly introduced using a functional form which interpolates
702: between a harmonic potential for transition state values of the parameter
703: \( b \), and zero for values of \( b \) characteristic for the reactant
704: or product configurations. The complete details for the construction of
705: the guiding potential for the solute can be found in reference {[}\citeonline{Thesis}{]}.
706:
707: The simplest practical means of incorporating the electrostatic interactions
708: between the solute and the solvent molecules in the molecular
709: mechanics potential is to fit partial charges to atomic sites in the
710: solute to reproduce the gas-phase electronic distribution. However,
711: since the electron distribution of the solute varies appreciably with
712: the configuration of the solute, the guiding potential must incorporate
713: solute charges which vary as the reaction proceeds.
714: In their bond evolution theory analysis of the tautomerization of
715: malonaldehyde, Krokidis et al.\cite{Krokidis} found that the
716: total charge in the basins of attraction of the proton and oxygen
717: atoms varies approximately linearly with a control parameter similar
718: to the parameter \( b \).
719: Therefore, it can reasonably be assumed that most of the variation of the
720: charges on atomic sites in the solute can be explained by a linear
721: variation with \( b \).
722:
723: An alternative approach to incorporate the solute-solvent interactions
724: can be constructed using simulated tempering
725: methods\cite{MarinariParisi,Geyer1995}. The advantage of this procedure
726: is that it does not rely on any approximation for the
727: variation of the fitted charges on the solute during the reaction.
728: The method essentially consists of using an extended state space to
729: gradually turn on electrostatic interactions between the solute and
730: the solvent. This approach is represented schematically in Figure 4.
731:
732: In the simulated tempering method, a Markov chain is constructed whose
733: states (\( i,\mathbf{r} \)) are defined on the space formed by the
734: direct product between a finite set of ``electrostatic'' indices, \(
735: i=1,2,\cdots , i_{max} \)
736: and the entire solvent plus the solute configurational space. In vector notation,
737: the states \( \mathbf{r} \) will be denoted \( \mathbf{r}=(\mathbf{r}^{s},\mathbf{r}^{S}) \),
738: where \( \mathbf{r}^{s} \) and \( \mathbf{r}^{S} \) represent the
739: solute and solvent degrees of freedom, respectively. This Markov chain,
740: whose generic state is denoted (\( i,\mathbf{r} \)), is constructed
741: to produce states asymptotically distributed according to the
742: probability density
743:
744: \begin{equation}
745: \label{p(i,x)}
746: p(i,{\mathbf{r}})=w_{i}p_{i}({\mathbf{r}}),
747: \end{equation}
748: where \( w_{i} \) are constants which will be referred to as the
749: weights for the unnormalized probability densities \( p_{i}({\mathbf{r}}) \) defined by
750:
751: \begin{equation}
752: \label{p_i(x)}
753: p_{i}({\mathbf{r}}) = \exp \left( -\beta E_{i}({\mathbf{r}})\right) ,
754: \end{equation}
755: where $\beta=1/k_{B}T$.
756: The potentials \( E_{i}({\mathbf{r}}) \) contain five components: the
757: \emph{ab initio} potential \( E^{s}({\mathbf{r}}^{s}) \) calculated
758: for the gas-phase solute configuration, the molecular mechanics potential
759: \( E^{S}(\mathbf{r}^{S}) \) describing the interaction between the
760: explicit solvent atoms, the Lennard-Jones potential \( E_{LJ}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S}) \)
761: describing the dispersive and short-ranged interactions between solute
762: and solvent atoms, the Coulomb electrostatic interaction \( E_{i}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S}) \)
763: between some charges on the solute atoms and the permanent and induced
764: charges on the solvent atoms, and the reaction field energies
765: $E_{RF}({\mathbf{r}}^{s},{\mathbf{r}}^{S})$ describing
766: the long-range solute-solvent electrostatic interactions:
767:
768: \begin{equation}
769: \label{E_i(x)}
770: E_{i}({\mathbf{r}}^{s},{\mathbf{r}}^{S})=E^{s}({\mathbf{r}}^{s})+E^{S}({\mathbf{r}}^{S})
771: +E_{LJ}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S})+
772: E_{i}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S})+E_{RF}({\mathbf{r}}^{s},{\mathbf{r}}^{S}).
773: \end{equation}
774:
775: The Coulomb solute-solvent electrostatic interaction potential between
776: charges $q^s$ on the solute and $q^S$ on the solvent, given by
777: \begin{equation}
778: E_{i}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S}) =
779: \lambda_i \sum^{N}_{J=1}\sum ^{n}_{j=1}
780: \frac{q_{J}^{S} q_{j}^{s}}{4\pi \epsilon _{0}d_{Jj}},
781: \label{electroSoluteSolvent}
782: \end{equation}
783: where $n$ and $N$ are the number of charge sites on the solute and
784: solvent molecules, respectively, and $\lambda_i$ is a scaling factor
785: henceforth called the charge fraction. Note that this interaction potential
786: depends on the choice of charges assigned to the solvent and solute
787: molecules. In our calculations, the charges on the solute were fitted using the
788: Kollman-Singh procedure\cite{KollmanSingh} from the gas-phase electron
789: distribution of the solute. The use of charges fitted from the
790: gas-phase calculation of the electron distribution does not account
791: for the polarization of the solute by the solvent. To partially compensate for the
792: neglect of this effect, the total charge on the solvent molecules used
793: in Eq.~(\ref{electroSoluteSolvent}) is set to the sum of the permanent
794: and induced charges on the solvent, $q=q_p+q_{in}$. This is in contrast
795: to the calculation of the electrostatic interaction between solvent
796: molecules in which the effective charges are set to $q=q_p+q_{in} /2$.
797: This approximation is consistent with first-order perturbation theory
798: in which the polarization energy of the solute by the solvent is
799: approximated by the polarization energy of the solvent by the
800: solute\cite{Thesis}.
801:
802: Note that the only difference between the potentials \( E_{i}({\mathbf{r}}^{s},{\mathbf{r}}^{S}) \)
803: for different electrostatic indices \( i \) consists of the form of the electrostatic interaction
804: energy \( E_{i}^{sS}({\mathbf{r}}^{s},{\mathbf{r}}^{S}) \) between the
805: solute and the explicit solvent atoms.
806: The energy \( E^{sS}_{1}({\mathbf{r}}) \) is chosen to be zero in this
807: work implying that $\lambda_1=0$,
808: which amounts to turning off solute charges. For this system all
809: electrostatic interactions between the solvent and the solute are
810: scaled to zero although the solvent and solute still interact through
811: Lennard-Jones potentials. By design, the last electrostatic index
812: is set to unity, $\lambda_{i_{max}}=1$ so that
813: the energy \( E^{sS}_{i_{max}}({\mathbf{r}}) \) corresponds
814: to calculating the electrostatic interaction
815: between the solvent and solute atoms using the fitted configuration-dependent
816: solute charges obtained \emph{via} the Kollman-Singh method. The additional dimension
817: of the solute plus solvent state space represented by the
818: electrostatic index \( i=1,\cdots ,i_{max} \)
819: ensures that the solvent configuration adapts smoothly to the solute
820: via a stepwise process in which the charges of the solute atoms
821: gradually interact with the other charges in the system. It should be
822: noted that an equivalent result can be achieved by using a parallel
823: tempering scheme in which the label \( i=1,\cdots ,i_{max} \) corresponds
824: to a stepwise decrease of a ``sampling temperature''
825: associated with the solute-solvent electrostatic interaction.
826:
827: In our implementation of the importance sampling, a Markov chain of
828: extended state space configurations was generated by two types of
829: transitions. In transitions of the first type the electrostatic index
830: \( i \) was kept fixed while the configuration of the system
831: \textbf{\( {\mathbf{r}}=({\mathbf{r}}^{s},{\mathbf{r}}^{S}) \)} was
832: updated using a transition matrix which leaves \( p_{i}({\mathbf{r}})
833: \) invariant. The way in which the configurations were updated for
834: fixed electrostatic index was dependent on the electrostatic scaling
835: factor. When the scaling factor was zero both the configuration of
836: the solute and the solvent were updated simultaneously using the MMBIF
837: approach. The background molecular mechanics simulations used to
838: guide the updates were run so as to produce effectively independent
839: but energetically reasonable configurations of the entire system.
840: The simultaneous update of both solute and solvent configuration is
841: possible in the absence of electrostatic interactions between the
842: solute and solvent since the fitted charges are
843: not used in the molecular mechanics auxiliary chain.
844: However, after the calculation of the \emph{ab initio} DFT energy
845: in the acceptance-rejection step of the MMBIF method,
846: the fitted charges for the given solute configuration
847: can be calculated. When the electrostatic scaling factor is nonzero
848: and solute-solvent electrostatic interactions occur, the solvent was
849: allowed to adjust to the charge distribution on the solute by updating
850: the solvent configuration while maintaining the configuration of the
851: core reactive region unchanged.
852:
853: Transitions of the second type move the system through the auxiliary
854: parameter space by applying a transition matrix that changes the
855: electrostatic index while leaving the configuration of the solute and
856: solvent unchanged. In this study, the method for
857: updating the electrostatic index \( i \) consisted
858: of using a Metropolis algorithm, with a proposal distribution in which
859: the proposed indices \( i_{\textrm{new}}=i_{\textrm{old}}+1 \) and \( i_{\textrm{new}}=i_{\textrm{old}}-1 \)
860: are equally probable. The proposed change of electrostatic index is rejected if \( i_{\textrm{new}} \)
861: is outside the valid range \( i=1,\cdots ,i_{max} \), otherwise it is accepted
862: with probability
863:
864: \begin{equation}
865: \label{AcceptanceCriterion}
866: \min \left[ 1,\frac{p\left( i_{\textrm{new}},{\mathbf{r}}\right) }{p\left( i_{\textrm{old}},{\mathbf{r}}\right)
867: }\right] =\min \left[ 1,\frac{w_{i_{\textrm{new}}}}{\textrm{w}_{i_{\textrm{old}}}}
868: \exp \left( -\beta \left( E_{i_{\textrm{new}}}-E_{i_{\textrm{old}}}\right) \right) \right] .
869: \end{equation}
870: The marginal distribution of \( i \) with respect to the equilibrium
871: distribution is given by
872:
873: \begin{equation}
874: \label{Choicew}
875: p(i)=\int p(i,{\mathbf{r}})d{\mathbf{r}} = \int w_{i}p_{i}({\mathbf{r}})d{\mathbf{r}}=w_{i}Z_{i},
876: \end{equation}
877: where the configuration partition function \( Z_{i} \) is defined
878: as
879:
880: \begin{equation}
881: \label{Partition}
882: Z_{i}=\int \exp \left( -\beta E_{i}({\mathbf{r}})\right) d{\mathbf{r}}.
883: \end{equation}
884: If the distributions \( p(i,{\mathbf{r}}) \) are all to play a useful
885: role in sampling, the weights \( w_{i} \) should be chosen such that
886: a roughly uniform distribution over \( i \) is obtained. Since the
887: \( Z_{i} \) are initially unknown, suitable values for the weights
888: are found through a trial and error process using preliminary runs.
889: To do this, an iterative procedure can be used in which the Markov
890: chain is simulated using the current values for \( w_{i} \), and
891: the frequencies \( f_{i} \) with which each distribution is visited
892: are recorded.
893:
894: Next, new and improved weights \( w_{i'} \) are calculated as \( w_{i'}=w_{i}/f_{i} \)
895: for electrostatic index \( i \). If some of the frequencies
896: \( f_{i} \) are zero, various elaborations of the estimation procedure
897: can be used and some of them are summarized in reference {[}\citeonline{Geyer1995}{]}.
898: The number \( i_{max} \) of values of the electrostatic scaling
899: parameter $\lambda_i$ used and the actual values of \( w_{i} \),
900: \( i=1,\cdots ,i_{max} \) are chosen by minimizing the average computer
901: time necessary for a new solute configuration to appear with an
902: electrostatic scaling parameter $\lambda_{i_{max}}=1$. A good
903: starting estimate for the numbers \( w_{i} \) and
904: \( i_{max} \) can be obtained by optimizing the weights and the number
905: of intermediate chains using only the molecular mechanics guiding
906: potential with a reasonable set of atomic charges on the solute atoms.
907:
908:
909: \section{Results }
910:
911: To explore the issue of how importance sampling methods can
912: be effectively utilized in simulations of reactions in condensed
913: phases, two different implementations of the MMBIF sampling method
914: were analyzed. In the first implementation, the dependence of the solute
915: charges on the reactive state of the system was taken to vary linearly
916: with the control parameter $b$ defined in
917: Eq.~(\ref{a_b_1andb_2Definitions}). For this simulation, the
918: electrostatic interactions between the interpolated charges on the
919: solute and the charges in the solvent were taken to be Coulombic. In
920: the second simulation, the simulated tempering method described above
921: was applied to the solvated acetic acid-methanol cluster. Both
922: simulations generate chains of states asymptotically distributed
923: according to a Boltzmann distribution based on Eq.~(\ref{InteractionSoluteSolventCorrect}) in which
924: the solute-solvent electrostatic interactions are modeled by
925: calculating the Coulomb interaction between the gas phase electron
926: distribution with the charges in the solvent. In both simulations, the desired distribution
927: for the chain of states based on the universal Kohn-Sham functional
928: $F[\rho_s (x)]$ can be recovered at the end of the simulation by
929: re-weighting each of the $N_T$ total configurations by a configuration dependent factor
930: \begin{equation}
931: \label{reweight}
932: W (x_i) = \frac{ e^{-\beta \Delta E_{pol}(x_i)}}{\displaystyle{\sum_{i=1}^{N_T}e^{-\beta \Delta E_{pol}(x_i)}}},
933: \end{equation}
934: where
935: \begin{equation}
936: \label{DeltaEpol}
937: \Delta E_{pol} (x_i) = F[\rho_s(x_i)] - \left( F[\rho_0 (x_i)]+V
938: \right) = F[\rho_s(x_i)]-E^{s}(r_{i}^{s}) - E^{sS}(r_{i}^{s},r_{i}^{S})
939: \end{equation}
940: is the difference in the polarization energy of the solute by the
941: solvent estimated by calculating the energy of the ground state
942: electron distribution in the presence of the solvent charges and the
943: energy of a gas-phase electron distribution interacting with the
944: optimized solvent charges.
945:
946: In all simulations, the calculations for
947: the solvated cluster (shown in Fig. 3)
948: have been carried out by treating all nuclei as classical point particles.
949: The calculations were conducted in the isobaric-isothermal ensemble at \( p=1 \) atm and \( T=298 \) K.
950: In order to improve the sampling along the reaction coordinate, an
951: umbrella potential\cite{UmbrellaSampling} was constructed for the
952: guiding potential using a self-adaptive scheme. Simulations biased by
953: the converged umbrella potential yielded a uniform sampling of the
954: important regions of the reaction coordinate even though the
955: activation barriers for the proton transfer reaction was on the order
956: of $28$ $k_{B}T$.
957:
958: Two thirds of the \emph{ab initio} Markov chain transition steps in
959: the simulation using the linearly-interpolated charges on the core were generated using the
960: MMBIF method. The same fraction of base transitions were generated
961: using the MMBIF updates when the electrostatic index is zero in the
962: simulated tempering simulation (see
963: Fig. 4). The rest of the \emph{ab initio} Markov chain transitions
964: were performed utilizing Metropolis single-variable updates using
965: the \emph{ab initio} DFT energy. As demonstrated in a previous study
966: {[}\citeonline{Iftimie1}{]}, the role of the Metropolis DFT updates
967: is to prevent the guiding molecular mechanics Markov chain of states
968: from spending a large number of successive steps in those regions
969: of the configuration space where the molecular mechanics density of
970: states in the solute configuration space substantially underestimates
971: the corresponding \emph{ab initio} DFT density of states.
972:
973: In the previous studies of the formic acid-water\cite{Iftimie1} and
974: of the malonaldehyde\cite{Iftimie2} systems it was demonstrated that
975: a useful strategy for optimizing the MMBIF method was to separate
976: the variables to be updated in a classical MC step into several groups,
977: with strongly correlated variables grouped together. In the case of
978: the double proton transfer in the cyclic cluster formed by acetic
979: acid and methanol, the vibrations of the two methyl groups should
980: be relatively uncoupled from the motions of the other atoms in the
981: cluster. Applying this separation of which variables are updated in
982: the MMBIF method, the percent of rejections of proposed
983: configurations obtained with the MMBIF method was about \( 30\% \).
984: The percent of rejections of the transitions which employed single-variable
985: Metropolis updates using the \emph{ab initio} potential was \( 45\% \).
986:
987: It is important to compare the computational effort of performing
988: liquid-phase versus gas-phase simulations of chemical reactions
989: within the MMBIF approach.
990: The potentials of mean force obtained using the reaction coordinate
991:
992: \begin{equation}
993: \label{ReactionCoordinateMalonaldehyde2}
994: \xi =b=d_{\textrm{O}_{7}\textrm{H}_{4}}-d_{\textrm{O}_{7}\textrm{H}_{6}},
995: \end{equation}
996: describing the double proton-transfer in the acetic acid-methanol system
997: in tetrahydrofuran obtained in the first simulation and in a gas-phase
998: simulation are represented in Fig. 5.
999: Although equal cpu times were spent in computing the two potentials
1000: of mean force depicted in Fig. 5,
1001: the statistical uncertainties for the activation energy calculated for the double
1002: proton reaction rate in the solvated system is approximately \( 2 \)
1003: times larger than the error bar for the activation energy calculated
1004: from the gas-phase simulation. This decrease in the statistical resolution
1005: of the computations in the solvated system with respect to the gas-phase
1006: system is due to a larger integrated correlation time of the overall
1007: Markov chain as well as to
1008: the fact that \( 50\% \) of the cpu time in the simulation of the
1009: solvated system is dedicated to calculating the molecular mechanics
1010: interactions between the \emph{solvent} molecules. Although a single
1011: \emph{ab initio} calculation is roughly $4$ orders of magnitude slower than a
1012: single update of the configuration of the molecular mechanical
1013: solvent, the long dielectric relaxation of the solvent required that
1014: many updates be carried out on the solvent before an independent
1015: solvent configuration was generated. This translated into an
1016: approximately equal amount of CPU time for the molecular mechanical
1017: and quantum mechanical components of the simulation.
1018:
1019: The approximation of the variation of the charges fitted using the
1020: Kollman-Singh procedure with the solute configuration by a linear
1021: variation with the parameter \( b \) in Equation (\ref{a_b_1andb_2Definitions})
1022: proved quite accurate. The standard deviation of the differences between
1023: the electrostatic interaction energies calculated using these fitted
1024: charges obtained via the Kollman-Singh procedure and the same energies
1025: calculated using solute charges which
1026: vary linearly with the parameter \( b \) defined in Equation (\ref{a_b_1andb_2Definitions})
1027: is \( 0.6 \) kcal/mol. This small deviation should be contrasted
1028: with the value of \( 4 \) kcal/mol representing the standard deviation
1029: of the values of the electrostatic interaction energy between the
1030: solute and solvent molecules (see Fig. 6).
1031:
1032: The distribution of the values of the polarization
1033: energy difference \( \Delta E_{\textrm{pol}} \) defined in Eq.~(\ref{DeltaEpol})
1034: relevant for the final re-weighting of data points is plotted in Fig. 7.
1035: As expected, the values
1036: of \( \Delta E_{\textrm{pol}} \) are distributed over a small range
1037: of energy values of approximately $0.4$ kcal/mol. In fact, calculation
1038: of the activation energy in the double proton reaction rate in tetrahydrofuran
1039: without re-weighting yields an activation energy which differs from
1040: the activation energy in Fig. 5
1041: only by a statistically irrelevant value of $0.1$ kcal/mol.
1042: It should be emphasized that the small values of \( \Delta E_{\textrm{pol}} \)
1043: in Fig. 7 are in part a consequence of approximating the polarization energy
1044: of the solute by the solvent by the polarization energy of the solvent by the solute.
1045: We estimated that if the solvent charges used in Eq. (11) were set to $q^{S}=q^{S}_p+q^{S}_{in}/2$,
1046: the standard deviation for \( \Delta E_{\textrm{pol}} \) would have been approximately $1.5$ kcal/mol.
1047:
1048: The potentials of mean force for the double proton transfer reaction in
1049: tetrahydrofuran calculated using the linearly-interpolated charge approach
1050: and using the simulated tempering method are identical
1051: within error bars. However, if the error bars for the activation energy
1052: in the linearly-interpolated charge method is $0.4$ kcal/mol, they are roughly four times
1053: larger in the simulation using the parallel tempering algorithm for
1054: comparable cpu times.
1055:
1056: The optimal number of electrostatic indices in the simulated tempering
1057: approach was found to be $i_{max}=4$. In the optimized setting
1058: for the simulated tempering method in which equal amounts of time
1059: were spent at all values of the electrostatic index,
1060: the Gibbs free energy differences $\Delta G_{i,i+1}$
1061: computed for consecutive values of the index were found to be approximately $0.7$
1062: kcal/mol. The Gibbs free energy $G(\lambda)$ of the electrostatic interaction between the
1063: solute and solvent charges have been estimated using Equation(\ref{Partition})
1064: and are plotted in Fig. 8 as a function of the
1065: fraction of solute charge $\lambda_i$ being turned on.
1066:
1067: Several comments are worth making about the results in Fig. 8.
1068: First, the partition functions $Z_{i}$ for the umbrella potential-biased
1069: ensemble from which the free energies were computed correspond to a
1070: particular value of the electrostatic fraction parameter
1071: $\lambda_i$. It should also be noted that although a difference in free energy between
1072: adjacent values of the electrostatic index of the order of $k_{B}T$ seems to suggest a low probability
1073: of rejecting swaps between consecutive values of the index, the actual rejection rate
1074: was found to be significantly higher.
1075: The relatively large rejection rate is due to the fact that differences in free energy reflect the
1076: \emph{average} energetic \emph{and} entropic differences
1077: between thermodynamic states with different $\lambda $,
1078: whereas the acceptance probability in Equation (\ref{AcceptanceCriterion})
1079: involves only the difference
1080: in the energies of the actual configurations to be swapped. In particular, it
1081: was observed that the enthalpic $H(\lambda)$ and entropic $-TS(\lambda)$ contributions
1082: to the Gibbs free energy $G(\lambda)$
1083: vary in opposite directions as the fraction of the solute charge $\lambda$
1084: increases from zero to one.
1085: These variations of enthalpic and entropic terms with the charge fraction
1086: can be visualized by comparing the results plotted in
1087: Fig. 8 and Fig. 9.
1088: From the two figures it appears that $H(\lambda)$ increases and $-TS(\lambda)$
1089: decreases with $\lambda$. The opposite directions in which enthalpic and entropic
1090: terms vary with $\lambda$ is reflected in the mobility of the state $(i,{\mathbf{r}})$ of the
1091: simulated tempering algorithm in the $i$ (or $\lambda$) subspace.
1092: Turning our attention to Equation (\ref{AcceptanceCriterion}), note that
1093: an increase of $-TS(\lambda)$ with $\lambda$ suggests that for a large fraction of
1094: configurations (i,{\textbf{r}}) transitions in which $i$ is \emph{increased} are
1095: accepted only if the ratio
1096: \begin{equation}
1097: \label{r}
1098: w_{\uparrow}(i)=\frac{w_{i+1}}{w_{i}}
1099: \end{equation}
1100: is approximately unity, $w_{\uparrow}(i) \approx 1$. Such is the case for an
1101: important number of transitions from electrostatic index $i=1$ to
1102: electrostatic index $i=2$, for example, for which an increase in $\lambda$ is accompanied
1103: only by a small decrease and occasionally even an increase in energy.
1104: On the other hand, for a $\lambda$ near unity, a decrease of $H(\lambda)$ with $\lambda$ suggests that
1105: a large fraction of transitions in which $i$ is \emph{decreased} are rejected
1106: unless $w_{\uparrow}(i) \ll 1$. Such is the case in a large number of transitions which are attempted from
1107: the final index $i_{max}$ to $i_{max}-1$ for example, which are accompanied by a significant
1108: increase in energy.
1109:
1110: The choice of the weights $w_{i}$ suggested in Equation (\ref{Choicew}) represents
1111: a good compromise in the sense that transitions which increase and which decrease
1112: $i$ have an equal probability of being accepted on average. Nevertheless, this
1113: analysis points to the fact that if one uses the simulated tempering approach for
1114: the study of chemical reactions in solution, one should
1115: try to avoid turning on the charges of the reactive core all the way
1116: from zero to their final values. As enthalpic and entropic terms will always vary in opposite directions
1117: during the charging process, the sampling could become quite inefficient, especially
1118: when studying reactions in which there is a difference in the net charge between
1119: reactants and transition states, and not just in their dipole moments as in the present
1120: case.
1121:
1122: However, this conclusion does not mean that the MMBIF approach used in
1123: the simulation with linearly interpolated charges
1124: will always be more efficient than the simulated tempering method. The efficiency
1125: of the MMBIF approach relies heavily on the appropriateness of the postulated variation of the
1126: charges in a reactive system with the reaction coordinate. In the present study,
1127: the approximation of the variation of the fitted Kollman-Singh charges
1128: with the solute configuration by a linear
1129: variation with the parameter \( b \) in Equation (\ref{a_b_1andb_2Definitions})
1130: proved quite accurate. However, such a simple relation could break down, especially
1131: when studying reactions with a net transfer of charge. In this case, we suggest
1132: using an approach which combines the benefits of both methods illustrated
1133: in the present study: the use of a simulated tempering approach in which
1134: the solute charge is gradually modified from a linear variation with the reaction coordinate
1135: to their actual values obtained via the Kollman-Singh approach.
1136:
1137:
1138:
1139: \section{Discussion and Conclusions}
1140:
1141: In this article, two important issues on the applicability of the
1142: molecular mechanics based importance sampling method to the study of
1143: reactive events in condensed phase environments have been addressed.
1144: One major concern in developing a successful implementation of the
1145: MMBIF approach is the ease of development of a sufficiently accurate
1146: molecular mechanics potential to guide the sampling. Although a fully
1147: automated approach to generating guiding potentials for general
1148: reactions is still not available, it is encouraging to note that the use of the same
1149: principles of bond-evolution theory\cite{Krokidis} as in our earlier study of the
1150: malonaldehyde system\cite{Iftimie2} were adequate for designing a molecular mechanics
1151: potential for the acetic acid-methanol system. This success is
1152: particularly impressive in light of the fact that the malonaldehyde and acetic
1153: acid-methanol systems differ substantially not only in their barrier
1154: heights (by a factor of $4$), but also in the qualitative nature of
1155: the chemical event (one proton versus two protons transferred).
1156: It suggests that bond-evolution theory guidelines are likely to be practical in developing molecular
1157: mechanics potentials for other proton transfer reactions, and for possibly
1158: other types of chemical events.
1159:
1160: The study of the double proton transfer reaction between acetic acid
1161: and methanol in tetrahydrofuran also demonstrates that the MMBIF
1162: method can be applied in reaction rate calculations
1163: of chemical transformations in solvents of medium polarity. In
1164: particular, an increase in the CPU time of factors of $4$ and $15$
1165: with respect to gas-phase calculations were obtained using two different
1166: sampling methods. Hence, we conclude that it should be possible in many instances
1167: to compute solvent mediated reaction rates with statistical accuracies
1168: comparable to those obtained in gas-phase calculations, even though the complexity
1169: of the calculation is increased enormously by the presence of the solvent.
1170:
1171: It should be emphasized that the contribution of the solvent to the
1172: total activation energy in the present study is only on the order of a
1173: few factors of $k_{B}T$ at room temperature. As a result, the actual
1174: mechanism of the proton transfer event is virtually unchanged from the
1175: process in gas phase. This simplification allowed the separation of
1176: the reactive \emph{ab initio} core and the solvent degrees of freedom
1177: into effectively disjoint sets which were updated in isolation in the
1178: parallel tempering implementation of the sampling.
1179: In many cases, the structure of the solvent plays a more substantial
1180: role in the reactive process. Under such circumstances, care must be
1181: taken to devise methods in which the reactive core and the solvent
1182: structure are updated in a more correlated fashion. For example, for
1183: a reaction in a solvent with a larger dielectric constant, the
1184: simulated tempering approach can be implemented by incorporating simultaneous
1185: MMBIF updates of the solute and solvent degrees of freedom at each
1186: electrostatic index.
1187:
1188: It is informative to compare and contrast the advantages and disadvantages
1189: of using the MMBIF method versus using present day molecular dynamics
1190: based methods for calculating reaction rates in solution using a hybrid
1191: QM/MM approach. The practical use of the simple molecular dynamics
1192: methods for sampling configurations of a system containing a reactive
1193: core which is described using \emph{ab initio} electronic structure
1194: methods and the solvent molecules which are described using a molecular
1195: mechanics potential is limited by the fact that calculating the time-evolution
1196: of the system necessitates the computation of the time-consuming
1197: \emph{ab initio} forces acting on the core atoms. These time-consuming
1198: calculations must be performed in traditional molecular dynamics calculations
1199: even when only the solvent degrees of freedom change significantly.
1200:
1201: Several methods have been proposed for circumventing the inefficiency
1202: of the traditional QM/MM molecular dynamics calculations which are
1203: based on an ``artificial'' separation of time scales
1204: associated with the solute and solvent atoms\cite{TuckermanJCP2002,RothlisbergerJPCB2001}.
1205: The central idea of these methods is to use a large mass in conjunction
1206: with a high temperature thermostat for the solute atoms, whereas the
1207: solvent atoms have usual masses and are in contact with a room temperature
1208: thermostat\cite{TuckermanJCP2002,RothlisbergerJPCB2001}. The large
1209: mass of the core atoms is chosen in such a way that the solvent degrees
1210: of freedom relax on a time scale which is much smaller than the time
1211: scale of the massive core atoms. Multiple-time scale arguments can
1212: be utilized to demonstrate that the integration of Hamilton's equations
1213: describing the evolution of the solute plus solvent system can be performed
1214: by computing the forces acting on the solvent degrees of freedom significantly
1215: more often than the \emph{ab initio} forces acting on the solute without
1216: altering the asymptotic Boltzmann distribution of the configurations
1217: of the system. In addition, the decoupling between the solute and
1218: solvent time scales enables the use of a large temperature thermostat
1219: coupled to the solute atoms without introducing an irreversible heat
1220: flow, thereby ensuring that the average kinetic energy of the core
1221: atoms is comparable with the magnitude of the reaction barrier which
1222: separates reactant and product configurations on the potential energy
1223: surface. Therefore, a relatively small number of \emph{ab initio}
1224: calculations must be performed before a reactive event occurs.
1225:
1226: Both the MMBIF and the modified molecular dynamics approach succinctly
1227: described above have a number of advantages and shortcomings when
1228: studying chemical reactions in solution which are essentially driven
1229: by fluctuations in the structure of the \emph{solute} using a QM/MM
1230: approach. The main disadvantage of the MMBIF method consists of the
1231: fact that a reasonable molecular mechanics description of the reactive
1232: event in the \emph{gas-phase} solute is required, and this molecular
1233: mechanics potential must usually be created from scratch. Nevertheless,
1234: once such molecular mechanics potential has been constructed, the
1235: cpu time necessary for calculating reaction rates in solution is essentially
1236: independent of the characteristics of the solvent such as its dielectric
1237: relaxation time. In contrast, the molecular dynamics approach does
1238: not necessitate prior information with respect to the potential energy
1239: surface of the cluster, although if such information exists, it can
1240: be used to improve the efficiency of the sampling\cite{TuckermanJCP2002,RothlisbergerJPCB2001}.
1241: However, the average time needed to observe a reaction event increases
1242: as the square root of the effective mass of the reactive degree of
1243: freedom. On the other hand, given the requirement
1244: of the separation of time scales between the solvent and solute degrees
1245: of freedom, the characteristic time scale of the solute atom motion
1246: and therefore the lower bound for the value of the mass needed for
1247: the core atoms is \emph{determined} by the duration of the solvent
1248: dielectric relaxation time. Therefore, it appears that the efficiency
1249: of the above mentioned molecular dynamics scheme decreases with the
1250: increase in the solvent dielectric relaxation time.
1251:
1252: The methodology proposed in the present work can be combined with the ideas
1253: presented in reference {[}\citeonline{Iftimie2}{]} to include nuclear quantum effects
1254: via centroid transition state theory with a supplementary increase in the cpu time
1255: by a factor of $2$ times the number of path-integral beads. Such a
1256: combination of procedures provides a rigorous and practical platform
1257: for calculation of kinetic isotope effects. An important goal for future
1258: work is to clarify the mechanistic origin of the relation between the breakdown of the
1259: rule of geometric mean in multiple proton transfer reactions and tunneling effects.
1260:
1261: \acknowledgements
1262: This work was supported by a grant from the Natural Sciences and Engineering
1263: Research Council of Canada. R. Iftimie would also like to thank the Ontario
1264: Ministry of Education for financial support.
1265:
1266:
1267:
1268: \begin{thebibliography}{10}
1269: \bibitem{FlemingPhysToday90}G. R. Fleming and P. G. Wolynes, Physics Today \textbf{43}, 36 (1990).
1270: \bibitem{ZewailExperimental}D. Zhong and A. H. Zewail, J. Phys. Chem. A, \textbf{102}, 4031 (1998).
1271: \bibitem{PermanScience98}B. Perman, V. Srajer, Z. Ren, T. Teng, C. Pradervand, T. Ursby, D.
1272: Bourgeois, F. Schotte, M. Wulff, R. Kort, K. Hellingwerf and K. Moffat,
1273: Science \textbf{279}, 1946 (1998).
1274: \bibitem{RosskyNature94}P. J. Rossky and J. D. Simon, Nature \textbf{370}, 263 (1994).
1275: \bibitem{ClaryTheoreticalGasPhase}D. C. Clary, Science \textbf{279}, 1879 (1998).
1276: \bibitem{KlinmanBiophys.}A. Kohen, R. Cannio, S. Bartolucci and J. P. Klinman, Nature \textbf{399},
1277: 496 (1999); B. J. Bahnson, T. D. Colby, J. K. Chin, B. M. Goldstein
1278: and J. P. Klinman, Proc. Natl. Acad. Sci. U.S.A. \textbf{94}, 12797
1279: (1997); Y. Cha, C. J. Murray and J. P. Klinman, Science \textbf{243},
1280: 1325 (1989).
1281: \bibitem{PageWilliams}M. Page and A. Williams, \emph{Organic and Bio-organic Mechanisms},
1282: (Addison Wesley Longman, Harlow, England, 1997).
1283: \bibitem{UmbrellaSampling}G. M. Torrie and J. P. Valleau, Chem. Phys. Lett. \textbf{28}, 578
1284: (1974).
1285: \bibitem{BlueMoonMethod}E. A. Carter, G. Ciccotti, J. T. Hynes and R. Kapral, Chem. Phys.
1286: Lett. \textbf{156}, 472 (1989).
1287: \bibitem{ProjectionMethods}S. Melchionna, Phys. Rev. E \textbf{62}, 8762 (2000).
1288: \bibitem{VariableTransform}Z. Zhu, M. E. Tuckerman and G. J. Martyna, Phys. Rev. Lett. \textbf{88},
1289: 100201 (2002).
1290: \bibitem{Iftimie1}R. Iftimie, D. Salahub, D. Wei and J. Schofield, J. Chem. Phys. \textbf{113},
1291: 4852 (2000).
1292: \bibitem{Rothlisberger2000}J. Van de Vondele and U. Rothlisberger, J. Chem. Phys. \textbf{113},
1293: 4863 (2000).
1294: \bibitem{Iftimie2}R. Iftimie and J. Schofield, J. Chem. Phys. \textbf{114}, 6763 (2001).
1295: \bibitem{Iftimie3}R. Iftimie and J. Schofield, J. Chem. Phys. \textbf{115}, 5891 (2001).
1296: \bibitem{Iftimie4}R. Iftimie and J. Schofield, Int. J. Quant. Chem. \textbf{91}, 404 (2003).
1297: \bibitem{relaxationWater}N. Nandi, S. Roy and B. Bagchi, J. Chem. Phys. \textbf{102}, 1390 (1995).
1298: \bibitem{relaxationTHF}A. Chaudhari, P. Khirade, R. Singh, S. N. Helambe, N. K. Narain and
1299: S. C. Mehrotra, J. Mol. Liq \textbf{82}, 245 (1999).
1300: \bibitem{KohenJACS2002}A. Kohen and J. H. Jensen, J. Am. Chem. Soc. \textbf{124}, 3858 (2002).
1301: \bibitem{LimbachBook}H. H. Limbach, in \emph{Electron and Proton Transfer in Chemistry
1302: and Biology}, (A. M\( \ddot{\textrm{u}} \)ler, H. Ratajczak, W. Junge
1303: and E. Diemann, eds., Elsevier, New York, 1992).
1304: \bibitem{KlinmanJACS2001}M. J. Knapp, K. Rickert and J. Klinman, J. Am. Chem. Soc. \textbf{124},
1305: 3865 (2002). For a simplified treatment of the problem of the rule
1306: of geometric mean see also reference {[}\citeonline{KohenJACS2002}{]}.
1307: \bibitem{Dugas}H. Dugas, \emph{Bioorganic Chemistry, a Chemical Approach to Enzyme
1308: Action}, (\( 3^{\textrm{rd}} \) edition, Springer-Verlag, New York,
1309: 1996). One of the textbook examples of bifunctional catalysis in solution
1310: chemistry is C. G. Swain and J. F. Brown, J. Am. Chem. Soc. \textbf{74},
1311: 2534 (1952).
1312: \bibitem{Limbach84}D. Gerritzen and H. H. Limbach, J. Am. Chem. Soc. \textbf{106}, 869
1313: (1984).
1314: \bibitem{Neumann}M. Neumann, Mol. Phys. \textbf{50}, 841 (1983).
1315: \bibitem{Watts}R. O. Watts, Mol. Phys. \textbf{28}, 1069 (1974).
1316: \bibitem{Barone}V. Barone and C. Adamo, J. Chem. Phys. \textbf{105}, 11007, (1996).
1317: \bibitem{Proynov}E. I. Proynov, A. Vela and D. R. Salahub, Chem. Phys. Lett. \textbf{230}, 419,
1318: (1994); \textbf{234}, 462(E), (1995).
1319: \bibitem{Sirois}S. Sirois, E. I. Proynov, D. T. Nguyen and D. R. Salahub, J. Chem. Phys. \textbf{107},
1320: 6770, (1997).
1321: \bibitem{Smedarchina}A. Fern\'{a}ndez-Ramos, Z. Smedarchina and
1322: J. Rodr\'{\i}guez-Otero, J. Chem. Phys \textbf{114}, 1567 (2001).
1323: \bibitem{Proynov2}E. I. Proynov, S. Sirois and D. R. Salahub, Int. J. Quantum Chem. \textbf{64},
1324: 427, (1997).
1325: \bibitem{deMon}
1326: For details on the deMon quantum chemistry package,
1327: see http://cvs.demon-software.com/public\_html/
1328: \bibitem{Jorgensen}W. L. Jorgensen, D. S. Maxwell and J. Tirado-Rives, J. Am. Chem. Soc.
1329: \textbf{118}, 11225 (1996).
1330: \bibitem{Thesis}R. Iftimie, Thesis, University of Toronto, (2001).
1331: \bibitem{RickBerne}S. W. Rick, S. J. Stuart and B. J. Berne, J. Chem. Phys. \textbf{101},
1332: 6141, (1994).
1333: \bibitem{Roitberg}M.C. Gonz\'{a}lez Lebrero, D.E. Bikiel, M. Dolores
1334: Elola, D.A. Estrin and A.E. Roitberg, J. Chem. Phys. \textbf{117},
1335: 2718 (2002).
1336: \bibitem{HohenbergKohn}P. Hohenberg and W. Kohn, Phys. Rev. \textbf{136}, B864 (1964).
1337: \bibitem{KohnSham}W. Kohn and L. J. Sham, Phys. Rev. \textbf{140}, A1133 (1965).
1338: \bibitem{Levy}M. Levy, Proc. Natl. Acad. Sci. USA \textbf{76}, 6062
1339: (1979).
1340: \bibitem{Rothlisberger} A. Laio, J. VandeVondele, and
1341: U. Rothlisberger, J. Chem. Phys. \textbf{116}, 6941 (2002).
1342: \bibitem{Reichardt}C. Reichardt, \emph{Solvents and Solvent Effects in Organic Chemistry},
1343: (VCH, New York, 1988, Appendix A, Table A-1).
1344: \bibitem{Krokidis}X. Krokidis, V. Goncalves, A. Savin, B. Silvi, J. Phys. Chem. A, \textbf{102},
1345: 5065 (1998); see in particular the discussion in page 5070 and Figure
1346: 8.; X. Krokidis, S. Nouri, B. Silvi, J. Phys. Chem. A, \textbf{101},
1347: 7277 (1997).
1348: \bibitem{MarinariParisi}E. Marinari and G. Parisi, Europhys. Lett. \textbf{19}, 451 (1992).
1349: \bibitem{Geyer1995}C. J. Geyer and E. A. Thompson, J. Am. Stat. Assoc. \textbf{90}, 909
1350: (1995).
1351: \bibitem{KollmanSingh}U. C. Singh and P. A. Kollman, J. Comp. Chem. \textbf{5}, 129 (1984).
1352: \bibitem{TuckermanJCP2002}L. Russo, P. Min\'{a}ry, Z. Zhu and M. E. Tuckerman, J. Chem. Phys.
1353: \textbf{116}, 4389 (2002).
1354: \bibitem{RothlisbergerJPCB2001}J. V. Vondele and U. Rothlisberger, J. Phys. Chem. B \textbf{106},
1355: 203 (2001).
1356: \bibitem{Atkins}P. W. Atkins, \emph{Molecular Quantum Mechanics}, (Chapter 8, Oxford
1357: University Press, New York, 1987).
1358: \bibitem{Polarization2002}G. Tabacchi, C. J. Mundy, J. H\( \ddot{\textrm{u}} \)tter and M.
1359: Parrinello, J. Chem. Phys. \textbf{117}, 1416 (2002).
1360: \bibitem{Bottcher}C. J. F. B\( \ddot{\textrm{o}} \)ttcher, Theory of electric polarization,
1361: vol. I, ($2^{\text{nd}}$ edition,
1362: Elsevier, New York, 1973), especially Chapter 4.
1363: \end{thebibliography}
1364:
1365: \newpage
1366: \begin{center}
1367: Figure Captions
1368: \end{center}
1369: Figure 1: The formation of two cyclic clusters involving one or
1370: two acetic acid molecules and one molecule of methanol. The double-proton
1371: transfer reaction studied here
1372: takes place along the hydrogen bonds of the cyclic cluster formed from
1373: one molecule of acetic acid and one molecule of methanol.
1374:
1375: \vspace{.1in}
1376: \noindent
1377: Figure 2: A pictorial view of the QM/MM/continuum solvent
1378: method and of the reaction field approach. The ``reactive
1379: core'' region contains a quantum representation (i.e: nuclei
1380: + electrons) of the atoms which are involved in the actual covalent
1381: bond-breaking and bond-forming events. The ``explicit solvent''
1382: region contains an atomic representation of the first few shells of
1383: solvent molecules. The effects of the solvent molecules which are
1384: far from the reaction center are included in an implicit manner in
1385: the QM/MM/continuum approach using the reaction field method. In our
1386: implementation, the reaction field method consists of calculating
1387: an effective electrostatic interaction between the dipole moments
1388: of the molecules inside the first two regions (see Equation {[}\ref{ReactionFieldEnergy}{]}).
1389:
1390: \vspace{.1in}
1391: \noindent
1392: Figure 3: The structures of the reactant, transition
1393: state and product in the gas-phase double-proton transfer reaction.
1394:
1395: \vspace{.1in}
1396: \noindent
1397: Figure 4: A schematic representation of
1398: the simulated tempering method which uses the MMBIF approach to generate
1399: configurations of the system in a simulation where solute polarizability
1400: is neglected. The values of the electrostatic scaling parameter $\lambda_i$ for electrostatic indices $i=1 \cdots
1401: i_{max}$ are gradually increased from zero to one.
1402:
1403: \vspace{.1in}
1404: \noindent
1405: Figure 5: The calculated potentials of
1406: mean force for the double-proton transfer reaction in the acetic acid-methanol
1407: cluster in gas-phase and in a solution of tetrahydrofuran using the
1408: reaction coordinate \protect\( \xi \protect \) defined in Equation
1409: (\ref{ReactionCoordinateMalonaldehyde2}). Note that the difference
1410: between the activation energies is approximately \protect\( 0.8\protect \)
1411: kcal/mol. This difference is larger than the width of the \protect\( 75\%\protect \)
1412: confidence intervals for the activation energies, which have been
1413: estimated to be \protect\( 0.2\protect \) kcal/mol for the gas-phase
1414: and \protect\( 0.4\protect \) kcal/mol for the liquid-phase
1415: simulations (inset).
1416:
1417: \vspace{.1in}
1418: \noindent
1419: Figure 6: The values of the
1420: electrostatic interaction energy between the solute and solvent molecules
1421: obtained for $\lambda=1$ in the simulated tempering method in which
1422: the solute charges are the gas-phase Kollman-Singh charges. Note that
1423: the values of the solute-solvent electrostatic interaction energies
1424: are scattered over an energy range of approximately \protect\( 4\protect \)
1425: kcal/mol. In contrast, the differences in the electrostatic interaction
1426: energies calculated using the Kollman-Singh charges and the same energies
1427: calculated using solute charges which vary linearly with the parameter
1428: \protect\( b\protect \) defined in Equation (\ref{a_b_1andb_2Definitions})
1429: are scattered over an interval of only \protect\( 0.6\protect \)
1430: kcal/mol (data not shown).
1431:
1432: \vspace{.1in}
1433: \noindent
1434: Figure 7: The values of the
1435: polarization energy difference \protect\( \Delta E_{\textrm{pol}}\protect \)
1436: defined in Equation (\ref{DeltaEpol}). Note
1437: that the distribution of values is centered roughly at \protect\( 0\protect \)
1438: kcal/mol and has a small standard deviation of \protect\( 0.4\protect \)
1439: kcal/mol.
1440:
1441: \vspace{.1in}
1442: \noindent
1443: Figure 8: The free electrostatic interaction energy
1444: between an acetic acid-methanol complex and molecules
1445: of tetrahydrofuran solvent as a function of the charge
1446: state of the complex. $\lambda=0$ and $\lambda=1$ correspond to
1447: fully uncharged and charged solute. The stars and diamonds represent
1448: the estimated values of the free energy obtained from the preliminary
1449: and from the optimized simulated tempering runs, respectively. The continous
1450: line represents a spline interpolation between the computed values.
1451:
1452: \vspace{.1in}
1453: \noindent
1454: Figure 9: The normalized probability density
1455: of the solvent-solute electrostatic interaction energy on for
1456: electrostatic indices $i=2$ (solid curve),
1457: $i=3$ (long dashed curve) and $i=4$ (dashed curve). Note that the enthalpies,
1458: calculated as the average solute-solvent electrostatic interaction energies,
1459: are approximately $-1.0$, $-2.4$ and
1460: $-3.5$ kcal/mol. Comparing these results with the corresponding free energies in
1461: Fig. 8 one obtains the corresponding entropies as
1462: being $0.3$, $1.0$ and $1.4$ kcal/mol. The variations in the mechanical work
1463: $pV$ with the number of chain are negligible.
1464:
1465:
1466: \newpage
1467: \begin{table}
1468:
1469: \caption{\label{OPLSAAExperimentalModOPLSAA}The values of the gas-phase dipole
1470: moment \protect\( \mu _{g}\protect \) expressed in Debyes, and of
1471: the static dielectric constant \protect\( \epsilon _{r}\protect \),
1472: obtained by considering the OPLS-AA charges as permanent charges,
1473: obtained from our modified version of the OPLS-AA force field (see
1474: text) which accounts approximately for electronic polarization effects,
1475: and from experimental data from reference {[}\citeonline{Reichardt}{]}.}
1476:
1477: \vspace{0.3cm}
1478:
1479: \begin{centering}
1480:
1481: \setlength{\extrarowheight}{6pt}
1482:
1483: \begin{tabular}{|c|c|c|c|}
1484: \hline
1485: &
1486: OPLS-AA&
1487: Modified OPLS-AA&
1488: Experimental\\
1489: \hline
1490: \hline
1491: \( \mu _{g}(D) \)&
1492: \( 1.97 \)&
1493: \( 1.76 \)&
1494: \( 1.75 \)\\
1495: \hline
1496: \( \epsilon _{r} \)&
1497: \( 6.15\pm 0.3 \)&
1498: \( 7.61\pm 0.38 \)&
1499: \( 7.58 \)\\
1500: \hline
1501: \end{tabular}
1502:
1503: \end{centering}
1504: \end{table}
1505:
1506:
1507:
1508: \end{document}
1509:
1510: