ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipoleSFPaper/multipoleSFPaper.tex
Revision: 3073
Committed: Fri Nov 3 20:05:01 2006 UTC (17 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 56229 byte(s)
Log Message:
final version after Steve's edits

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 \documentclass[11pt]{article}
3 \usepackage{tabls}
4 %\usepackage{endfloat}
5 \usepackage[tbtags]{amsmath}
6 \usepackage{amsmath,bm}
7 \usepackage{amssymb}
8 \usepackage{mathrsfs}
9 \usepackage{setspace}
10 \usepackage{tabularx}
11 \usepackage{graphicx}
12 \usepackage{booktabs}
13 \usepackage{colortbl}
14 \usepackage[ref]{overcite}
15 \pagestyle{plain}
16 \pagenumbering{arabic}
17 \oddsidemargin 0.0cm \evensidemargin 0.0cm
18 \topmargin -21pt \headsep 10pt
19 \textheight 9.0in \textwidth 6.5in
20 \brokenpenalty=10000
21 \renewcommand{\baselinestretch}{1.2}
22 \renewcommand\citemid{\ } % no comma in optional reference note
23 %\AtBeginDelayedFloats{\renewcommand{\baselinestretch}{2}} %doublespace captions
24 %\let\Caption\caption
25 %\renewcommand\caption[1]{%
26 % \Caption[#1]{}%
27 %}
28 %\setlength{\abovecaptionskip}{1.2in}
29 %\setlength{\belowcaptionskip}{1.2in}
30
31 \begin{document}
32
33 \title{Pairwise Alternatives to the Ewald Sum: Applications
34 and Extension to Point Multipoles}
35
36 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
37 Department of Chemistry and Biochemistry\\
38 University of Notre Dame\\
39 Notre Dame, Indiana 46556}
40
41 \date{\today}
42
43 \maketitle
44 %\doublespacing
45
46 \begin{abstract}
47 The damped, shifted-force electrostatic potential has been shown to
48 give nearly quantitative agreement with smooth particle mesh Ewald for
49 energy differences between configurations as well as for atomic force
50 and molecular torque vectors. In this paper, we extend this technique
51 to handle interactions between electrostatic multipoles. We also
52 investigate the effects of damped and shifted electrostatics on the
53 static, thermodynamic, and dynamic properties of liquid water and
54 various polymorphs of ice. Additionally, we provide a way of choosing
55 the optimal damping strength for a given cutoff radius that reproduces
56 the static dielectric constant in a variety of water models.
57 \end{abstract}
58
59 \newpage
60
61 %\narrowtext
62
63 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64 % BODY OF TEXT
65 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66
67 \section{Introduction}
68
69 Over the past several years, there has been increasing interest in
70 pairwise methods for correcting electrostatic interactions in computer
71 simulations of condensed molecular
72 systems.\cite{Wolf99,Zahn02,Kast03,Beck05,Ma05,Fennell06} These
73 techniques were developed from the observations and efforts of Wolf
74 {\it et al.} and their work towards an $\mathcal{O}(N)$ Coulombic
75 sum.\cite{Wolf99} Wolf's method of cutoff neutralization is able to
76 obtain excellent agreement with Madelung energies in ionic
77 crystals.\cite{Wolf99}
78
79 In a recent paper, we showed that simple modifications to Wolf's
80 method could give nearly quantitative agreement with smooth particle
81 mesh Ewald (SPME) for quantities of interest in Monte Carlo
82 (i.e. configurational energy differences) and Molecular Dynamics
83 (i.e. atomic force and molecular torque vectors).\cite{Fennell06} We
84 described the undamped and damped shifted potential (SP) and shifted
85 force (SF) techniques. The potential for the damped form of the SP
86 method, where $\alpha$ is the adjustable damping parameter, is given
87 by
88 \begin{equation}
89 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
90 - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
91 \quad r_{ij}\leqslant R_\textrm{c},
92 \label{eq:DSPPot}
93 \end{equation}
94 while the damped form of the SF method is given by
95 \begin{equation}
96 \begin{split}
97 V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
98 \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
99 - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
100 &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
101 + \frac{2\alpha}{\pi^{1/2}}
102 \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
103 \right)\left(r_{ij}-R_\mathrm{c}\right)\ \Biggr{]}
104 \quad r_{ij}\leqslant R_\textrm{c}.
105 \label{eq:DSFPot}
106 \end{split}
107 \end{equation}
108 In these potentials and in all electrostatic quantities that follow,
109 the standard $4 \pi \epsilon_{0}$ has been omitted for clarity.
110
111 The damped SF method is an improvement over the SP method because
112 there is no discontinuity in the forces as particles move out of the
113 cutoff radius ($R_\textrm{c}$). This is accomplished by shifting the
114 forces (and potential) to zero at $R_\textrm{c}$. This is analogous to
115 neutralizing the charge as well as the force effect of the charges
116 within $R_\textrm{c}$.
117
118 To complete the charge neutralization procedure, a self-neutralization
119 term is added to the potential. This term is constant (as long as the
120 charges and cutoff radius do not change), and exists outside the
121 normal pair-loop. It can be thought of as a contribution from a
122 charge opposite in sign, but equal in magnitude, to the central
123 charge, which has been spread out over the surface of the cutoff
124 sphere. This term is calculated via a single loop over all charges in
125 the system. For the undamped case, the self term can be written as
126 \begin{equation}
127 V_\textrm{self} = - \frac{1}{2 R_\textrm{c}} \sum_{i=1}^N q_i^{2},
128 \label{eq:selfTerm}
129 \end{equation}
130 while for the damped case it can be written as
131 \begin{equation}
132 V_\textrm{self} = - \left(\frac{\alpha}{\sqrt{\pi}}
133 + \frac{\textrm{erfc}(\alpha
134 R_\textrm{c})}{2R_\textrm{c}}\right) \sum_{i=1}^N q_i^{2}.
135 \label{eq:dampSelfTerm}
136 \end{equation}
137 The first term within the parentheses of equation
138 (\ref{eq:dampSelfTerm}) is identical to the self term in the Ewald
139 summation, and comes from the utilization of the complimentary error
140 function for electrostatic damping.\cite{deLeeuw80,Wolf99}
141
142 The SF, SP, and Wolf methods operate by neutralizing the total charge
143 contained within the cutoff sphere surrounding each particle. This is
144 accomplished by shifting the potential functions to generate image
145 charges on the surface of the cutoff sphere for each pair interaction
146 computed within $R_\textrm{c}$. The damping function applied to the
147 potential is also an important method for accelerating convergence.
148 In the case of systems involving electrostatic distributions of higher
149 order than point charges, the question remains: How will the shifting
150 and damping need to be modified in order to accommodate point
151 multipoles?
152
153 \section{Electrostatic Damping for Point
154 Multipoles}\label{sec:dampingMultipoles}
155
156 To apply the SF method for systems involving point multipoles, we
157 consider separately the two techniques (shifting and damping) which
158 contribute to the effectiveness of the DSF potential.
159
160 As noted above, shifting the potential and forces is employed to
161 neutralize the total charge contained within each cutoff sphere;
162 however, in a system composed purely of point multipoles, each cutoff
163 sphere is already neutral, so shifting becomes unnecessary.
164
165 In a mixed system of charges and multipoles, the undamped SF potential
166 needs only to shift the force terms between charges and smoothly
167 truncate the multipolar interactions with a switching function. The
168 switching function is required for energy conservation, because a
169 discontinuity will exist in both the potential and forces at
170 $R_\textrm{c}$ in the absence of shifting terms.
171
172 To dampen the SF potential for point multipoles, we need to incorporate
173 the complimentary error function term into the standard forms of the
174 multipolar potentials. We can determine the necessary damping
175 functions by replacing $1/r_{ij}$ with $\mathrm{erfc}(\alpha r_{ij})/r_{ij}$ in the
176 multipole expansion. This procedure quickly becomes quite complex
177 with ``two-center'' systems, like the dipole-dipole potential, and is
178 typically approached using spherical harmonics.\cite{Hirschfelder67} A
179 simpler method for determining damped multipolar interaction
180 potentials arises when we adopt the tensor formalism described by
181 Stone.\cite{Stone02}
182
183 The tensor formalism for electrostatic interactions involves obtaining
184 the multipole interactions from successive gradients of the monopole
185 potential. Thus, tensors of rank zero through two are
186 \begin{equation}
187 T = \frac{1}{r_{ij}},
188 \label{eq:tensorRank1}
189 \end{equation}
190 \begin{equation}
191 T_\alpha = \nabla_\alpha \frac{1}{r_{ij}},
192 \label{eq:tensorRank2}
193 \end{equation}
194 \begin{equation}
195 T_{\alpha\beta} = \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
196 \label{eq:tensorRank3}
197 \end{equation}
198 where the form of the first tensor is the charge-charge potential, the
199 second gives the charge-dipole potential, and the third gives the
200 charge-quadrupole and dipole-dipole potentials.\cite{Stone02} Since
201 the force is $-\nabla V$, the forces for each potential come from the
202 next higher tensor.
203
204 As one would do with the multipolar expansion, we can replace $r_{ij}^{-1}$
205 with $\mathrm{erfc}(\alpha r_{ij})/r_{ij}$ to obtain damped forms of the
206 electrostatic potentials. Equation \ref{eq:tensorRank2} generates a
207 damped charge-dipole potential,
208 \begin{equation}
209 V_\textrm{Dcd} = -q_i\frac{\mathbf{r}_{ij}\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
210 c_1(r_{ij}),
211 \label{eq:dChargeDipole}
212 \end{equation}
213 where $c_1(r_{ij})$ is
214 \begin{equation}
215 c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
216 + \textrm{erfc}(\alpha r_{ij}).
217 \label{eq:c1Func}
218 \end{equation}
219 Equation \ref{eq:tensorRank3} generates a damped dipole-dipole potential,
220 \begin{equation}
221 V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
222 (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
223 c_2(r_{ij}) -
224 \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
225 c_1(r_{ij}),
226 \label{eq:dampDipoleDipole}
227 \end{equation}
228 where
229 \begin{equation}
230 c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
231 + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
232 + \textrm{erfc}(\alpha r_{ij}).
233 \end{equation}
234 Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
235 term. Continuing with higher rank tensors, we can obtain the damping
236 functions for higher multipole potentials and forces. Each subsequent
237 damping function includes one additional term, and we can simplify the
238 procedure for obtaining these terms by writing out the following
239 recurrence relation,
240 \begin{equation}
241 c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
242 {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
243 \label{eq:dampingGeneratingFunc}
244 \end{equation}
245 where,
246 \begin{equation}
247 m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
248 m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
249 m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
250 1 & m = -1\textrm{ or }0,
251 \end{array}\right.
252 \label{eq:doubleFactorial}
253 \end{equation}
254 and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
255 is similar in form to those obtained by Smith and Aguado and Madden
256 for the application of the Ewald sum to
257 multipoles.\cite{Smith82,Smith98,Aguado03}
258
259 Returning to the dipole-dipole example, the potential consists of a
260 portion dependent upon $r_{ij}^{-5}$ and another dependent upon
261 $r_{ij}^{-3}$. $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
262 respectively. The forces for the damped dipole-dipole interaction, are
263 obtained from the next higher tensor, $T_{\alpha \beta \gamma}$,
264 \begin{equation}
265 \begin{split}
266 F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
267 (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
268 c_3(r_{ij})\\ &-
269 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})\cdot\boldsymbol{\mu}_j +
270 (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\cdot\boldsymbol{\mu}_i +
271 \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
272 {r^5_{ij}} c_2(r_{ij}),
273 \end{split}
274 \label{eq:dampDipoleDipoleForces}
275 \end{equation}
276 Using the tensor formalism, we can dampen higher order multipolar
277 interactions using the same effective damping function that we use for
278 charge-charge interactions. This allows us to include multipoles in
279 simulations involving damped electrostatic interactions. In general,
280 if the multipolar potentials are left in $\mathbf{r}_{ij}/r_{ij}$
281 form, instead of reducing them to the more common angular forms which
282 use $\hat{\mathbf{r}}_{ij}$ (or the resultant angles), one may simply replace
283 any $1/r_{ij}^{2n+1}$ dependence with $c_n(r_{ij}) / r_{ij}^{2n+1}$ to
284 obtain the damped version of that multipolar potential.
285
286 As a practical consideration, we note that the evaluation of the
287 complementary error function inside a pair loop can become quite
288 costly. In practice, we pre-compute the $c_n(r)$ functions over a
289 grid of $r$ values and use cubic spline interpolation to obtain
290 estimates of these functions when necessary. Using this procedure,
291 the computational cost of damped electrostatics is only marginally
292 more costly than the undamped case.
293
294 \section{Applications of Damped Shifted-Force Electrostatics}
295
296 Our earlier work on the SF method showed that it can give nearly
297 quantitive agreement with SPME-derived configurational energy
298 differences. The force and torque vectors in identical configurations
299 are also nearly equivalent under the damped SF potential and
300 SPME.\cite{Fennell06} Although these measures bode well for the
301 performance of the SF method in both Monte Carlo and Molecular
302 Dynamics simulations, it would be helpful to have direct comparisons
303 of structural, thermodynamic, and dynamic quantities. To address
304 this, we performed a detailed analysis of a group of simulations
305 involving water models (both point charge and multipolar) under a
306 number of different simulation conditions.
307
308 To provide the most difficult test for the damped SF method, we have
309 chosen a model that has been optimized for use with Ewald sum, and
310 have compared the simulated properties to those computed via Ewald.
311 It is well known that water models parametrized for use with the Ewald
312 sum give calculated properties that are in better agreement with
313 experiment.\cite{vanderSpoel98,Horn04,Rick04} For these reasons, we
314 chose the TIP5P-E water model for our comparisons involving point
315 charges.\cite{Rick04}
316
317 The TIP5P-E water model is a variant of Mahoney and Jorgensen's
318 five-point transferable intermolecular potential (TIP5P) model for
319 water.\cite{Mahoney00} TIP5P was developed to reproduce the density
320 maximum in liquid water near 4$^\circ$C. As with many previous point
321 charge water models (such as ST2, TIP3P, TIP4P, SPC, and SPC/E), TIP5P
322 was parametrized using a simple cutoff with no long-range
323 electrostatic
324 correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
325 Without this correction, the pressure term on the central particle
326 from the surroundings is missing. When this correction is included,
327 the system expands to compensate for this added pressure and therefore
328 under-predicts the density of water under standard conditions. In
329 developing TIP5P-E, Rick preserved the geometry and point charge
330 magnitudes in TIP5P and focused on altering the Lennard-Jones
331 parameters to correct the density at 298~K. With the density
332 corrected, he compared common water properties for TIP5P-E using the
333 Ewald sum with TIP5P using a 9~\AA\ cutoff.
334
335 In the following sections, we compare these same properties calculated
336 from TIP5P-E using the Ewald sum with TIP5P-E using the damped SF
337 technique. Our comparisons include the SF technique with a 12~\AA\
338 cutoff and an $\alpha$ = 0.0, 0.1, and 0.2~\AA$^{-1}$, as well as a
339 9~\AA\ cutoff with an $\alpha$ = 0.2~\AA$^{-1}$.
340
341 Moving beyond point-charge electrostatics, the soft sticky dipole
342 (SSD) family of water models is the perfect test case for the
343 point-multipolar extension of damped electrostatics. SSD water models
344 are single point molecules that consist of a ``soft'' Lennard-Jones
345 sphere, a point-dipole, and a tetrahedral function for capturing the
346 hydrogen bonding nature of water - a spherical harmonic term for
347 water-water tetrahedral interactions and a point-quadrupole for
348 interactions with surrounding charges. Detailed descriptions of these
349 models can be found in other
350 studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
351
352 In deciding which version of the SSD model to use, we need only
353 consider that the SF technique was presented as a pairwise replacement
354 for the Ewald summation. It has been suggested that models
355 parametrized for the Ewald summation (like TIP5P-E) would be
356 appropriate for use with a reaction field and vice versa.\cite{Rick04}
357 Therefore, we decided to test the SSD/RF water model, which was
358 parametrized for use with a reaction field, with the damped
359 electrostatic technique to see how the calculated properties change.
360
361 \subsection{The Density Maximum of TIP5P-E}\label{sec:t5peDensity}
362
363 To compare densities, $NPT$ simulations were performed with the same
364 temperatures as those selected by Rick in his Ewald summation
365 simulations.\cite{Rick04} In order to improve statistics around the
366 density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
367 25$^\circ$C, while 2~ns trajectories were obtained at all other
368 temperatures. The average densities were calculated from the latter
369 three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
370 method for accumulating statistics, these sequences were spliced into
371 200 segments, each providing an average density. These 200 density
372 values were used to calculate the average and standard deviation of
373 the density at each temperature.\cite{Mahoney00}
374
375 \begin{figure}
376 \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
377 \caption{Density versus temperature for the TIP5P-E water model when
378 using the Ewald summation (Ref. \citen{Rick04}) and the SF method with
379 varying cutoff radii and damping coefficients. The pressure term from
380 the image-charge shell is larger than that provided by the
381 reciprocal-space portion of the Ewald summation, leading to slightly
382 lower densities. This effect is more visible with the 9~\AA\ cutoff,
383 where the image charges exert a greater force on the central
384 particle. The representative error bar for the SF methods shows the
385 average one-sigma uncertainty of the density measurement, and this
386 uncertainty is the same for all the SF curves.}
387 \label{fig:t5peDensities}
388 \end{figure}
389 Figure \ref{fig:t5peDensities} shows the densities calculated for
390 TIP5P-E using differing electrostatic corrections overlaid with the
391 experimental values.\cite{CRC80} The densities when using the SF
392 technique are close to, but typically lower than, those calculated
393 using the Ewald summation. These slightly reduced densities indicate
394 that the pressure component from the image charges at R$_\textrm{c}$
395 is larger than that exerted by the reciprocal-space portion of the
396 Ewald summation. Bringing the image charges closer to the central
397 particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
398 preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
399 charge interactions on the central particle and results in a further
400 reduction of the densities.
401
402 Because the strength of the image charge interactions has a noticeable
403 effect on the density, we would expect the use of electrostatic
404 damping to also play a role in these calculations. Larger values of
405 $\alpha$ weaken the pair-interactions; and since electrostatic damping
406 is distance-dependent, force components from the image charges will be
407 reduced more than those from particles close the the central
408 charge. This effect is visible in figure \ref{fig:t5peDensities} with
409 the damped SF sums showing slightly higher densities than the undamped
410 case; however, it is clear that the choice of cutoff radius plays a
411 much more important role in the resulting densities.
412
413 All of the above density calculations were performed with systems of
414 512 water molecules. Rick observed a system size dependence of the
415 computed densities when using the Ewald summation, most likely due to
416 his tying of the convergence parameter to the box
417 dimensions.\cite{Rick04} For systems of 256 water molecules, the
418 calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
419 system size of 256 molecules would force the use of a shorter
420 R$_\textrm{c}$ when using the SF technique, and this would also lower
421 the densities. Moving to larger systems, as long as the R$_\textrm{c}$
422 remains at a fixed value, we would expect the densities to remain
423 constant.
424
425 \subsection{Liquid Structure of TIP5P-E}\label{sec:t5peLiqStructure}
426
427 The experimentally-determined oxygen-oxygen pair correlation function
428 ($g_\textrm{OO}(r)$) for liquid water has been compared in great
429 detail with predictions of the various common water models, and TIP5P
430 was found to be in better agreement than other rigid, non-polarizable
431 models.\cite{Sorenson00} This excellent agreement with experiment was
432 maintained when Rick developed TIP5P-E.\cite{Rick04} To check whether
433 the choice of using the Ewald summation or the SF technique alters the
434 liquid structure, we calculated this correlation function at 298~K and
435 1~atm for the parameters used in the previous section.
436
437 \begin{figure}
438 \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
439 \caption{The oxygen-oxygen pair correlation functions calculated for
440 TIP5P-E at 298~K and 1~atm while using the Ewald summation
441 (Ref. \citen{Rick04}) and the SF technique with varying
442 parameters. Even with the lower densities obtained using the SF
443 technique, the correlation functions are essentially identical.}
444 \label{fig:t5peGofRs}
445 \end{figure}
446 The pair correlation functions calculated for TIP5P-E while using the
447 SF technique with various parameters are overlaid on the same function
448 obtained while using the Ewald summation in
449 figure~\ref{fig:t5peGofRs}. The differences in density do not appear
450 to have any effect on the liquid structure as the correlation
451 functions are indistinguishable. These results do indicate that
452 $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
453 correction.
454
455 \subsection{Thermodynamic Properties of TIP5P-E}\label{sec:t5peThermo}
456
457 In addition to the density and structual features of the liquid, there
458 are a variety of thermodynamic quantities that can be calculated for
459 water and compared directly to experimental values. Some of these
460 additional quantities include the latent heat of vaporization ($\Delta
461 H_\textrm{vap}$), the constant pressure heat capacity ($C_p$), the
462 isothermal compressibility ($\kappa_T$), the thermal expansivity
463 ($\alpha_p$), and the static dielectric constant ($\epsilon$). All of
464 these properties were calculated for TIP5P-E with the Ewald summation,
465 so they provide a good set of reference data for comparisons involving
466 the SF technique.
467
468 The $\Delta H_\textrm{vap}$ is the enthalpy change required to
469 transform one mole of substance from the liquid phase to the gas
470 phase.\cite{Berry00} In molecular simulations, this quantity can be
471 determined via
472 \begin{equation}
473 \begin{split}
474 \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq} \\
475 &= E_\textrm{gas} - E_\textrm{liq}
476 + P(V_\textrm{gas} - V_\textrm{liq}) \\
477 &\approx -\frac{\langle U_\textrm{liq}\rangle}{N}+ RT,
478 \end{split}
479 \label{eq:DeltaHVap}
480 \end{equation}
481 where $E$ is the total energy, $U$ is the potential energy, $P$ is the
482 pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
483 the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
484 seen in the last line of equation (\ref{eq:DeltaHVap}), we can
485 approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
486 state. This allows us to cancel the kinetic energy terms, leaving only
487 the $U_\textrm{liq}$ term. Additionally, the $PV$ term for the gas is
488 several orders of magnitude larger than that of the liquid, so we can
489 neglect the liquid $PV$ term.
490
491 The remaining thermodynamic properties can all be calculated from
492 fluctuations of the enthalpy, volume, and system dipole
493 moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
494 enthalpy in constant pressure simulations via
495 \begin{equation}
496 \begin{split}
497 C_p = \left(\frac{\partial H}{\partial T}\right)_{N,P}
498 = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
499 \end{split}
500 \label{eq:Cp}
501 \end{equation}
502 where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
503 \begin{equation}
504 \begin{split}
505 \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
506 = \frac{({\langle V^2\rangle}_{NPT} - {\langle V\rangle}^{2}_{NPT})}
507 {k_BT\langle V\rangle_{NPT}},
508 \end{split}
509 \label{eq:kappa}
510 \end{equation}
511 and $\alpha_p$ can be calculated via
512 \begin{equation}
513 \begin{split}
514 \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
515 = \frac{(\langle VH\rangle_{NPT}
516 - \langle V\rangle_{NPT}\langle H\rangle_{NPT})}
517 {k_BT^2\langle V\rangle_{NPT}}.
518 \end{split}
519 \label{eq:alpha}
520 \end{equation}
521 Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
522 be calculated for systems of non-polarizable substances via
523 \begin{equation}
524 \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
525 \label{eq:staticDielectric}
526 \end{equation}
527 where $\epsilon_0$ is the permittivity of free space and $\langle
528 M^2\rangle$ is the fluctuation of the system dipole
529 moment.\cite{Allen87} The numerator in the fractional term in equation
530 (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
531 dipole moment, identical to the quantity calculated in the
532 finite-system Kirkwood $g$ factor ($G_k$):
533 \begin{equation}
534 G_k = \frac{\langle M^2\rangle}{N\mu^2},
535 \label{eq:KirkwoodFactor}
536 \end{equation}
537 where $\mu$ is the dipole moment of a single molecule of the
538 homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
539 fluctuation term in both equation (\ref{eq:staticDielectric}) and
540 (\ref{eq:KirkwoodFactor}) is calculated as follows,
541 \begin{equation}
542 \begin{split}
543 \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
544 - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
545 &= \langle M_x^2+M_y^2+M_z^2\rangle
546 - (\langle M_x\rangle^2 + \langle M_x\rangle^2
547 + \langle M_x\rangle^2).
548 \end{split}
549 \label{eq:fluctBoxDipole}
550 \end{equation}
551 This fluctuation term can be accumulated during the simulation;
552 however, it converges rather slowly, thus requiring multi-nanosecond
553 simulation times.\cite{Horn04} In the case of tin-foil boundary
554 conditions, the dielectric/surface term of the Ewald summation is
555 equal to zero. Since the SF method also lacks this
556 dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
557 valid for determining static dielectric constants.
558
559 All of the above properties were calculated from the same trajectories
560 used to determine the densities in section \ref{sec:t5peDensity}
561 except for the static dielectric constants. The $\epsilon$ values were
562 accumulated from 2~ns $NVE$ ensemble trajectories with system densities
563 fixed at the average values from the $NPT$ simulations at each of the
564 temperatures. The resulting values are displayed in figure
565 \ref{fig:t5peThermo}.
566 \begin{figure}
567 \centering
568 \includegraphics[width=5.8in]{./figures/t5peThermo.pdf}
569 \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
570 and the SF techniques along with the experimental values. Units
571 for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
572 cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
573 and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
574 reference \citen{Rick04}. Experimental values for $\Delta
575 H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
576 \citen{Kell75}. Experimental values for $C_p$ are from reference
577 \citen{Wagner02}. Experimental values for $\epsilon$ are from reference
578 \citen{Malmberg56}.}
579 \label{fig:t5peThermo}
580 \end{figure}
581
582 For all of the properties computed, the trends with temperature
583 obtained when using the Ewald summation are reproduced with the SF
584 technique. One noticeable difference between the properties calculated
585 using the two methods are the lower $\Delta H_\textrm{vap}$ values
586 when using SF. This is to be expected due to the direct weakening of
587 the electrostatic interaction through forced neutralization. This
588 results in an increase of the intermolecular potential producing lower
589 values from equation (\ref{eq:DeltaHVap}). The slopes of these values
590 with temperature are similar to that seen using the Ewald summation;
591 however, they are both steeper than the experimental trend, indirectly
592 resulting in the inflated $C_p$ values at all temperatures.
593
594 Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
595 all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
596 and $C_p$ results, the deviations between experiment and simulation in
597 this region are not the fault of the electrostatic summation methods
598 but are due to the geometry and parameters of the TIP5P class of water
599 models. Like most rigid, non-polarizable, point-charge water models,
600 the density decreases with temperature at a much faster rate than
601 experiment (see figure \ref{fig:t5peDensities}). This reduced density
602 leads to the inflated compressibility and expansivity values at higher
603 temperatures seen here in figure \ref{fig:t5peThermo}. Incorporation
604 of polarizability and many-body effects are required in order for
605 water models to overcome differences between simulation-based and
606 experimentally determined densities at these higher
607 temperatures.\cite{Laasonen93,Donchev06}
608
609 At temperatures below the freezing point for experimental water, the
610 differences between SF and the Ewald summation results are more
611 apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
612 indicate a more pronounced transition in the supercooled regime,
613 particularly in the case of SF without damping. This points to the
614 onset of a more frustrated or glassy behavior for the undamped and
615 weakly-damped SF simulations of TIP5P-E at temperatures below 250~K
616 than is seen from the Ewald sum at these temperatures. Undamped SF
617 electrostatics has a stronger contribution from nearby charges.
618 Damping these local interactions or using a reciprocal-space method
619 makes the water less sensitive to ordering on a shorter length scale.
620 We can recover nearly quantitative agreement with the Ewald results by
621 increasing the damping constant.
622
623 The final thermodynamic property displayed in figure
624 \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
625 between the Ewald and SF methods (and with experiment). It is known
626 that the dielectric constant is dependent upon and is quite sensitive
627 to the imposed boundary conditions.\cite{Neumann80,Neumann83} This is
628 readily apparent in the converged $\epsilon$ values accumulated for
629 the SF simulations. Lack of a damping function results in dielectric
630 constants significantly smaller than those obtained using the Ewald
631 sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
632 agreement considerably. It should be noted that the choice of the
633 ``Ewald coefficient'' ($\kappa$) and real-space cutoff values also
634 have a significant effect on the calculated static dielectric constant
635 when using the Ewald summation. In the simulations of TIP5P-E with the
636 Ewald sum, this screening parameter was tethered to the simulation box
637 size (as was the $R_\textrm{c}$).\cite{Rick04} In general, systems
638 with larger screening parameters reported larger dielectric constant
639 values, the same behavior we see here with SF; however, the
640 choice of cutoff radius also plays an important role.
641
642 \subsection{Optimal Damping Coefficients for Damped
643 Electrostatics}\label{sec:t5peDielectric}
644
645 In the previous section, we observed that the choice of damping
646 coefficient plays a major role in the calculated dielectric constant
647 for the SF method. Similar damping parameter behavior was observed in
648 the long-time correlated motions of the NaCl crystal.\cite{Fennell06}
649 The static dielectric constant is calculated from the long-time
650 fluctuations of the system's accumulated dipole moment
651 (Eq. (\ref{eq:staticDielectric})), so it is quite sensitive to the
652 choice of damping parameter. Since $\alpha$ is an adjustable
653 parameter, it would be best to choose optimal damping constants such
654 that any arbitrary choice of cutoff radius will properly capture the
655 dielectric behavior of the liquid.
656
657 In order to find these optimal values, we mapped out the static
658 dielectric constant as a function of both the damping parameter and
659 cutoff radius for TIP5P-E and for a point-dipolar water model
660 (SSD/RF). To calculate the static dielectric constant, we performed
661 5~ns $NPT$ calculations on systems of 512 water molecules and averaged
662 over the converged region (typically the latter 2.5~ns) of equation
663 (\ref{eq:staticDielectric}). The selected cutoff radii ranged from 9,
664 10, 11, and 12~\AA , and the damping parameter values ranged from 0.1
665 to 0.35~\AA$^{-1}$.
666
667 \begin{table}
668 \centering
669 \caption{Static dielectric constants for the TIP5P-E and SSD/RF water models at 298~K and 1~atm as a function of damping coefficient $\alpha$ and
670 cutoff radius $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
671 \vspace{6pt}
672 \begin{tabular}{ lccccccccc }
673 \toprule
674 \toprule
675 & \multicolumn{4}{c}{TIP5P-E} & & \multicolumn{4}{c}{SSD/RF} \\
676 & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} & & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
677 \cmidrule(lr){2-5} \cmidrule(lr){7-10}
678 $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 & & 9 & 10 & 11 & 12 \\
679 \midrule
680 0.35 & \cellcolor[rgb]{1, 0.788888888888889, 0.5} 87.0 & \cellcolor[rgb]{1, 0.695555555555555, 0.5} 91.2 & \cellcolor[rgb]{1, 0.717777777777778, 0.5} 90.2 & \cellcolor[rgb]{1, 0.686666666666667, 0.5} 91.6 & & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 119.2 & \cellcolor[rgb]{1, 0.5, 0.5} 131.4 & \cellcolor[rgb]{1, 0.5, 0.5} 130 \\
681 & \cellcolor[rgb]{1, 0.892222222222222, 0.5} & \cellcolor[rgb]{1, 0.704444444444444, 0.5} & \cellcolor[rgb]{1, 0.72, 0.5} & \cellcolor[rgb]{1, 0.6666666666667, 0.5} & & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} & \cellcolor[rgb]{1, 0.5, 0.5} \\
682 0.3 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.713333333333333, 0.5} 90.4 & \cellcolor[rgb]{1, 0.646666666666667, 0.5} 93.4 & & \cellcolor[rgb]{1, 0.5, 0.5} 100 & \cellcolor[rgb]{1, 0.5, 0.5} 118.8 & \cellcolor[rgb]{1, 0.5, 0.5} 116 & \cellcolor[rgb]{1, 0.5, 0.5} 122 \\
683 0.275 & \cellcolor[rgb]{0.653333333333333, 1, 0.5} 61.9 & \cellcolor[rgb]{1, 0.933333333333333, 0.5} 80.5 & \cellcolor[rgb]{1, 0.811111111111111, 0.5} 86.0 & \cellcolor[rgb]{1, 0.766666666666667, 0.5} 88 & & \cellcolor[rgb]{1, 1, 0.5} 77.5 & \cellcolor[rgb]{1, 0.5, 0.5} 105 & \cellcolor[rgb]{1, 0.5, 0.5} 118 & \cellcolor[rgb]{1, 0.5, 0.5} 125.2 \\
684 0.25 & \cellcolor[rgb]{0.537777777777778, 1, 0.5} 56.7 & \cellcolor[rgb]{0.795555555555555, 1, 0.5} 68.3 & \cellcolor[rgb]{1, 0.966666666666667, 0.5} 79 & \cellcolor[rgb]{1, 0.704444444444445, 0.5} 90.8 & & \cellcolor[rgb]{0.5, 1, 0.582222222222222} 51.3 & \cellcolor[rgb]{1, 0.993333333333333, 0.5} 77.8 & \cellcolor[rgb]{1, 0.522222222222222, 0.5} 99 & \cellcolor[rgb]{1, 0.5, 0.5} 113 \\
685 0.225 & \cellcolor[rgb]{0.5, 1, 0.768888888888889} 42.9 & \cellcolor[rgb]{0.566666666666667, 1, 0.5} 58.0 & \cellcolor[rgb]{0.693333333333333, 1, 0.5} 63.7 & \cellcolor[rgb]{1, 0.937777777777778, 0.5} 80.3 & & \cellcolor[rgb]{0.5, 0.984444444444444, 1} 31.8 & \cellcolor[rgb]{0.5, 1, 0.586666666666667} 51.1 & \cellcolor[rgb]{1, 0.995555555555556, 0.5} 77.7 & \cellcolor[rgb]{1, 0.5, 0.5} 108.1 \\
686 0.2 & \cellcolor[rgb]{0.5, 0.973333333333333, 1} 31.3 & \cellcolor[rgb]{0.5, 1, 0.842222222222222} 39.6 & \cellcolor[rgb]{0.54, 1, 0.5} 56.8 & \cellcolor[rgb]{0.735555555555555, 1, 0.5} 65.6 & & \cellcolor[rgb]{0.5, 0.698666666666667, 1} 18.94 & \cellcolor[rgb]{0.5, 0.946666666666667, 1} 30.1 & \cellcolor[rgb]{0.5, 1, 0.704444444444445} 45.8 & \cellcolor[rgb]{0.893333333333333, 1, 0.5} 72.7 \\
687 & \cellcolor[rgb]{0.5, 0.848888888888889, 1} & \cellcolor[rgb]{0.5, 0.973333333333333, 1} & \cellcolor[rgb]{0.5, 1, 0.793333333333333} & \cellcolor[rgb]{0.5, 1, 0.624444444444445} & & \cellcolor[rgb]{0.5, 0.599333333333333, 1} & \cellcolor[rgb]{0.5, 0.732666666666667, 1} & \cellcolor[rgb]{0.5, 0.942111111111111, 1} & \cellcolor[rgb]{0.5, 1, 0.695555555555556} \\
688 0.15 & \cellcolor[rgb]{0.5, 0.724444444444444, 1} 20.1 & \cellcolor[rgb]{0.5, 0.788888888888889, 1} 23.0 & \cellcolor[rgb]{0.5, 0.873333333333333, 1} 26.8 & \cellcolor[rgb]{0.5, 1, 0.984444444444444} 33.2 & & \cellcolor[rgb]{0.5, 0.5, 1} 8.29 & \cellcolor[rgb]{0.5, 0.518666666666667, 1} 10.84 & \cellcolor[rgb]{0.5, 0.588666666666667, 1} 13.99 & \cellcolor[rgb]{0.5, 0.715555555555556, 1} 19.7 \\
689 & \cellcolor[rgb]{0.5, 0.696111111111111, 1} & \cellcolor[rgb]{0.5, 0.736333333333333, 1} & \cellcolor[rgb]{0.5, 0.775222222222222, 1} & \cellcolor[rgb]{0.5, 0.860666666666667, 1} & & \cellcolor[rgb]{0.5, 0.5, 1} & \cellcolor[rgb]{0.5, 0.509333333333333, 1} & \cellcolor[rgb]{0.5, 0.544333333333333, 1} & \cellcolor[rgb]{0.5, 0.607777777777778, 1} \\
690 0.1 & \cellcolor[rgb]{0.5, 0.667777777777778, 1} 17.55 & \cellcolor[rgb]{0.5, 0.683777777777778, 1} 18.27 & \cellcolor[rgb]{0.5, 0.677111111111111, 1} 17.97 & \cellcolor[rgb]{0.5, 0.705777777777778, 1} 19.26 & & \cellcolor[rgb]{0.5, 0.5, 1} 4.96 & \cellcolor[rgb]{0.5, 0.5, 1} 5.46 & \cellcolor[rgb]{0.5, 0.5, 1} 6.04 & \cellcolor[rgb]{0.5,0.5, 1} 6.82 \\
691 \bottomrule
692 \end{tabular}
693 \label{tab:DielectricMap}
694 \end{table}
695
696 The results of these calculations are displayed in table
697 \ref{tab:DielectricMap}. The dielectric constants for both models
698 decrease with increasing cutoff radii ($R_\textrm{c}$) and increase
699 with increasing damping ($\alpha$). Another point to note is that
700 choosing $\alpha$ and $R_\textrm{c}$ identical to those used with the
701 Ewald summation results in the same calculated dielectric constant. As
702 an example, in the paper outlining the development of TIP5P-E, the
703 real-space cutoff and Ewald coefficient were tethered to the system
704 size, and for a 512 molecule system are approximately 12~\AA\ and
705 0.25~\AA$^{-1}$ respectively.\cite{Rick04} These parameters resulted
706 in a dielectric constant of 92$\pm$14, while with SF these parameters
707 give a dielectric constant of 90.8$\pm$0.9. Another example comes from
708 the TIP4P-Ew paper where $\alpha$ and $R_\textrm{c}$ were chosen to be
709 9.5~\AA\ and 0.35~\AA$^{-1}$, and these parameters resulted in a
710 dielectric constant equal to 63$\pm$1.\cite{Horn04} Calculations using
711 SF with these parameters and this water model give a dielectric
712 constant of 61$\pm$1. Since the dielectric constant is dependent on
713 $\alpha$ and $R_\textrm{c}$ with the SF technique, it might be
714 interesting to investigate the dependence of the static dielectric
715 constant on the choice of convergence parameters ($R_\textrm{c}$ and
716 $\kappa$) utilized in most implementations of the Ewald sum.
717
718 It is also apparent from this table that electrostatic damping has a
719 more pronounced effect on the dipolar interactions of SSD/RF than the
720 monopolar interactions of TIP5P-E. The dielectric constant covers a
721 much wider range and has a steeper slope with increasing damping
722 parameter.
723
724 Although it is tempting to choose damping parameters equivalent to the
725 Ewald examples to obtain quantitative agreement, the results of our
726 previous work indicate that values this high are destructive to both
727 the energetics and dynamics.\cite{Fennell06} Ideally, $\alpha$ should
728 not exceed 0.3~\AA$^{-1}$ for any of the cutoff values in this
729 range. If the optimal damping parameter is chosen to be midway between
730 0.275 and 0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff,
731 then the linear trend with $R_\textrm{c}$ will always keep $\alpha$
732 below 0.3~\AA$^{-1}$ for the studied cutoff radii. This linear
733 progression would give values of 0.2875, 0.2625, 0.2375, and
734 0.2125~\AA$^{-1}$ for cutoff radii of 9, 10, 11, and 12~\AA. Setting
735 this to be the default behavior for the damped SF technique will
736 result in consistent dielectric behavior for these and other condensed
737 molecular systems, regardless of the chosen cutoff radius. The static
738 dielectric constants for TIP5P-E simulations with 9 and 12\AA\
739 $R_\textrm{c}$ values using their respective damping parameters are
740 76$\pm$1 and 75$\pm$2. These values are lower than the values reported
741 for TIP5P-E with the Ewald sum; however, they are more in line with
742 the values reported by Mahoney {\it et al.} for TIP5P while using a
743 reaction field (RF) with an infinite RF dielectric constant
744 (81.5$\pm$1.6).\cite{Mahoney00}
745
746 Using the same linear relationship utilized with TIP5P-E above, the
747 static dielectric constants for SSD/RF with $R_\textrm{c}$ values of 9
748 and 12~\AA\ are 88$\pm$8 and 82.6$\pm$0.6. These values compare
749 favorably with the experimental value of 78.3.\cite{Malmberg56} These
750 results are also not surprising given that early studies of the SSD
751 model indicated a static dielectric constant around 81.\cite{Liu96}
752
753 As a final note on optimal damping parameters, aside from a slight
754 lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
755 not alter any of the other thermodynamic properties.
756
757 \subsection{Dynamic Properties of TIP5P-E}\label{sec:t5peDynamics}
758
759 To look at the dynamic properties of TIP5P-E when using the SF method,
760 200~ps $NVE$ simulations were performed for each temperature at the
761 average density obtained from the $NPT$ simulations. $R_\textrm{c}$
762 values of 9 and 12~\AA\ and the ideal $\alpha$ values determined above
763 (0.2875 and 0.2125~\AA$^{-1}$) were used for the damped
764 electrostatics. The self-diffusion constants (D) were calculated from
765 linear fits to the long-time portion of the mean square displacement
766 ($\langle r^{2}(t) \rangle$).\cite{Allen87}
767
768 In addition to translational diffusion, orientational relaxation times
769 were calculated for comparisons with the Ewald simulations and with
770 experiments. These values were determined from the same 200~ps $NVE$
771 trajectories used for translational diffusion by calculating the
772 orientational time correlation function,
773 \begin{equation}
774 C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\gamma(t)
775 \cdot\hat{\mathbf{u}}_i^\gamma(0)\right]\right\rangle,
776 \label{eq:OrientCorr}
777 \end{equation}
778 where $P_l$ is the Legendre polynomial of order $l$ and
779 $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
780 axis $\gamma$. The body-fixed reference frame used for our
781 orientational correlation functions has the $z$-axis running along the
782 HOH bisector, and the $y$-axis connecting the two hydrogen atoms.
783 $C_l^y$ is therefore calculated from the time evolution of a vector of
784 unit length pointing between the two hydrogen atoms.
785
786 From the orientation autocorrelation functions, we can obtain time
787 constants for rotational relaxation. The relatively short time
788 portions (between 1 and 3~ps for water) of these curves can be fit to
789 an exponential decay to obtain these constants, and they are directly
790 comparable to water orientational relaxation times from nuclear
791 magnetic resonance (NMR). The relaxation constant obtained from
792 $C_2^y(t)$ is of particular interest because it describes the
793 relaxation of the principle axis connecting the hydrogen atoms. Thus,
794 $C_2^y(t)$ can be compared to the intermolecular portion of the
795 dipole-dipole relaxation from a proton NMR signal and should provide
796 the best estimate of the NMR relaxation time constant.\cite{Impey82}
797
798 \begin{figure}
799 \centering
800 \includegraphics[width=5.8in]{./figures/t5peDynamics.pdf}
801 \caption{Diffusion constants ({\it upper}) and reorientational time
802 constants ({\it lower}) for TIP5P-E using the Ewald sum and SF
803 technique compared with experiment. Data at temperatures less than
804 0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
805 easier comparisons in the more relevant temperature regime.}
806 \label{fig:t5peDynamics}
807 \end{figure}
808 Results for the diffusion constants and orientational relaxation times
809 are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
810 apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
811 the Ewald sum are reproduced with the SF technique. The enhanced
812 diffusion (relative to experiment) at high temperatures are again the
813 product of the lower simulated densities and do not provide any
814 special insight into differences between the electrostatic summation
815 techniques. Though not apparent in this figure, SF values at the
816 lowest temperature are approximately twice as slow as $D$ values
817 obtained using the Ewald sum. These values support the observation
818 from section \ref{sec:t5peThermo} that the SF simulations result in a
819 slightly more viscous supercooled region than is obtained using the
820 Ewald sum.
821
822 The $\tau_2^y$ results in the lower frame of figure
823 \ref{fig:t5peDynamics} show a much greater difference between the SF
824 results and the Ewald results. At all temperatures shown, TIP5P-E
825 relaxes faster than experiment with the Ewald sum while tracking
826 experiment fairly well when using the SF technique (independent of the
827 choice of damping constant). There are several possible reasons for
828 this deviation between techniques. The Ewald results were calculated
829 using shorter trajectories (10~ps) than the SF results (200~ps).
830 Calculation of these SF values from 10~ps trajectories (with
831 subsequently lower accuracy) showed a 0.4~ps drop in $\tau_2^y$,
832 placing the result more in line with that obtained using the Ewald
833 sum. Recomputing correlation times to meet a lower statistical
834 standard is counter-productive, however. Assuming the Ewald results
835 are not entirely the product of poor statistics, differences in
836 techniques to integrate the orientational motion could also play a
837 role. {\sc shake} is the most commonly used technique for
838 approximating rigid-body orientational motion,\cite{Ryckaert77}
839 whereas in {\sc oopse}, we maintain and integrate the entire rotation
840 matrix using the {\sc dlm} method.\cite{Meineke05} Since {\sc shake}
841 is an iterative constraint technique, if the convergence tolerances
842 are raised for increased performance, error will accumulate in the
843 orientational motion. Finally, the Ewald results were calculated using
844 the $NVT$ ensemble, while the $NVE$ ensemble was used for SF
845 calculations. The motion due to the extended variable (the thermostat)
846 will always alter the dynamics, resulting in differences between $NVT$
847 and $NVE$ results. These differences are increasingly noticeable as
848 the time constant for the thermostat decreases.
849
850 \subsection{Comparison of Reaction Field and Damped Electrostatics for
851 SSD/RF}
852
853 SSD/RF was parametrized for use with a reaction field, which is a
854 common and relatively inexpensive way of handling long-range
855 electrostatic corrections in dipolar systems.\cite{Onsager36}
856 Although there is no reason to expect that damped electrostatics will
857 behave in a similar fashion to the reaction field, it is well known
858 that model that are parametrized for use with Ewald do better than
859 unoptimized models under the influence of a reaction
860 field.\cite{Rick04} We compared a number of properties of SSD/RF that
861 had previously been computed using a reaction field with those same
862 values under damped electrostatics.
863
864 The properties shown in table \ref{tab:dampedSSDRF} show that using
865 damped electrostatics can result in even better agreement with
866 experiment than is obtained via reaction field. The average density
867 shows a modest increase when using damped electrostatics in place of
868 the reaction field. This comes about because we neglect the pressure
869 effect due to the surroundings outside of the cutoff, instead relying
870 on screening effects to neutralize electrostatic interactions at long
871 distances. The $C_p$ also shows a slight increase, indicating greater
872 fluctuation in the enthalpy at constant pressure. The only other
873 differences between the damped electrostatics and the reaction field
874 results are the dipole reorientational time constants, $\tau_1$ and
875 $\tau_2$. When using damped electrostatics, the water molecules relax
876 more quickly and exhibit behavior closer to the experimental
877 values. These results indicate that since there is no need to specify
878 an external dielectric constant with the damped electrostatics, it is
879 almost certainly a better choice for dipolar simulations than the
880 reaction field method. Additionally, by using damped electrostatics
881 instead of reaction field, SSD/RF can be used effectively with mixed
882 charge / dipolar systems, such as dissolved ions, dissolved organic
883 molecules, or even proteins.
884
885 \begin{table}
886 \caption{Properties of SSD/RF when using reaction field or damped
887 electrostatics. Simulations were carried out at 298~K, 1~atm, and
888 with 512 molecules.}
889 \footnotesize
890 \centering
891 \begin{tabular}{ llccc }
892 \toprule
893 \toprule
894 & & Reaction Field (Ref. \citen{Fennell04}) & Damped Electrostatics &
895 Experiment [Ref.] \\
896 & & $\epsilon = 80$ & $R_\textrm{c} = 12$\AA ; $\alpha = 0.2125$~\AA$^{-1}$ & \\
897 \midrule
898 $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 [\citen{CRC80}]\\
899 $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 [\citen{Wagner02}] \\
900 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 [\citen{Mills73}]\\
901 $n_C$ & & 4.4 & 4.2 & 4.7 [\citen{Hura00}]\\
902 $n_H$ & & 3.7 & 3.7 & 3.5 [\citen{Soper86}]\\
903 $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 [\citen{Eisenberg69}]\\
904 $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 [\citen{Krynicki66}]\\
905 \bottomrule
906 \end{tabular}
907 \label{tab:dampedSSDRF}
908 \end{table}
909
910 \subsection{Predictions of Ice Polymorph Stability}
911
912 In an earlier paper, we performed a series of free energy calculations
913 on several ice polymorphs which are stable or metastable at low
914 pressures, one of which (Ice-$i$) we observed in spontaneous
915 crystallizations of an early version of the SSD/RF water
916 model.\cite{Fennell05} In this study, a distinct dependence of the
917 computed free energies on the cutoff radius and electrostatic
918 summation method was observed. Since the SF technique can be used as
919 a simple and efficient replacement for the Ewald summation, our final
920 test of this method is to see if it is capable of addressing the
921 spurious stability of the Ice-$i$ phases with the various common water
922 models. To this end, we have performed thermodynamic integrations of
923 all of the previously discussed ice polymorphs using the SF technique
924 with a cutoff radius of 12~\AA\ and an $\alpha$ of 0.2125~\AA . These
925 calculations were performed on TIP5P-E and TIP4P-Ew (variants of the
926 TIP5P and TIP4P models optimized for the Ewald summation) as well as
927 SPC/E and SSD/RF.
928
929 \begin{table}
930 \centering
931 \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
932 using the damped SF electrostatic correction method with a
933 variety of water models.}
934 \begin{tabular}{ lccccc }
935 \toprule
936 \toprule
937 Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
938 \cmidrule(lr){2-6}
939 & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
940 \midrule
941 TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
942 TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
943 SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
944 SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
945 \bottomrule
946 \end{tabular}
947 \label{tab:dampedFreeEnergy}
948 \end{table}
949 The results of these calculations in table \ref{tab:dampedFreeEnergy}
950 show similar behavior to the Ewald results in the previous
951 study.\cite{Fennell05} The Helmholtz free energies of the ice
952 polymorphs with SSD/RF order in the same fashion, with Ice-$i$ having
953 the lowest free energy; however, the Ice-$i$ and ice B free energies
954 are quite a bit closer (nearly isoenergetic). The SPC/E results show
955 the different polymorphs to be nearly isoenergetic. This is the same
956 behavior observed using an Ewald correction.\cite{Fennell05} Ice B has
957 the lowest Helmholtz free energy for SPC/E; however, all the polymorph
958 results overlap within the error estimates.
959
960 The most interesting results from these calculations come from the
961 more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
962 optimized for use with an electrostatic correction and are
963 geometrically arranged to mimic water using drastically different
964 charge distributions. In TIP5P-E, the primary location for the
965 negative charge in the molecule is assigned to the lone-pairs of the
966 oxygen, while TIP4P-Ew places the negative charge near the
967 center-of-mass along the H-O-H bisector. There is some debate as to
968 which is the proper choice for the negative charge location, and this
969 has in part led to a six-site water model that balances both of these
970 options.\cite{Vega05,Nada03} The limited results in table
971 \ref{tab:dampedFreeEnergy} support the results of Vega {\it et al.},
972 which indicate the TIP4P charge location geometry performs better
973 under some circumstances.\cite{Vega05} With the TIP4P-Ew water model,
974 the experimentally observed polymorph (ice I$_\textrm{h}$) is the
975 preferred form with ice I$_\textrm{c}$ slightly higher in energy,
976 though overlapping within error. Additionally, the spurious ice B and
977 Ice-$i^\prime$ structures are destabilized relative to these
978 polymorphs. TIP5P-E shows similar behavior to SPC/E, where there is no
979 real free energy distinction between the various polymorphs. While ice
980 B is close in free energy to the other polymorphs, these results fail
981 to support the findings of other researchers indicating the preferred
982 form of TIP5P at 1~atm is a structure similar to ice
983 B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we were
984 looking at TIP5P-E rather than TIP5P, and the differences in the
985 Lennard-Jones parameters could cause this discrepancy. Overall, these
986 results indicate that TIP4P-Ew is a better mimic of the solid forms of
987 water than some of the other models.
988
989 \section{Conclusions}
990
991 This investigation of pairwise electrostatic summation techniques
992 shows that there is a viable and computationally efficient alternative
993 to the Ewald summation. The SF method (equation (\ref{eq:DSFPot}))
994 has proven itself capable of reproducing structural, thermodynamic,
995 and dynamic quantities that are nearly quantitative matches to results
996 from far more expensive methods. Additionally, we have now extended
997 the damping formalism to electrostatic multipoles, so the damped SF
998 potential can be used in systems that contain mixtures of charges and
999 point multipoles.
1000
1001 We have also provided a simple linear prescription for choosing
1002 optimal damping parameters given a choice of cutoff radius. The
1003 damping parameters were chosen to obtain a static dielectric constant
1004 as close as possible to the experimental value, which should be useful
1005 for simulating the electrostatic screening properties of liquid water
1006 accurately. The linear formula for optimal damping was the same for
1007 a complicated multipoint model as it was for a simple point-dipolar
1008 model.
1009
1010 As in all purely pairwise cutoff methods, the damped SF method is
1011 expected to scale approximately {\it linearly} with system size, and
1012 is easily parallelizable. This should result in substantial
1013 reductions in the computational cost of performing large simulations.
1014 With the proper use of pre-computation and spline interpolation, the
1015 damped SF method is essentially the same cost as a simple real-space
1016 cutoff.
1017
1018 We are not suggesting that there is any flaw with the Ewald sum; in
1019 fact, it is the standard by which the damped SF method has been
1020 judged. However, these results provide further evidence that in the
1021 typical simulations performed today, the Ewald summation may no longer
1022 be required to obtain the level of accuracy most researchers have come
1023 to expect.
1024
1025 \section{Acknowledgments}
1026 Support for this project was provided by the National Science
1027 Foundation under grant CHE-0134881. Computation time was provided by
1028 the Notre Dame Center for Research Computing. The authors would like
1029 to thank Steve Corcelli and Ed Maginn for helpful discussions and
1030 comments.
1031
1032 \newpage
1033
1034 \bibliographystyle{achemso}
1035 \bibliography{multipoleSFPaper}
1036
1037
1038 \end{document}