ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipoleSFPaper/multipoleSFPaper.tex
Revision: 3066
Committed: Tue Oct 24 16:11:42 2006 UTC (17 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 54396 byte(s)
Log Message:
working on the paper

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 \\
37 Department of Chemistry and Biochemistry\\ University of Notre Dame\\
38 Notre Dame, Indiana 46556}
39
40 \date{\today}
41
42 \maketitle
43 %\doublespacing
44
45 \begin{abstract}
46 The damped, shifted-force electrostatic potential has been shown to
47 give nearly quantitative agreement with smooth particle mesh Ewald for
48 energy differences between configurations as well as for atomic force
49 and molecular torque vectors. In this paper, we extend this technique
50 to handle interactions between electrostatic multipoles. We also
51 investigate the effects of damped and shifted electrostatics on the
52 static, thermodynamic, and dynamic properties of liquid water and
53 various polymorphs of ice. We provide a way of choosing the optimal
54 damping strength for a given cutoff radius that reproduces the static
55 dielectric constant in a variety of water models.
56 \end{abstract}
57
58 %\narrowtext
59
60 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61 % BODY OF TEXT
62 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63
64 \section{Introduction}
65
66 Over the past several years, there has been increasing interest in
67 pairwise methods for correcting electrostatic interactions in computer
68 simulations of condensed molecular
69 systems.\cite{Wolf99,Zahn02,Kast03,Beck05,Ma05,Fennell06} These
70 techniques were developed from the observations and efforts of Wolf
71 {\it et al.} and their work towards an $\mathcal{O}(N)$ Coulombic
72 sum.\cite{Wolf99} Wolf's method of cutoff neutralization is able to
73 obtain excellent agreement with Madelung energies in ionic
74 crystals.\cite{Wolf99}
75
76 In a recent paper, we showed that simple modifications to Wolf's
77 method could give nearly quantitative agreement with smooth particle
78 mesh Ewald (SPME) for quantities of interest in Monte Carlo
79 (i.e. configurational energy differences) and Molecular Dynamics
80 (i.e. atomic force and molecular torque vectors).\cite{Fennell06} We
81 described the undamped and damped shifted potential (SP) and shifted
82 force (SF) techniques. The potential for damped form of the SF method
83 is given by
84 \begin{equation}
85 \begin{split}
86 V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
87 \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
88 - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
89 &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
90 + \frac{2\alpha}{\pi^{1/2}}
91 \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
92 \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
93 \quad r\leqslant R_\textrm{c}.
94 \label{eq:DSFPot}
95 \end{split}
96 \end{equation}
97 (In this potential and in all electrostatic quantities that follow,
98 the standard $4 \pi \epsilon_{0}$ has been omitted for clarity.)
99
100 The damped SF method is an improvement over the SP method because
101 there is no discontinuity in the forces as particles move out of the
102 cutoff radius ($R_\textrm{c}$). This is accomplished by shifting the
103 forces (and potential) to zero at $R_\textrm{c}$. This is analogous to
104 neutralizing the charge as well as the force effect of the charges
105 within $R_\textrm{c}$.
106
107 To complete the charge neutralization procedure, a self-neutralization
108 term needs to be included in the potential. This term is constant (as
109 long as the charges and cutoff radius do not change), and exists
110 outside the normal pair-loop. It can be thought of as a contribution
111 from a charge opposite in sign and equal in magnitude to the central
112 charge, but which has been spread out over the surface of the cutoff
113 sphere. This term is calculated via a single loop over all charges in
114 the system. For the undamped case, the self term can be written as
115 \begin{equation}
116 V_\textrm{self} = \frac{1}{2 R_\textrm{c}} \sum_{i=1}^N q_i^{2},
117 \label{eq:selfTerm}
118 \end{equation}
119 while for the damped case it can be written as
120 \begin{equation}
121 V_\textrm{self} = \left(\frac{\alpha}{\sqrt{\pi}}
122 + \frac{\textrm{erfc}(\alpha
123 R_\textrm{c})}{2R_\textrm{c}}\right) \sum_{i=1}^N q_i^{2}.
124 \label{eq:dampSelfTerm}
125 \end{equation}
126 The first term within the parentheses of equation
127 (\ref{eq:dampSelfTerm}) is identical to the self term in the Ewald
128 summation, and comes from the utilization of the complimentary error
129 function for electrostatic damping.\cite{deLeeuw80,Wolf99}
130
131 The SF, SP, and Wolf methods operate by neutralizing the total charge
132 contained within the cutoff sphere surrounding each particle. This is
133 accomplished by creating image charges on the surface of the cutoff
134 sphere for each pair interaction computed within the sphere. The
135 damping function applied to the potential is also an important method
136 for accelerating convergence. In the case of systems involving
137 electrostatic distributions of higher order than point charges, the
138 question remains: How will the shifting and damping need to be
139 modified in order to accommodate point multipoles?
140
141 \section{Electrostatic Damping for Point
142 Multipoles}\label{sec:dampingMultipoles}
143
144 To apply the SF method for systems involving point multipoles, we
145 consider separately the two techniques (shifting and damping) which
146 contribute to the effectiveness of the DSF potential.
147
148 As noted above, shifting the potential and forces is employed to
149 neutralize the total charge contained within each cutoff sphere;
150 however, in a system composed purely of point multipoles, each cutoff
151 sphere is already neutral, so shifting becomes unnecessary.
152
153 In a mixed system of charges and multipoles, the undamped SF potential
154 needs only to shift the force terms between charges and smoothly
155 truncate the multipolar interactions with a switching function. The
156 switching function is required for energy conservation, because a
157 discontinuity will exist in both the potential and forces at
158 $R_\textrm{c}$ in the absence of shifting terms.
159
160 To damp the SF potential for point multipoles, we need to incorporate
161 the complimentary error function term into the standard forms of the
162 multipolar potentials. We can determine the necessary damping
163 functions by replacing $1/r$ with $\mathrm{erfc}(\alpha r)/r$ in the
164 multipole expansion. This procedure quickly becomes quite complex
165 with ``two-center'' systems, like the dipole-dipole potential, and is
166 typically approached using spherical harmonics.\cite{Hirschfelder67} A
167 simpler method for determining damped multipolar interaction
168 potentials arises when we adopt the tensor formalism described by
169 Stone.\cite{Stone02}
170
171 The tensor formalism for electrostatic interactions involves obtaining
172 the multipole interactions from successive gradients of the monopole
173 potential. Thus, tensors of rank zero through two are
174 \begin{equation}
175 T = \frac{1}{r_{ij}},
176 \label{eq:tensorRank1}
177 \end{equation}
178 \begin{equation}
179 T_\alpha = \nabla_\alpha \frac{1}{r_{ij}},
180 \label{eq:tensorRank2}
181 \end{equation}
182 \begin{equation}
183 T_{\alpha\beta} = \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
184 \label{eq:tensorRank3}
185 \end{equation}
186 where the form of the first tensor is the charge-charge potential, the
187 second gives the charge-dipole potential, and the third gives the
188 charge-quadrupole and dipole-dipole potentials.\cite{Stone02} Since
189 the force is $-\nabla V$, the forces for each potential come from the
190 next higher tensor.
191
192 As one would do with the multipolar expansion, we can replace $r^{-1}$
193 with $\mathrm{erfc}(\alpha r)\cdot r^{-1}$ to obtain damped forms of the
194 electrostatic potentials. Equation \ref{eq:tensorRank2} generates a
195 damped charge-dipole potential,
196 \begin{equation}
197 V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
198 \label{eq:dChargeDipole}
199 \end{equation}
200 where $c_1(r_{ij})$ is
201 \begin{equation}
202 c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
203 + \textrm{erfc}(\alpha r_{ij}).
204 \label{eq:c1Func}
205 \end{equation}
206
207 Equation \ref{eq:tensorRank3} generates a damped dipole-dipole potential,
208 \begin{equation}
209 V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
210 (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
211 c_2(r_{ij}) -
212 \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
213 c_1(r_{ij}),
214 \label{eq:dampDipoleDipole}
215 \end{equation}
216 where
217 \begin{equation}
218 c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
219 + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
220 + \textrm{erfc}(\alpha r_{ij}).
221 \end{equation}
222 Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
223 term. Continuing with higher rank tensors, we can obtain the damping
224 functions for higher multipole potentials and forces. Each subsequent
225 damping function includes one additional term, and we can simplify the
226 procedure for obtaining these terms by writing out the following
227 generating function,
228 \begin{equation}
229 c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
230 {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
231 \label{eq:dampingGeneratingFunc}
232 \end{equation}
233 where,
234 \begin{equation}
235 m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
236 m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
237 m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
238 1 & m = -1\textrm{ or }0,
239 \end{array}\right.
240 \label{eq:doubleFactorial}
241 \end{equation}
242 and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
243 is similar in form to those obtained by Smith and Aguado and Madden
244 for the application of the Ewald sum to
245 multipoles.\cite{Smith82,Smith98,Aguado03}
246
247 Returning to the dipole-dipole example, the potential consists of a
248 portion dependent upon $r^{-5}$ and another dependent upon
249 $r^{-3}$.
250 $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
251 respectively. The forces for the damped dipole-dipole interaction, are
252 obtained from the next higher tensor, $T_{\alpha \beta \gamma}$,
253 \begin{equation}
254 \begin{split}
255 F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
256 (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
257 c_3(r_{ij})\\ &-
258 3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
259 \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
260 \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
261 {r^5_{ij}} c_2(r_{ij}),
262 \end{split}
263 \label{eq:dampDipoleDipoleForces}
264 \end{equation}
265 Using the tensor formalism, we can dampen higher order multipolar
266 interactions using the same effective damping function that we use for
267 charge-charge interactions. This allows us to include multipoles in
268 simulations involving damped electrostatic interactions.
269
270 \section{Applications of DSF Electrostatics}
271
272 Our earlier work on the SF method showed that it can give nearly
273 quantitive agreement with SPME-derived configurational energy
274 differences. The force and torque vectors in identical configurations
275 are also nearly equivalent under the damped SF potential and
276 SPME.\cite{Fennell06} Although these measures bode well for the
277 performance of the SF method in both Monte Carlo and Molecular
278 Dynamics simulations, it would be helpful to have direct comparisons
279 of structural, thermodynamic, and dynamic quantities. To address
280 this, we performed a detailed analysis of a group of simulations
281 involving water models (both point charge and multipolar) under a
282 number of different simulation conditions.
283
284 To provide the most difficult test for the damped SF method, we have
285 chosen a model that has been optimized for use with Ewald sum, and
286 have compared the simulated properties to those computed via Ewald.
287 It is well known that water models parametrized for use with the Ewald
288 sum give calculated properties that are in better agreement with
289 experiment.\cite{vanderSpoel98,Horn04,Rick04} For these reasons, we
290 chose the TIP5P-E water model for our comparisons involving point
291 charges.\cite{Rick04}
292
293 The soft sticky dipole (SSD) family of water models is the perfect
294 test case for the point-multipolar extension of damped electrostatics.
295 SSD water models are single point molecules that consist of a ``soft''
296 Lennard-Jones sphere, a point-dipole, and a tetrahedral function for
297 capturing the hydrogen bonding nature of water - a spherical harmonic
298 term for water-water tetrahedral interactions and a point-quadrupole
299 for interactions with surrounding charges. Detailed descriptions of
300 these models can be found in other
301 studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
302
303 In deciding which version of the SSD model to use, we need only
304 consider that the SF technique was presented as a pairwise replacement
305 for the Ewald summation. It has been suggested that models
306 parametrized for the Ewald summation (like TIP5P-E) would be
307 appropriate for use with a reaction field and vice versa.\cite{Rick04}
308 Therefore, we decided to test the SSD/RF water model, which was
309 parametrized for use with a reaction field, with the damped
310 electrostatic technique to see how the calculated properties change.
311
312 The TIP5P-E water model is a variant of Mahoney and Jorgensen's
313 five-point transferable intermolecular potential (TIP5P) model for
314 water.\cite{Mahoney00} TIP5P was developed to reproduce the density
315 maximum in liquid water near 4$^\circ$C. As with many previous point
316 charge water models (such as ST2, TIP3P, TIP4P, SPC, and SPC/E), TIP5P
317 was parametrized using a simple cutoff with no long-range
318 electrostatic
319 correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
320 Without this correction, the pressure term on the central particle
321 from the surroundings is missing. When this correction is included,
322 systems of these particles expand to compensate for this added
323 pressure term and under-predict the density of water under standard
324 conditions. In developing TIP5P-E, Rick preserved the geometry and
325 point charge magnitudes in TIP5P and focused on altering the
326 Lennard-Jones parameters to correct the density at 298~K. With the
327 density corrected, he compared common water properties for TIP5P-E
328 using the Ewald sum with TIP5P using a 9~\AA\ cutoff.
329
330 In the following sections, we compare these same properties calculated
331 from TIP5P-E using the Ewald sum with TIP5P-E using the damped SF
332 technique. Our comparisons include the SF technique with a 12~\AA\
333 cutoff and an $\alpha$ = 0.0, 0.1, and 0.2~\AA$^{-1}$, as well as a
334 9~\AA\ cutoff with an $\alpha$ = 0.2~\AA$^{-1}$.
335
336 \subsection{The Density Maximum of TIP5P-E}\label{sec:t5peDensity}
337
338 To compare densities, $NPT$ simulations were performed with the same
339 temperatures as those selected by Rick in his Ewald summation
340 simulations.\cite{Rick04} In order to improve statistics around the
341 density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
342 25$^\circ$C, while 2~ns trajectories were obtained at all other
343 temperatures. The average densities were calculated from the later
344 three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
345 method for accumulating statistics, these sequences were spliced into
346 200 segments, each providing an average density. These 200 density
347 values were used to calculate the average and standard deviation of
348 the density at each temperature.\cite{Mahoney00}
349
350 \begin{figure}
351 \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
352 \caption{Density versus temperature for the TIP5P-E water model when
353 using the Ewald summation (Ref. \citen{Rick04}) and the SF method with
354 varying cutoff radii and damping coefficients. The pressure term from
355 the image-charge shell is larger than that provided by the
356 reciprocal-space portion of the Ewald summation, leading to slightly
357 lower densities. This effect is more visible with the 9~\AA\ cutoff,
358 where the image charges exert a greater force on the central
359 particle. The error bars for the SF methods show the average one-sigma
360 uncertainty of the density measurement, and this uncertainty is the
361 same for all the SF curves.}
362 \label{fig:t5peDensities}
363 \end{figure}
364 Figure \ref{fig:t5peDensities} shows the densities calculated for
365 TIP5P-E using differing electrostatic corrections overlaid with the
366 experimental values.\cite{CRC80} The densities when using the SF
367 technique are close to, but typically lower than, those calculated
368 using the Ewald summation. These slightly reduced densities indicate
369 that the pressure component from the image charges at R$_\textrm{c}$
370 is larger than that exerted by the reciprocal-space portion of the
371 Ewald summation. Bringing the image charges closer to the central
372 particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
373 preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
374 charge interactions on the central particle and results in a further
375 reduction of the densities.
376
377 Because the strength of the image charge interactions has a noticeable
378 effect on the density, we would expect the use of electrostatic
379 damping to also play a role in these calculations. Larger values of
380 $\alpha$ weaken the pair-interactions; and since electrostatic damping
381 is distance-dependent, force components from the image charges will be
382 reduced more than those from particles close the the central
383 charge. This effect is visible in figure \ref{fig:t5peDensities} with
384 the damped SF sums showing slightly higher densities; however, it is
385 clear that the choice of cutoff radius plays a much more important
386 role in the resulting densities.
387
388 All of the above density calculations were performed with systems of
389 512 water molecules. Rick observed a system size dependence of the
390 computed densities when using the Ewald summation, most likely due to
391 his tying of the convergence parameter to the box
392 dimensions.\cite{Rick04} For systems of 256 water molecules, the
393 calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
394 system size of 256 molecules would force the use of a shorter
395 R$_\textrm{c}$ when using the SF technique, and this would also lower
396 the densities. Moving to larger systems, as long as the R$_\textrm{c}$
397 remains at a fixed value, we would expect the densities to remain
398 constant.
399
400 \subsection{Liquid Structure of TIP5P-E}\label{sec:t5peLiqStructure}
401
402 The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
403 been compared in great detail with the various common water models,
404 and TIP5P was found to be in better agreement than other rigid,
405 non-polarizable models.\cite{Sorenson00} This excellent agreement with
406 experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
407 check whether the choice of using the Ewald summation or the SF
408 technique alters the liquid structure, the $g_\textrm{OO}(r)$s at
409 298~K and 1~atm were determined for the systems compared in the
410 previous section.
411
412 \begin{figure}
413 \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
414 \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and
415 1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
416 sf} technique with varying parameters. Even with the reduced densities
417 using the SF technique, the $g_\textrm{OO}(r)$s are essentially
418 identical.}
419 \label{fig:t5peGofRs}
420 \end{figure}
421 The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
422 sf} technique with a various parameters are overlaid on the
423 $g_\textrm{OO}(r)$ while using the Ewald summation in
424 figure~\ref{fig:t5peGofRs}. The differences in density do not appear
425 to have any effect on the liquid structure as the $g_\textrm{OO}(r)$s
426 are indistinguishable. These results do indicate that
427 $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
428 correction.
429
430 \subsection{Thermodynamic Properties of TIP5P-E}\label{sec:t5peThermo}
431
432 In addition to the density and structual features of the liquid, there
433 are a variety of thermodynamic quantities that can be calculated for
434 water and compared directly to experimental values. Some of these
435 additional quantities include the latent heat of vaporization ($\Delta
436 H_\textrm{vap}$), the constant pressure heat capacity ($C_p$), the
437 isothermal compressibility ($\kappa_T$), the thermal expansivity
438 ($\alpha_p$), and the static dielectric constant ($\epsilon$). All of
439 these properties were calculated for TIP5P-E with the Ewald summation,
440 so they provide a good set for comparisons involving the SF technique.
441
442 The $\Delta H_\textrm{vap}$ is the enthalpy change required to
443 transform one mole of substance from the liquid phase to the gas
444 phase.\cite{Berry00} In molecular simulations, this quantity can be
445 determined via
446 \begin{equation}
447 \begin{split}
448 \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq} \\
449 &= E_\textrm{gas} - E_\textrm{liq}
450 + P(V_\textrm{gas} - V_\textrm{liq}) \\
451 &\approx -\frac{\langle U_\textrm{liq}\rangle}{N}+ RT,
452 \end{split}
453 \label{eq:DeltaHVap}
454 \end{equation}
455 where $E$ is the total energy, $U$ is the potential energy, $P$ is the
456 pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
457 the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
458 seen in the last line of equation (\ref{eq:DeltaHVap}), we can
459 approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
460 state. This allows us to cancel the kinetic energy terms, leaving only
461 the $U_\textrm{liq}$ term. Additionally, the $pV$ term for the gas is
462 several orders of magnitude larger than that of the liquid, so we can
463 neglect the liquid $pV$ term.
464
465 The remaining thermodynamic properties can all be calculated from
466 fluctuations of the enthalpy, volume, and system dipole
467 moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
468 enthalpy in constant pressure simulations via
469 \begin{equation}
470 \begin{split}
471 C_p = \left(\frac{\partial H}{\partial T}\right)_{N,P}
472 = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
473 \end{split}
474 \label{eq:Cp}
475 \end{equation}
476 where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
477 \begin{equation}
478 \begin{split}
479 \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
480 = \frac{(\langle V^2\rangle_{NPT} - \langle V\rangle^2_{NPT})}
481 {k_BT\langle V\rangle_{NPT}},
482 \end{split}
483 \label{eq:kappa}
484 \end{equation}
485 and $\alpha_p$ can be calculated via
486 \begin{equation}
487 \begin{split}
488 \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
489 = \frac{(\langle VH\rangle_{NPT}
490 - \langle V\rangle_{NPT}\langle H\rangle_{NPT})}
491 {k_BT^2\langle V\rangle_{NPT}}.
492 \end{split}
493 \label{eq:alpha}
494 \end{equation}
495 Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
496 be calculated for systems of non-polarizable substances via
497 \begin{equation}
498 \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
499 \label{eq:staticDielectric}
500 \end{equation}
501 where $\epsilon_0$ is the permittivity of free space and $\langle
502 M^2\rangle$ is the fluctuation of the system dipole
503 moment.\cite{Allen87} The numerator in the fractional term in equation
504 (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
505 dipole moment, identical to the quantity calculated in the
506 finite-system Kirkwood $g$ factor ($G_k$):
507 \begin{equation}
508 G_k = \frac{\langle M^2\rangle}{N\mu^2},
509 \label{eq:KirkwoodFactor}
510 \end{equation}
511 where $\mu$ is the dipole moment of a single molecule of the
512 homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
513 fluctuation term in both equation (\ref{eq:staticDielectric}) and
514 \ref{eq:KirkwoodFactor} is calculated as follows,
515 \begin{equation}
516 \begin{split}
517 \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
518 - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
519 &= \langle M_x^2+M_y^2+M_z^2\rangle
520 - (\langle M_x\rangle^2 + \langle M_x\rangle^2
521 + \langle M_x\rangle^2).
522 \end{split}
523 \label{eq:fluctBoxDipole}
524 \end{equation}
525 This fluctuation term can be accumulated during the simulation;
526 however, it converges rather slowly, thus requiring multi-nanosecond
527 simulation times.\cite{Horn04} In the case of tin-foil boundary
528 conditions, the dielectric/surface term of the Ewald summation is
529 equal to zero. Since the SF method also lacks this
530 dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
531 valid for determining static dielectric constants.
532
533 All of the above properties were calculated from the same trajectories
534 used to determine the densities in section \ref{sec:t5peDensity}
535 except for the static dielectric constants. The $\epsilon$ values were
536 accumulated from 2~ns $NVE$ ensemble trajectories with system densities
537 fixed at the average values from the $NPT$ simulations at each of the
538 temperatures. The resulting values are displayed in figure
539 \ref{fig:t5peThermo}.
540 \begin{figure}
541 \centering
542 \includegraphics[width=4.5in]{./figures/t5peThermo.pdf}
543 \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
544 and the SF techniques along with the experimental values. Units
545 for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
546 cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
547 and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
548 reference \cite{Rick04}. Experimental values for $\Delta
549 H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
550 \cite{Kell75}. Experimental values for $C_p$ are from reference
551 \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
552 \cite{Malmberg56}.}
553 \label{fig:t5peThermo}
554 \end{figure}
555
556 For all of the properties computed, the trends with temperature
557 obtained when using the Ewald summation are reproduced with the SF
558 technique. One noticeable difference between the properties calculated
559 using the two methods are the lower $\Delta H_\textrm{vap}$ values
560 when using SF. This is to be expected due to the direct weakening of
561 the electrostatic interaction through forced neutralization. This
562 results in an increase of the intermolecular potential producing lower
563 values from equation (\ref{eq:DeltaHVap}). The slopes of these values
564 with temperature are similar to that seen using the Ewald summation;
565 however, they are both steeper than the experimental trend, indirectly
566 resulting in the inflated $C_p$ values at all temperatures.
567
568 Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
569 all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
570 and $C_p$ results, the deviations between experiment and simulation in
571 this region are not the fault of the electrostatic summation methods
572 but are due to the geometry and parameters of the TIP5P class of water
573 models. Like most rigid, non-polarizable, point-charge water models,
574 the density decreases with temperature at a much faster rate than
575 experiment (see figure \ref{fig:t5peDensities}). This reduced density
576 leads to the inflated compressibility and expansivity values at higher
577 temperatures seen here in figure \ref{fig:t5peThermo}. Incorporation
578 of polarizability and many-body effects are required in order for
579 water models to overcome differences between simulation-based and
580 experimentally determined densities at these higher
581 temperatures.\cite{Laasonen93,Donchev06}
582
583 At temperatures below the freezing point for experimental water, the
584 differences between SF and the Ewald summation results are more
585 apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
586 indicate a more pronounced transition in the supercooled regime,
587 particularly in the case of SF without damping.
588
589 This points to the onset of a more frustrated or glassy behavior for
590 the undamped and weakly-damped SF simulations of TIP5P-E at
591 temperatures below 250~K than is seen from the Ewald sum. Undamped SF
592 electrostatics has a stronger contribution from nearby charges.
593 Damping these local interactions or using a reciprocal-space method
594 makes the water less sensitive to ordering on a short length scale.
595 We can recover nearly quantitative agreement with the Ewald results by
596 increasing the damping constant.
597
598 The final thermodynamic property displayed in figure
599 \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
600 between the Ewald and SF methods (and with experiment). It is known
601 that the dielectric constant is dependent upon and is quite sensitive
602 to the imposed boundary conditions.\cite{Neumann80,Neumann83} This is
603 readily apparent in the converged $\epsilon$ values accumulated for
604 the SF simulations. Lack of a damping function results in dielectric
605 constants significantly smaller than those obtained using the Ewald
606 sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
607 agreement considerably. It should be noted that the choice of the
608 ``Ewald coefficient'' ($\kappa$) and real-space cutoff values also
609 have a significant effect on the calculated static dielectric constant
610 when using the Ewald summation. In the simulations of TIP5P-E with the
611 Ewald sum, this screening parameter was tethered to the simulation box
612 size (as was the $R_\textrm{c}$).\cite{Rick04} In general, systems
613 with larger screening parameters reported larger dielectric constant
614 values, the same behavior we see here with {\sc sf}; however, the
615 choice of cutoff radius also plays an important role.
616
617 \subsubsection{Dielectric Constants for TIP5P-E and SSD/RF}\label{sec:t5peDielectric}
618
619 In the previous section, we observed that the choice of damping
620 coefficient plays a major role in the calculated dielectric constant
621 for the SF method. Similar damping parameter behavior was observed in
622 the long-time correlated motions of the NaCl crystal.\cite{Fennell06}
623 The static dielectric constant is calculated from the long-time
624 fluctuations of the system's accumulated dipole moment
625 (Eq. (\ref{eq:staticDielectric})), so it is quite sensitive to the
626 choice of damping parameter. Since $\alpha$ is an adjustable
627 parameter, it would be best to choose optimal damping constants such
628 that any arbitrary choice of cutoff radius will properly capture the
629 dielectric behavior of the liquid.
630
631 In order to find these optimal values, we mapped out the static
632 dielectric constant as a function of both the damping parameter and
633 cutoff radius for TIP5P-E and for a point-dipolar water model
634 (SSD/RF). To calculate the static dielectric constant, we performed
635 5~ns $NPT$ calculations on systems of 512 water molecules and averaged
636 over the converged region (typically the later 2.5~ns) of equation
637 (\ref{eq:staticDielectric}). The selected cutoff radii ranged from 9,
638 10, 11, and 12~\AA , and the damping parameter values ranged from 0.1
639 to 0.35~\AA$^{-1}$.
640
641 \begin{table}
642 \centering
643 \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
644 cutoff radius $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
645 \vspace{6pt}
646 \begin{tabular}{ lccccccccc }
647 \toprule
648 \toprule
649 & \multicolumn{4}{c}{TIP5P-E} & & \multicolumn{4}{c}{SSD/RF} \\
650 & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} & & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
651 \cmidrule(lr){2-5} \cmidrule(lr){7-10}
652 $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 & & 9 & 10 & 11 & 12 \\
653 \midrule
654 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 \\
655 & \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} \\
656 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 \\
657 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 \\
658 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 \\
659 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 \\
660 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 \\
661 & \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} \\
662 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 \\
663 & \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} \\
664 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 \\
665 \bottomrule
666 \end{tabular}
667 \label{tab:DielectricMap}
668 \end{table}
669
670 The results of these calculations are displayed in table
671 \ref{tab:DielectricMap}. The dielectric constants for both models
672 decrease linearly with increasing cutoff radii ($R_\textrm{c}$) and
673 increase linearly with increasing damping ($\alpha$). Another point
674 to note is that choosing $\alpha$ and $R_\textrm{c}$ identical to
675 those used with the Ewald summation results in the same calculated
676 dielectric constant. As an example, in the paper outlining the
677 development of TIP5P-E, the real-space cutoff and Ewald coefficient
678 were tethered to the system size, and for a 512 molecule system are
679 approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04}
680 These parameters resulted in a dielectric constant of 92$\pm$14, while
681 with SF these parameters give a dielectric constant of
682 90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
683 $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
684 0.35~\AA$^{-1}$, and these parameters resulted in a dielectric
685 constant equal to 63$\pm$1.\cite{Horn04} Calculations using SF with
686 these parameters and this water model give a dielectric constant of
687 61$\pm$1. Since the dielectric constant is dependent on $\alpha$ and
688 $R_\textrm{c}$ with the SF technique, it might be interesting to
689 investigate the dielectric dependence of the real-space Ewald
690 parameters.
691
692 We have also mapped out the static dielectric constant of SSD/RF as a
693 function of $R_\textrm{c}$ and $\alpha$. It is apparent from this
694 table that electrostatic damping has a more pronounced effect on the
695 dipolar interactions of SSD/RF than the monopolar interactions of
696 TIP5P-E. The dielectric constant covers a much wider range and has a
697 steeper slope with increasing damping parameter.
698
699 Although it is tempting to choose damping parameters equivalent to the
700 Ewald examples, the results of our previous work indicate that values
701 this high are destructive to both the energetics and
702 dynamics.\cite{Fennell06} Ideally, $\alpha$ should not exceed
703 0.3~\AA$^{-1}$ for any of the cutoff values in this range. If the
704 optimal damping parameter is chosen to be midway between 0.275 and
705 0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff, then the
706 linear trend with $R_\textrm{c}$ will always keep $\alpha$ below
707 0.3~\AA$^{-1}$ for the studied cutoff radii. This linear progression
708 would give values of 0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for
709 cutoff radii of 9, 10, 11, and 12~\AA. Setting this to be the default
710 behavior for the damped SF technique will result in consistent
711 dielectric behavior for these and other condensed molecular systems,
712 regardless of the chosen cutoff radius. The static dielectric
713 constants for TIP5P-E simulations with 9 and 12\AA\ $R_\textrm{c}$
714 values using their respective damping parameters are 76$\pm$1 and
715 75$\pm$2. These values are lower than the values reported for TIP5P-E
716 with the Ewald sum; however, they are more in line with the values
717 reported by Mahoney {\it et al.} for TIP5P while using a reaction
718 field (RF) with an infinite RF dielectric constant
719 (81.5$\pm$1.6).\cite{Mahoney00}
720
721 We can use the same trend of $\alpha$ with $R_\textrm{c}$ for SSD/RF
722 and for a 12~\AA\ $R_\textrm{c}$, the resulting dielectric constant is
723 82.6$\pm$0.6. This value compares very favorably with the experimental
724 value of 78.3.\cite{Malmberg56} This is not surprising given that
725 early studies of the SSD model indicate a static dielectric constant
726 around 81.\cite{Liu96} The static dielectric constants for SSD/RF
727 simulations with 9 and 12\AA\ $R_\textrm{c}$ values using their
728 respective damping parameters are 82.6$\pm$0.6 and 75$\pm$2.
729
730 As a final note, aside from a slight
731 lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
732 not alter the other other thermodynamic properties.
733
734 \subsubsection{Dynamic Properties}\label{sec:t5peDynamics}
735
736 To look at the dynamic properties of TIP5P-E when using the SF
737 method, 200~ps $NVE$ simulations were performed for each temperature
738 at the average density reported by the $NPT$ simulations using 9 and
739 12~\AA\ $R_\textrm{c}$ values using the ideal $\alpha$ values
740 determined above (0.2875 and 0.2125~\AA$^{-1}$). The self-diffusion
741 constants ($D$) were calculated using the mean square displacement
742 (MSD) form of the Einstein relation,
743 \begin{equation}
744 D = \lim_{t\rightarrow\infty}
745 \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
746 \label{eq:MSD}
747 \end{equation}
748 where $t$ is time, and $\mathbf{r}_i$ is the position of particle
749 $i$.\cite{Allen87}
750
751 \begin{figure}
752 \centering
753 \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
754 \caption{Body-fixed coordinate frame for a water molecule. The
755 respective molecular principle axes point in the direction of the
756 labeled frame axes.}
757 \label{fig:waterFrame}
758 \end{figure}
759 In addition to translational diffusion, orientational relaxation times
760 were calculated for comparisons with the Ewald simulations and with
761 experiments. These values were determined from the same 200~ps $NVE$
762 trajectories used for translational diffusion by calculating the
763 orientational time correlation function,
764 \begin{equation}
765 C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
766 \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
767 \label{eq:OrientCorr}
768 \end{equation}
769 where $P_l$ is the Legendre polynomial of order $l$ and
770 $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
771 principle axis $\alpha$. The principle axis frame for these water
772 molecules is shown in figure \ref{fig:waterFrame}. As an example,
773 $C_l^y$ is calculated from the time evolution of the unit vector
774 connecting the two hydrogen atoms.
775
776 \begin{figure}
777 \centering
778 \includegraphics[width=3.5in]{./figures/exampleOrientCorr.pdf}
779 \caption{Example plots of the orientational autocorrelation functions
780 for the first and second Legendre polynomials. These curves show the
781 time decay of the unit vector along the $y$ principle axis.}
782 \label{fig:OrientCorr}
783 \end{figure}
784 From the orientation autocorrelation functions, we can obtain time
785 constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
786 some example plots of orientational autocorrelation functions for the
787 first and second Legendre polynomials. 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=3.5in]{./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 at high temperatures are again the product of the lower
813 densities in comparison with experiment and do not provide any special
814 insight into differences between the electrostatic summation
815 techniques. With the undamped SF technique, TIP5P-E tends to
816 diffuse a little faster than with the Ewald sum; however, use of light
817 to moderate damping results in indistinguishable $D$ values. Though
818 not apparent in this figure, SF values at the lowest temperature
819 are approximately twice as slow as $D$ values obtained using the Ewald
820 sum. These values support the observation from section
821 \ref{sec:t5peThermo} that there appeared to be a change to a more
822 glassy-like phase with the SF technique at these lower
823 temperatures, though this change seems to be more prominent with the
824 {\it undamped} SF method, which has stronger local pairwise
825 electrostatic interactions.
826
827 The $\tau_2^y$ results in the lower frame of figure
828 \ref{fig:t5peDynamics} show a much greater difference between the {\sc
829 sf} results and the Ewald results. At all temperatures shown, TIP5P-E
830 relaxes faster than experiment with the Ewald sum while tracking
831 experiment fairly well when using the SF technique, independent
832 of the choice of damping constant. Their are several possible reasons
833 for this deviation between techniques. The Ewald results were
834 calculated using shorter (10ps) trajectories than the SF results
835 (25ps). A quick calculation from a 10~ps trajectory with SF with
836 an $\alpha$ of 0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in
837 $\tau_2^y$, placing the result more in line with that obtained using
838 the Ewald sum. This example supports this explanation; however,
839 recomputing the results to meet a poorer statistical standard is
840 counter-productive. Assuming the Ewald results are not entirely the
841 product of poor statistics, differences in techniques to integrate the
842 orientational motion could also play a role. {\sc shake} is the most
843 commonly used technique for approximating rigid-body orientational
844 motion,\cite{Ryckaert77} whereas in {\sc oopse}, we maintain and
845 integrate the entire rotation matrix using the {\sc dlm}
846 method.\cite{Meineke05} Since {\sc shake} is an iterative constraint
847 technique, if the convergence tolerances are raised for increased
848 performance, error will accumulate in the orientational
849 motion. Finally, the Ewald results were calculated using the $NVT$
850 ensemble, while the $NVE$ ensemble was used for SF
851 calculations. The additional mode of motion due to the thermostat will
852 alter the dynamics, resulting in differences between $NVT$ and $NVE$
853 results. These differences are increasingly noticeable as the
854 thermostat time constant decreases.
855
856
857 \subsection{SSD/RF}
858
859 In section \ref{sec:dampingMultipoles}, we described a method for
860 applying the damped SF technique to systems containing point
861 multipoles. The soft sticky dipole (SSD) family of water models is the
862 perfect test case because of the dipole-dipole (and
863 charge-dipole/quadrupole) interactions that are present. As alluded to
864 in the name, soft sticky dipole water models are single point
865 molecules that consist of a ``soft'' Lennard-Jones sphere, a
866 point-dipole, and a tetrahedral function for capturing the hydrogen
867 bonding nature of water - a spherical harmonic term for water-water
868 tetrahedral interactions and a point-quadrupole for interactions with
869 surrounding charges. Detailed descriptions of these models can be
870 found in other studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
871
872 In deciding which version of the SSD model to use, we need only
873 consider that the SF technique was presented as a pairwise
874 replacement for the Ewald summation. It has been suggested that models
875 parametrized for the Ewald summation (like TIP5P-E) would be
876 appropriate for use with a reaction field and vice versa.\cite{Rick04}
877 Therefore, we decided to test the SSD/RF water model, which was
878 parametrized for use with a reaction field, with this damped
879 electrostatic technique to see how the calculated properties change.
880
881 \subsubsection{Dipolar Damping}
882
883 \begin{table}
884 \caption{Properties of SSD/RF when using different electrostatic
885 correction methods.}
886 \footnotesize
887 \centering
888 \begin{tabular}{ llccc }
889 \toprule
890 \toprule
891 & & Reaction Field [Ref. \citen{Fennell04}] & Damped Electrostatics & Experiment [Ref.] \\
892 & & $\epsilon = 80$ & $R_\textrm{c} = 12$\AA ; $\alpha = 0.2125$~\AA$^{-1}$ & \\
893 \midrule
894 $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 [\citen{CRC80}]\\
895 $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 [\citen{Wagner02}] \\
896 $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 [\citen{Mills73}]\\
897 $n_C$ & & 4.4 & 4.2 & 4.7 [\citen{Hura00}]\\
898 $n_H$ & & 3.7 & 3.7 & 3.5 [\citen{Soper86}]\\
899 $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 [\citen{Eisenberg69}]\\
900 $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 [\citen{Krynicki66}]\\
901 \bottomrule
902 \end{tabular}
903 \label{tab:dampedSSDRF}
904 \end{table}
905 The properties shown in table \ref{tab:dampedSSDRF} compare
906 quite well. The average density shows a modest increase when
907 using damped electrostatics in place of the reaction field. This comes
908 about because we neglect the pressure effect due to the surroundings
909 outside of the cutoff, instead relying on screening effects to
910 neutralize electrostatic interactions at long distances. The $C_p$
911 also shows a slight increase, indicating greater fluctuation in the
912 enthalpy at constant pressure. The only other differences between the
913 damped and reaction field results are the dipole reorientational time
914 constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
915 the water molecules relax more quickly and exhibit behavior very
916 similar to the experimental values. These results indicate that not
917 only is it reasonable to use damped electrostatics with SSD/RF, it is
918 recommended if capturing realistic dynamics is of primary
919 importance. This is an encouraging result because the damping methods
920 are more generally applicable than reaction field. Using damping,
921 SSD/RF can be used effectively with mixed systems, such as dissolved
922 ions, dissolved organic molecules, or even proteins.
923
924 \section{Application of Pairwise Electrostatic Corrections: Imaginary Ice}
925
926 In an earlier work, we performed a series of free energy calculations
927 on several ice polymorphs which are stable or metastable at low
928 pressures, one of which (Ice-$i$) we observed in spontaneous
929 crystallizations of an SSD type single point water
930 model.\cite{Fennell05} In this study, a distinct dependence of the
931 free energies on the interaction cutoff and correction technique was
932 observed. Being that the SF technique can be used as a simple
933 and efficient replacement for the Ewald summation, it would be
934 interesting to apply it in addressing the question of the free
935 energies of these ice polymorphs with varying water models. To this
936 end, we set up thermodynamic integrations of all of the previously
937 discussed ice polymorphs using the SF technique with a cutoff
938 radius of 12~\AA\ and an $\alpha$ of 0.2125~\AA . These calculations
939 were performed on TIP5P-E and TIP4P-Ew (variants of the root models
940 optimized for the Ewald summation) as well as SPC/E, and SSD/RF.
941
942 \begin{table}
943 \centering
944 \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
945 using the damped SF electrostatic correction method with a
946 variety of water models.}
947 \begin{tabular}{ lccccc }
948 \toprule
949 \toprule
950 Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
951 \cmidrule(lr){2-6}
952 & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
953 \midrule
954 TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
955 TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
956 SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
957 SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
958 \bottomrule
959 \end{tabular}
960 \label{tab:dampedFreeEnergy}
961 \end{table}
962 The results of these calculations in table \ref{tab:dampedFreeEnergy}
963 show similar behavior to the Ewald results in the previous study, at
964 least for SSD/RF and SPC/E which are present in both.\cite{Fennell05}
965 The Helmholtz free energies of the ice polymorphs with SSD/RF order in
966 the same fashion, with Ice-$i$ having the lowest free energy; however,
967 the Ice-$i$ and ice B free energies are quite a bit closer (nearly
968 isoenergetic). The SPC/E results show the near isoenergetic behavior
969 when using the Ewald summation.\cite{Fennell05} Ice B has the lowest
970 Helmholtz free energy; however, all the polymorph results overlap
971 within error.
972
973 The most interesting results from these calculations come from the
974 more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
975 optimized for use with an electrostatic correction and are
976 geometrically arranged to mimic water following two different
977 ideas. In TIP5P-E, the primary location for the negative charge in the
978 molecule is assigned to the lone-pairs of the oxygen, while TIP4P-Ew
979 places the negative charge near the center-of-mass along the H-O-H
980 bisector. There is some debate as to which is the proper choice for
981 the negative charge location, and this has in part led to a six-site
982 water model that balances both of these options.\cite{Vega05,Nada03}
983 The limited results in table \ref{tab:dampedFreeEnergy} support the
984 results of Vega {\it et al.}, which indicate the TIP4P charge location
985 geometry is more physically valid.\cite{Vega05} With the TIP4P-Ew
986 water model, the experimentally observed polymorph (ice
987 I$_\textrm{h}$) is the preferred form with ice I$_\textrm{c}$ slightly
988 higher in energy, though overlapping within error. Additionally, the
989 less realistic ice B and Ice-$i^\prime$ structures are destabilized
990 relative to these polymorphs. TIP5P-E shows similar behavior to SPC/E,
991 where there is no real free energy distinction between the various
992 polymorphs, because many overlap within error. While ice B is close in
993 free energy to the other polymorphs, these results fail to support the
994 findings of other researchers indicating the preferred form of TIP5P
995 at 1~atm is a structure similar to ice
996 B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we are
997 looking at TIP5P-E rather than TIP5P, and the differences in the
998 Lennard-Jones parameters could be a reason for this dissimilarity.
999 Overall, these results indicate that TIP4P-Ew is a better mimic of
1000 real water than these other models when studying crystallization and
1001 solid forms of water.
1002
1003 \section{Conclusions}
1004
1005 \section{Acknowledgments}
1006 Support for this project was provided by the National Science
1007 Foundation under grant CHE-0134881. Computation time was provided by
1008 the Notre Dame Center for Research Computing.
1009
1010 \newpage
1011
1012 \bibliographystyle{achemso}
1013 \bibliography{multipoleSFPaper}
1014
1015
1016 \end{document}