ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipoleSFPaper/multipoleSFPaper.tex
Revision: 3067
Committed: Tue Oct 24 19:46:30 2006 UTC (17 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 56149 byte(s)
Log Message:
more changes

File Contents

# User Rev Content
1 chrisfen 3064 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 3066 \documentclass[11pt]{article}
3 chrisfen 3064 \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 gezelter 3066 \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 chrisfen 3064
31     \begin{document}
32    
33 gezelter 3066 \title{Pairwise Alternatives to the Ewald Sum: Applications
34     and Extension to Point Multipoles}
35 chrisfen 3064
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 gezelter 3066 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 chrisfen 3064 \end{abstract}
57    
58     %\narrowtext
59    
60     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61     % BODY OF TEXT
62     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63    
64     \section{Introduction}
65    
66 gezelter 3066 Over the past several years, there has been increasing interest in
67     pairwise methods for correcting electrostatic interactions in computer
68 chrisfen 3064 simulations of condensed molecular
69 gezelter 3066 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 chrisfen 3064
76 gezelter 3066 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 gezelter 3067 force (SF) techniques. The potential for the damped form of the SP
83     method is given by
84 chrisfen 3064 \begin{equation}
85 gezelter 3067 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
86     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
87     \quad r\leqslant R_\textrm{c},
88     \label{eq:DSPPot}
89     \end{equation}
90     while the damped form of the SF method is given by
91     \begin{equation}
92 chrisfen 3064 \begin{split}
93     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
94     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
95     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
96     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
97     + \frac{2\alpha}{\pi^{1/2}}
98     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
99     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
100     \quad r\leqslant R_\textrm{c}.
101     \label{eq:DSFPot}
102     \end{split}
103     \end{equation}
104 gezelter 3067 (In these potentials and in all electrostatic quantities that follow,
105 gezelter 3066 the standard $4 \pi \epsilon_{0}$ has been omitted for clarity.)
106 chrisfen 3064
107 gezelter 3066 The damped SF method is an improvement over the SP method because
108     there is no discontinuity in the forces as particles move out of the
109     cutoff radius ($R_\textrm{c}$). This is accomplished by shifting the
110     forces (and potential) to zero at $R_\textrm{c}$. This is analogous to
111     neutralizing the charge as well as the force effect of the charges
112     within $R_\textrm{c}$.
113    
114     To complete the charge neutralization procedure, a self-neutralization
115 gezelter 3067 term is included in the potential. This term is constant (as long as
116     the charges and cutoff radius do not change), and exists outside the
117     normal pair-loop. It can be thought of as a contribution from a
118     charge opposite in sign, but equal in magnitude, to the central
119 gezelter 3066 charge, but which has been spread out over the surface of the cutoff
120     sphere. This term is calculated via a single loop over all charges in
121     the system. For the undamped case, the self term can be written as
122 chrisfen 3064 \begin{equation}
123 gezelter 3066 V_\textrm{self} = \frac{1}{2 R_\textrm{c}} \sum_{i=1}^N q_i^{2},
124 chrisfen 3064 \label{eq:selfTerm}
125     \end{equation}
126     while for the damped case it can be written as
127     \begin{equation}
128 gezelter 3066 V_\textrm{self} = \left(\frac{\alpha}{\sqrt{\pi}}
129     + \frac{\textrm{erfc}(\alpha
130     R_\textrm{c})}{2R_\textrm{c}}\right) \sum_{i=1}^N q_i^{2}.
131 chrisfen 3064 \label{eq:dampSelfTerm}
132     \end{equation}
133     The first term within the parentheses of equation
134 gezelter 3066 (\ref{eq:dampSelfTerm}) is identical to the self term in the Ewald
135 chrisfen 3064 summation, and comes from the utilization of the complimentary error
136     function for electrostatic damping.\cite{deLeeuw80,Wolf99}
137    
138 gezelter 3066 The SF, SP, and Wolf methods operate by neutralizing the total charge
139     contained within the cutoff sphere surrounding each particle. This is
140     accomplished by creating image charges on the surface of the cutoff
141     sphere for each pair interaction computed within the sphere. The
142     damping function applied to the potential is also an important method
143     for accelerating convergence. In the case of systems involving
144     electrostatic distributions of higher order than point charges, the
145     question remains: How will the shifting and damping need to be
146     modified in order to accommodate point multipoles?
147 chrisfen 3064
148 gezelter 3066 \section{Electrostatic Damping for Point
149     Multipoles}\label{sec:dampingMultipoles}
150 chrisfen 3064
151 gezelter 3066 To apply the SF method for systems involving point multipoles, we
152     consider separately the two techniques (shifting and damping) which
153     contribute to the effectiveness of the DSF potential.
154 chrisfen 3064
155 gezelter 3066 As noted above, shifting the potential and forces is employed to
156     neutralize the total charge contained within each cutoff sphere;
157     however, in a system composed purely of point multipoles, each cutoff
158     sphere is already neutral, so shifting becomes unnecessary.
159    
160     In a mixed system of charges and multipoles, the undamped SF potential
161     needs only to shift the force terms between charges and smoothly
162     truncate the multipolar interactions with a switching function. The
163     switching function is required for energy conservation, because a
164     discontinuity will exist in both the potential and forces at
165     $R_\textrm{c}$ in the absence of shifting terms.
166    
167     To damp the SF potential for point multipoles, we need to incorporate
168     the complimentary error function term into the standard forms of the
169     multipolar potentials. We can determine the necessary damping
170     functions by replacing $1/r$ with $\mathrm{erfc}(\alpha r)/r$ in the
171     multipole expansion. This procedure quickly becomes quite complex
172     with ``two-center'' systems, like the dipole-dipole potential, and is
173     typically approached using spherical harmonics.\cite{Hirschfelder67} A
174     simpler method for determining damped multipolar interaction
175     potentials arises when we adopt the tensor formalism described by
176     Stone.\cite{Stone02}
177    
178     The tensor formalism for electrostatic interactions involves obtaining
179     the multipole interactions from successive gradients of the monopole
180     potential. Thus, tensors of rank zero through two are
181 chrisfen 3064 \begin{equation}
182 gezelter 3066 T = \frac{1}{r_{ij}},
183     \label{eq:tensorRank1}
184 chrisfen 3064 \end{equation}
185     \begin{equation}
186 gezelter 3066 T_\alpha = \nabla_\alpha \frac{1}{r_{ij}},
187     \label{eq:tensorRank2}
188 chrisfen 3064 \end{equation}
189 gezelter 3066 \begin{equation}
190     T_{\alpha\beta} = \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
191     \label{eq:tensorRank3}
192     \end{equation}
193     where the form of the first tensor is the charge-charge potential, the
194     second gives the charge-dipole potential, and the third gives the
195     charge-quadrupole and dipole-dipole potentials.\cite{Stone02} Since
196     the force is $-\nabla V$, the forces for each potential come from the
197     next higher tensor.
198 chrisfen 3064
199 gezelter 3066 As one would do with the multipolar expansion, we can replace $r^{-1}$
200 gezelter 3067 with $\mathrm{erfc}(\alpha r)/r$ to obtain damped forms of the
201 gezelter 3066 electrostatic potentials. Equation \ref{eq:tensorRank2} generates a
202     damped charge-dipole potential,
203 chrisfen 3064 \begin{equation}
204     V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
205     \label{eq:dChargeDipole}
206     \end{equation}
207     where $c_1(r_{ij})$ is
208     \begin{equation}
209     c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
210     + \textrm{erfc}(\alpha r_{ij}).
211     \label{eq:c1Func}
212     \end{equation}
213    
214 gezelter 3066 Equation \ref{eq:tensorRank3} generates a damped dipole-dipole potential,
215 chrisfen 3064 \begin{equation}
216 gezelter 3066 V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
217     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
218     c_2(r_{ij}) -
219     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
220     c_1(r_{ij}),
221     \label{eq:dampDipoleDipole}
222 chrisfen 3064 \end{equation}
223 gezelter 3066 where
224 chrisfen 3064 \begin{equation}
225     c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
226     + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
227     + \textrm{erfc}(\alpha r_{ij}).
228     \end{equation}
229     Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
230     term. Continuing with higher rank tensors, we can obtain the damping
231     functions for higher multipole potentials and forces. Each subsequent
232     damping function includes one additional term, and we can simplify the
233     procedure for obtaining these terms by writing out the following
234     generating function,
235     \begin{equation}
236     c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
237     {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
238     \label{eq:dampingGeneratingFunc}
239     \end{equation}
240     where,
241     \begin{equation}
242     m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
243     m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
244     m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
245     1 & m = -1\textrm{ or }0,
246     \end{array}\right.
247     \label{eq:doubleFactorial}
248     \end{equation}
249     and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
250 gezelter 3066 is similar in form to those obtained by Smith and Aguado and Madden
251     for the application of the Ewald sum to
252 chrisfen 3064 multipoles.\cite{Smith82,Smith98,Aguado03}
253    
254     Returning to the dipole-dipole example, the potential consists of a
255     portion dependent upon $r^{-5}$ and another dependent upon
256 gezelter 3066 $r^{-3}$.
257 chrisfen 3064 $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
258 gezelter 3066 respectively. The forces for the damped dipole-dipole interaction, are
259     obtained from the next higher tensor, $T_{\alpha \beta \gamma}$,
260 chrisfen 3064 \begin{equation}
261     \begin{split}
262     F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
263     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
264     c_3(r_{ij})\\ &-
265     3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
266     \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
267     \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
268     {r^5_{ij}} c_2(r_{ij}),
269     \end{split}
270     \label{eq:dampDipoleDipoleForces}
271     \end{equation}
272 gezelter 3066 Using the tensor formalism, we can dampen higher order multipolar
273     interactions using the same effective damping function that we use for
274     charge-charge interactions. This allows us to include multipoles in
275     simulations involving damped electrostatic interactions.
276 chrisfen 3064
277 gezelter 3067 As a practical consideration, we note that the evaluation of the error
278     function inside a pair loop can become quite costly. In practice, we
279     pre-compute the $c_n(r)$ functions over a grid of $r$ values and use
280     cubic spline interpolation to obtain estimates of these functions when
281     necessary. Using this procedure, the computational cost of damped
282     electrostatics is essentially the same as the undamped case.
283 chrisfen 3064
284 gezelter 3067 \section{Applications of Damped Shifted-Froce Electrostatics}
285    
286 gezelter 3066 Our earlier work on the SF method showed that it can give nearly
287     quantitive agreement with SPME-derived configurational energy
288     differences. The force and torque vectors in identical configurations
289     are also nearly equivalent under the damped SF potential and
290     SPME.\cite{Fennell06} Although these measures bode well for the
291     performance of the SF method in both Monte Carlo and Molecular
292     Dynamics simulations, it would be helpful to have direct comparisons
293     of structural, thermodynamic, and dynamic quantities. To address
294     this, we performed a detailed analysis of a group of simulations
295     involving water models (both point charge and multipolar) under a
296     number of different simulation conditions.
297 chrisfen 3064
298 gezelter 3066 To provide the most difficult test for the damped SF method, we have
299     chosen a model that has been optimized for use with Ewald sum, and
300     have compared the simulated properties to those computed via Ewald.
301     It is well known that water models parametrized for use with the Ewald
302     sum give calculated properties that are in better agreement with
303     experiment.\cite{vanderSpoel98,Horn04,Rick04} For these reasons, we
304     chose the TIP5P-E water model for our comparisons involving point
305     charges.\cite{Rick04}
306 chrisfen 3064
307 gezelter 3066 The soft sticky dipole (SSD) family of water models is the perfect
308     test case for the point-multipolar extension of damped electrostatics.
309     SSD water models are single point molecules that consist of a ``soft''
310     Lennard-Jones sphere, a point-dipole, and a tetrahedral function for
311     capturing the hydrogen bonding nature of water - a spherical harmonic
312     term for water-water tetrahedral interactions and a point-quadrupole
313     for interactions with surrounding charges. Detailed descriptions of
314     these models can be found in other
315     studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
316    
317     In deciding which version of the SSD model to use, we need only
318     consider that the SF technique was presented as a pairwise replacement
319     for the Ewald summation. It has been suggested that models
320     parametrized for the Ewald summation (like TIP5P-E) would be
321     appropriate for use with a reaction field and vice versa.\cite{Rick04}
322     Therefore, we decided to test the SSD/RF water model, which was
323     parametrized for use with a reaction field, with the damped
324     electrostatic technique to see how the calculated properties change.
325    
326 chrisfen 3064 The TIP5P-E water model is a variant of Mahoney and Jorgensen's
327     five-point transferable intermolecular potential (TIP5P) model for
328     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
329 gezelter 3066 maximum in liquid water near 4$^\circ$C. As with many previous point
330     charge water models (such as ST2, TIP3P, TIP4P, SPC, and SPC/E), TIP5P
331     was parametrized using a simple cutoff with no long-range
332     electrostatic
333 chrisfen 3064 correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
334     Without this correction, the pressure term on the central particle
335     from the surroundings is missing. When this correction is included,
336     systems of these particles expand to compensate for this added
337     pressure term and under-predict the density of water under standard
338 gezelter 3066 conditions. In developing TIP5P-E, Rick preserved the geometry and
339     point charge magnitudes in TIP5P and focused on altering the
340     Lennard-Jones parameters to correct the density at 298~K. With the
341 chrisfen 3064 density corrected, he compared common water properties for TIP5P-E
342     using the Ewald sum with TIP5P using a 9~\AA\ cutoff.
343    
344 gezelter 3066 In the following sections, we compare these same properties calculated
345     from TIP5P-E using the Ewald sum with TIP5P-E using the damped SF
346     technique. Our comparisons include the SF technique with a 12~\AA\
347     cutoff and an $\alpha$ = 0.0, 0.1, and 0.2~\AA$^{-1}$, as well as a
348     9~\AA\ cutoff with an $\alpha$ = 0.2~\AA$^{-1}$.
349 chrisfen 3064
350 gezelter 3066 \subsection{The Density Maximum of TIP5P-E}\label{sec:t5peDensity}
351 chrisfen 3064
352     To compare densities, $NPT$ simulations were performed with the same
353     temperatures as those selected by Rick in his Ewald summation
354     simulations.\cite{Rick04} In order to improve statistics around the
355     density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
356     25$^\circ$C, while 2~ns trajectories were obtained at all other
357     temperatures. The average densities were calculated from the later
358     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
359     method for accumulating statistics, these sequences were spliced into
360     200 segments, each providing an average density. These 200 density
361     values were used to calculate the average and standard deviation of
362     the density at each temperature.\cite{Mahoney00}
363    
364     \begin{figure}
365     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
366     \caption{Density versus temperature for the TIP5P-E water model when
367 gezelter 3066 using the Ewald summation (Ref. \citen{Rick04}) and the SF method with
368     varying cutoff radii and damping coefficients. The pressure term from
369     the image-charge shell is larger than that provided by the
370     reciprocal-space portion of the Ewald summation, leading to slightly
371     lower densities. This effect is more visible with the 9~\AA\ cutoff,
372     where the image charges exert a greater force on the central
373     particle. The error bars for the SF methods show the average one-sigma
374     uncertainty of the density measurement, and this uncertainty is the
375     same for all the SF curves.}
376 chrisfen 3064 \label{fig:t5peDensities}
377     \end{figure}
378     Figure \ref{fig:t5peDensities} shows the densities calculated for
379 gezelter 3066 TIP5P-E using differing electrostatic corrections overlaid with the
380     experimental values.\cite{CRC80} The densities when using the SF
381     technique are close to, but typically lower than, those calculated
382 chrisfen 3064 using the Ewald summation. These slightly reduced densities indicate
383     that the pressure component from the image charges at R$_\textrm{c}$
384     is larger than that exerted by the reciprocal-space portion of the
385 gezelter 3066 Ewald summation. Bringing the image charges closer to the central
386 chrisfen 3064 particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
387     preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
388     charge interactions on the central particle and results in a further
389     reduction of the densities.
390    
391     Because the strength of the image charge interactions has a noticeable
392     effect on the density, we would expect the use of electrostatic
393     damping to also play a role in these calculations. Larger values of
394     $\alpha$ weaken the pair-interactions; and since electrostatic damping
395     is distance-dependent, force components from the image charges will be
396     reduced more than those from particles close the the central
397     charge. This effect is visible in figure \ref{fig:t5peDensities} with
398 gezelter 3066 the damped SF sums showing slightly higher densities; however, it is
399     clear that the choice of cutoff radius plays a much more important
400     role in the resulting densities.
401 chrisfen 3064
402 gezelter 3066 All of the above density calculations were performed with systems of
403     512 water molecules. Rick observed a system size dependence of the
404     computed densities when using the Ewald summation, most likely due to
405     his tying of the convergence parameter to the box
406 chrisfen 3064 dimensions.\cite{Rick04} For systems of 256 water molecules, the
407     calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
408     system size of 256 molecules would force the use of a shorter
409 gezelter 3066 R$_\textrm{c}$ when using the SF technique, and this would also lower
410     the densities. Moving to larger systems, as long as the R$_\textrm{c}$
411     remains at a fixed value, we would expect the densities to remain
412     constant.
413 chrisfen 3064
414 gezelter 3066 \subsection{Liquid Structure of TIP5P-E}\label{sec:t5peLiqStructure}
415 chrisfen 3064
416     The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
417     been compared in great detail with the various common water models,
418     and TIP5P was found to be in better agreement than other rigid,
419     non-polarizable models.\cite{Sorenson00} This excellent agreement with
420     experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
421 gezelter 3066 check whether the choice of using the Ewald summation or the SF
422 chrisfen 3064 technique alters the liquid structure, the $g_\textrm{OO}(r)$s at
423     298~K and 1~atm were determined for the systems compared in the
424     previous section.
425    
426     \begin{figure}
427     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
428     \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and
429     1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
430     sf} technique with varying parameters. Even with the reduced densities
431 gezelter 3066 using the SF technique, the $g_\textrm{OO}(r)$s are essentially
432 chrisfen 3064 identical.}
433     \label{fig:t5peGofRs}
434     \end{figure}
435     The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
436     sf} technique with a various parameters are overlaid on the
437     $g_\textrm{OO}(r)$ while using the Ewald summation in
438     figure~\ref{fig:t5peGofRs}. The differences in density do not appear
439     to have any effect on the liquid structure as the $g_\textrm{OO}(r)$s
440 gezelter 3066 are indistinguishable. These results do indicate that
441 chrisfen 3064 $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
442     correction.
443    
444 gezelter 3066 \subsection{Thermodynamic Properties of TIP5P-E}\label{sec:t5peThermo}
445 chrisfen 3064
446 gezelter 3066 In addition to the density and structual features of the liquid, there
447     are a variety of thermodynamic quantities that can be calculated for
448     water and compared directly to experimental values. Some of these
449     additional quantities include the latent heat of vaporization ($\Delta
450     H_\textrm{vap}$), the constant pressure heat capacity ($C_p$), the
451     isothermal compressibility ($\kappa_T$), the thermal expansivity
452     ($\alpha_p$), and the static dielectric constant ($\epsilon$). All of
453     these properties were calculated for TIP5P-E with the Ewald summation,
454     so they provide a good set for comparisons involving the SF technique.
455 chrisfen 3064
456     The $\Delta H_\textrm{vap}$ is the enthalpy change required to
457     transform one mole of substance from the liquid phase to the gas
458     phase.\cite{Berry00} In molecular simulations, this quantity can be
459     determined via
460     \begin{equation}
461     \begin{split}
462 gezelter 3066 \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq} \\
463     &= E_\textrm{gas} - E_\textrm{liq}
464     + P(V_\textrm{gas} - V_\textrm{liq}) \\
465     &\approx -\frac{\langle U_\textrm{liq}\rangle}{N}+ RT,
466 chrisfen 3064 \end{split}
467     \label{eq:DeltaHVap}
468     \end{equation}
469 gezelter 3066 where $E$ is the total energy, $U$ is the potential energy, $P$ is the
470 chrisfen 3064 pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
471     the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
472     seen in the last line of equation (\ref{eq:DeltaHVap}), we can
473     approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
474     state. This allows us to cancel the kinetic energy terms, leaving only
475 gezelter 3066 the $U_\textrm{liq}$ term. Additionally, the $pV$ term for the gas is
476 chrisfen 3064 several orders of magnitude larger than that of the liquid, so we can
477     neglect the liquid $pV$ term.
478    
479     The remaining thermodynamic properties can all be calculated from
480     fluctuations of the enthalpy, volume, and system dipole
481     moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
482     enthalpy in constant pressure simulations via
483     \begin{equation}
484     \begin{split}
485 gezelter 3066 C_p = \left(\frac{\partial H}{\partial T}\right)_{N,P}
486 chrisfen 3064 = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
487     \end{split}
488     \label{eq:Cp}
489     \end{equation}
490     where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
491     \begin{equation}
492     \begin{split}
493     \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
494 gezelter 3066 = \frac{(\langle V^2\rangle_{NPT} - \langle V\rangle^2_{NPT})}
495     {k_BT\langle V\rangle_{NPT}},
496 chrisfen 3064 \end{split}
497     \label{eq:kappa}
498     \end{equation}
499     and $\alpha_p$ can be calculated via
500     \begin{equation}
501     \begin{split}
502     \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
503 gezelter 3066 = \frac{(\langle VH\rangle_{NPT}
504     - \langle V\rangle_{NPT}\langle H\rangle_{NPT})}
505     {k_BT^2\langle V\rangle_{NPT}}.
506 chrisfen 3064 \end{split}
507     \label{eq:alpha}
508     \end{equation}
509     Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
510     be calculated for systems of non-polarizable substances via
511     \begin{equation}
512     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
513     \label{eq:staticDielectric}
514     \end{equation}
515     where $\epsilon_0$ is the permittivity of free space and $\langle
516     M^2\rangle$ is the fluctuation of the system dipole
517     moment.\cite{Allen87} The numerator in the fractional term in equation
518     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
519     dipole moment, identical to the quantity calculated in the
520     finite-system Kirkwood $g$ factor ($G_k$):
521     \begin{equation}
522     G_k = \frac{\langle M^2\rangle}{N\mu^2},
523     \label{eq:KirkwoodFactor}
524     \end{equation}
525     where $\mu$ is the dipole moment of a single molecule of the
526     homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
527     fluctuation term in both equation (\ref{eq:staticDielectric}) and
528     \ref{eq:KirkwoodFactor} is calculated as follows,
529     \begin{equation}
530     \begin{split}
531     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
532     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
533     &= \langle M_x^2+M_y^2+M_z^2\rangle
534     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
535     + \langle M_x\rangle^2).
536     \end{split}
537     \label{eq:fluctBoxDipole}
538     \end{equation}
539     This fluctuation term can be accumulated during the simulation;
540     however, it converges rather slowly, thus requiring multi-nanosecond
541     simulation times.\cite{Horn04} In the case of tin-foil boundary
542     conditions, the dielectric/surface term of the Ewald summation is
543 gezelter 3066 equal to zero. Since the SF method also lacks this
544 chrisfen 3064 dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
545     valid for determining static dielectric constants.
546    
547     All of the above properties were calculated from the same trajectories
548     used to determine the densities in section \ref{sec:t5peDensity}
549     except for the static dielectric constants. The $\epsilon$ values were
550     accumulated from 2~ns $NVE$ ensemble trajectories with system densities
551     fixed at the average values from the $NPT$ simulations at each of the
552     temperatures. The resulting values are displayed in figure
553     \ref{fig:t5peThermo}.
554     \begin{figure}
555     \centering
556 gezelter 3067 \includegraphics[width=5.8in]{./figures/t5peThermo.pdf}
557 chrisfen 3064 \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
558 gezelter 3066 and the SF techniques along with the experimental values. Units
559 chrisfen 3064 for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
560     cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
561     and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
562     reference \cite{Rick04}. Experimental values for $\Delta
563     H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
564     \cite{Kell75}. Experimental values for $C_p$ are from reference
565     \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
566     \cite{Malmberg56}.}
567     \label{fig:t5peThermo}
568     \end{figure}
569    
570 gezelter 3066 For all of the properties computed, the trends with temperature
571     obtained when using the Ewald summation are reproduced with the SF
572     technique. One noticeable difference between the properties calculated
573     using the two methods are the lower $\Delta H_\textrm{vap}$ values
574     when using SF. This is to be expected due to the direct weakening of
575     the electrostatic interaction through forced neutralization. This
576     results in an increase of the intermolecular potential producing lower
577     values from equation (\ref{eq:DeltaHVap}). The slopes of these values
578     with temperature are similar to that seen using the Ewald summation;
579     however, they are both steeper than the experimental trend, indirectly
580     resulting in the inflated $C_p$ values at all temperatures.
581 chrisfen 3064
582     Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
583     all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
584 gezelter 3066 and $C_p$ results, the deviations between experiment and simulation in
585     this region are not the fault of the electrostatic summation methods
586     but are due to the geometry and parameters of the TIP5P class of water
587     models. Like most rigid, non-polarizable, point-charge water models,
588     the density decreases with temperature at a much faster rate than
589     experiment (see figure \ref{fig:t5peDensities}). This reduced density
590     leads to the inflated compressibility and expansivity values at higher
591     temperatures seen here in figure \ref{fig:t5peThermo}. Incorporation
592     of polarizability and many-body effects are required in order for
593     water models to overcome differences between simulation-based and
594     experimentally determined densities at these higher
595 chrisfen 3064 temperatures.\cite{Laasonen93,Donchev06}
596    
597     At temperatures below the freezing point for experimental water, the
598 gezelter 3066 differences between SF and the Ewald summation results are more
599 chrisfen 3064 apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
600     indicate a more pronounced transition in the supercooled regime,
601 gezelter 3066 particularly in the case of SF without damping.
602 chrisfen 3064
603 gezelter 3066 This points to the onset of a more frustrated or glassy behavior for
604     the undamped and weakly-damped SF simulations of TIP5P-E at
605     temperatures below 250~K than is seen from the Ewald sum. Undamped SF
606     electrostatics has a stronger contribution from nearby charges.
607     Damping these local interactions or using a reciprocal-space method
608     makes the water less sensitive to ordering on a short length scale.
609     We can recover nearly quantitative agreement with the Ewald results by
610     increasing the damping constant.
611    
612 chrisfen 3064 The final thermodynamic property displayed in figure
613     \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
614 gezelter 3066 between the Ewald and SF methods (and with experiment). It is known
615     that the dielectric constant is dependent upon and is quite sensitive
616     to the imposed boundary conditions.\cite{Neumann80,Neumann83} This is
617     readily apparent in the converged $\epsilon$ values accumulated for
618     the SF simulations. Lack of a damping function results in dielectric
619 chrisfen 3064 constants significantly smaller than those obtained using the Ewald
620     sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
621     agreement considerably. It should be noted that the choice of the
622 gezelter 3066 ``Ewald coefficient'' ($\kappa$) and real-space cutoff values also
623     have a significant effect on the calculated static dielectric constant
624     when using the Ewald summation. In the simulations of TIP5P-E with the
625     Ewald sum, this screening parameter was tethered to the simulation box
626     size (as was the $R_\textrm{c}$).\cite{Rick04} In general, systems
627     with larger screening parameters reported larger dielectric constant
628     values, the same behavior we see here with {\sc sf}; however, the
629     choice of cutoff radius also plays an important role.
630 chrisfen 3064
631 gezelter 3067 \subsection{Dielectric Constants for TIP5P-E and SSD/RF}\label{sec:t5peDielectric}
632 chrisfen 3064
633 gezelter 3066 In the previous section, we observed that the choice of damping
634     coefficient plays a major role in the calculated dielectric constant
635     for the SF method. Similar damping parameter behavior was observed in
636     the long-time correlated motions of the NaCl crystal.\cite{Fennell06}
637     The static dielectric constant is calculated from the long-time
638     fluctuations of the system's accumulated dipole moment
639     (Eq. (\ref{eq:staticDielectric})), so it is quite sensitive to the
640     choice of damping parameter. Since $\alpha$ is an adjustable
641     parameter, it would be best to choose optimal damping constants such
642     that any arbitrary choice of cutoff radius will properly capture the
643     dielectric behavior of the liquid.
644 chrisfen 3064
645     In order to find these optimal values, we mapped out the static
646     dielectric constant as a function of both the damping parameter and
647 gezelter 3066 cutoff radius for TIP5P-E and for a point-dipolar water model
648     (SSD/RF). To calculate the static dielectric constant, we performed
649     5~ns $NPT$ calculations on systems of 512 water molecules and averaged
650     over the converged region (typically the later 2.5~ns) of equation
651     (\ref{eq:staticDielectric}). The selected cutoff radii ranged from 9,
652     10, 11, and 12~\AA , and the damping parameter values ranged from 0.1
653     to 0.35~\AA$^{-1}$.
654 chrisfen 3064
655     \begin{table}
656     \centering
657 gezelter 3066 \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
658     cutoff radius $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
659 chrisfen 3064 \vspace{6pt}
660 gezelter 3066 \begin{tabular}{ lccccccccc }
661 chrisfen 3064 \toprule
662     \toprule
663 gezelter 3066 & \multicolumn{4}{c}{TIP5P-E} & & \multicolumn{4}{c}{SSD/RF} \\
664     & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} & & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
665     \cmidrule(lr){2-5} \cmidrule(lr){7-10}
666     $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 & & 9 & 10 & 11 & 12 \\
667 chrisfen 3064 \midrule
668 gezelter 3066 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 \\
669     & \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} \\
670     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 \\
671     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 \\
672     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 \\
673     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 \\
674     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 \\
675     & \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} \\
676     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 \\
677     & \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} \\
678     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 \\
679 chrisfen 3064 \bottomrule
680     \end{tabular}
681 gezelter 3066 \label{tab:DielectricMap}
682 chrisfen 3064 \end{table}
683 gezelter 3066
684 chrisfen 3064 The results of these calculations are displayed in table
685 gezelter 3066 \ref{tab:DielectricMap}. The dielectric constants for both models
686     decrease linearly with increasing cutoff radii ($R_\textrm{c}$) and
687     increase linearly with increasing damping ($\alpha$). Another point
688     to note is that choosing $\alpha$ and $R_\textrm{c}$ identical to
689     those used with the Ewald summation results in the same calculated
690     dielectric constant. As an example, in the paper outlining the
691     development of TIP5P-E, the real-space cutoff and Ewald coefficient
692     were tethered to the system size, and for a 512 molecule system are
693     approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04}
694     These parameters resulted in a dielectric constant of 92$\pm$14, while
695     with SF these parameters give a dielectric constant of
696 chrisfen 3064 90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
697     $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
698     0.35~\AA$^{-1}$, and these parameters resulted in a dielectric
699 gezelter 3066 constant equal to 63$\pm$1.\cite{Horn04} Calculations using SF with
700     these parameters and this water model give a dielectric constant of
701     61$\pm$1. Since the dielectric constant is dependent on $\alpha$ and
702     $R_\textrm{c}$ with the SF technique, it might be interesting to
703     investigate the dielectric dependence of the real-space Ewald
704     parameters.
705 chrisfen 3064
706 gezelter 3066 We have also mapped out the static dielectric constant of SSD/RF as a
707     function of $R_\textrm{c}$ and $\alpha$. It is apparent from this
708     table that electrostatic damping has a more pronounced effect on the
709     dipolar interactions of SSD/RF than the monopolar interactions of
710     TIP5P-E. The dielectric constant covers a much wider range and has a
711     steeper slope with increasing damping parameter.
712    
713     Although it is tempting to choose damping parameters equivalent to the
714     Ewald examples, the results of our previous work indicate that values
715     this high are destructive to both the energetics and
716 chrisfen 3064 dynamics.\cite{Fennell06} Ideally, $\alpha$ should not exceed
717     0.3~\AA$^{-1}$ for any of the cutoff values in this range. If the
718     optimal damping parameter is chosen to be midway between 0.275 and
719     0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff, then the
720     linear trend with $R_\textrm{c}$ will always keep $\alpha$ below
721     0.3~\AA$^{-1}$ for the studied cutoff radii. This linear progression
722     would give values of 0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for
723     cutoff radii of 9, 10, 11, and 12~\AA. Setting this to be the default
724 gezelter 3066 behavior for the damped SF technique will result in consistent
725 chrisfen 3064 dielectric behavior for these and other condensed molecular systems,
726     regardless of the chosen cutoff radius. The static dielectric
727     constants for TIP5P-E simulations with 9 and 12\AA\ $R_\textrm{c}$
728     values using their respective damping parameters are 76$\pm$1 and
729     75$\pm$2. These values are lower than the values reported for TIP5P-E
730     with the Ewald sum; however, they are more in line with the values
731     reported by Mahoney {\it et al.} for TIP5P while using a reaction
732     field (RF) with an infinite RF dielectric constant
733 gezelter 3066 (81.5$\pm$1.6).\cite{Mahoney00}
734    
735     We can use the same trend of $\alpha$ with $R_\textrm{c}$ for SSD/RF
736     and for a 12~\AA\ $R_\textrm{c}$, the resulting dielectric constant is
737     82.6$\pm$0.6. This value compares very favorably with the experimental
738     value of 78.3.\cite{Malmberg56} This is not surprising given that
739     early studies of the SSD model indicate a static dielectric constant
740     around 81.\cite{Liu96} The static dielectric constants for SSD/RF
741     simulations with 9 and 12\AA\ $R_\textrm{c}$ values using their
742     respective damping parameters are 82.6$\pm$0.6 and 75$\pm$2.
743    
744     As a final note, aside from a slight
745 chrisfen 3064 lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
746     not alter the other other thermodynamic properties.
747    
748 gezelter 3067 \subsection{Dynamic Properties of TIP5P-E}\label{sec:t5peDynamics}
749 chrisfen 3064
750 gezelter 3066 To look at the dynamic properties of TIP5P-E when using the SF
751 chrisfen 3064 method, 200~ps $NVE$ simulations were performed for each temperature
752     at the average density reported by the $NPT$ simulations using 9 and
753     12~\AA\ $R_\textrm{c}$ values using the ideal $\alpha$ values
754     determined above (0.2875 and 0.2125~\AA$^{-1}$). The self-diffusion
755 gezelter 3067 constants ($D$) were calculated from linear fits to the long-time
756     portion of the mean square displacement ($\langle r^{2}(t)
757     \rangle$).\cite{Allen87}
758 chrisfen 3064
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 gezelter 3067 axis $\alpha$. The body-fixed reference frame used for our
772     orientational correlation functions has the $z$-axis running along the
773     HOH bisector, and the $y$-axis connecting the two hydrogen atoms.
774     $C_l^y$ is therefore calculated from the time evolution of a vector of
775     unit length pointing between the two hydrogen atoms.
776 chrisfen 3064
777     From the orientation autocorrelation functions, we can obtain time
778 gezelter 3067 constants for rotational relaxation. The relatively short time
779 chrisfen 3064 portions (between 1 and 3~ps for water) of these curves can be fit to
780     an exponential decay to obtain these constants, and they are directly
781     comparable to water orientational relaxation times from nuclear
782     magnetic resonance (NMR). The relaxation constant obtained from
783     $C_2^y(t)$ is of particular interest because it describes the
784     relaxation of the principle axis connecting the hydrogen atoms. Thus,
785     $C_2^y(t)$ can be compared to the intermolecular portion of the
786     dipole-dipole relaxation from a proton NMR signal and should provide
787     the best estimate of the NMR relaxation time constant.\cite{Impey82}
788    
789     \begin{figure}
790     \centering
791 gezelter 3067 \includegraphics[width=5.8in]{./figures/t5peDynamics.pdf}
792 chrisfen 3064 \caption{Diffusion constants ({\it upper}) and reorientational time
793 gezelter 3066 constants ({\it lower}) for TIP5P-E using the Ewald sum and SF
794 chrisfen 3064 technique compared with experiment. Data at temperatures less than
795     0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
796     easier comparisons in the more relevant temperature regime.}
797     \label{fig:t5peDynamics}
798     \end{figure}
799     Results for the diffusion constants and orientational relaxation times
800     are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
801     apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
802 gezelter 3066 the Ewald sum are reproduced with the SF technique. The enhanced
803 gezelter 3067 diffusion (relative to experiment) at high temperatures are again the
804     product of the lower simulated densities and do not provide any
805     special insight into differences between the electrostatic summation
806     techniques. With the undamped SF technique, TIP5P-E tends to diffuse a
807     little faster than with the Ewald sum; however, the use of light to
808     moderate damping results in indistinguishable $D$ values. Though not
809     apparent in this figure, SF values at the lowest temperature are
810     approximately twice as slow as $D$ values obtained using the Ewald
811 chrisfen 3064 sum. These values support the observation from section
812     \ref{sec:t5peThermo} that there appeared to be a change to a more
813 gezelter 3067 glassy phase with the SF technique at these lower temperatures, though
814     this change is more prominent with the {\it undamped} SF method, which
815     has stronger local pairwise electrostatic interactions.
816 chrisfen 3064
817     The $\tau_2^y$ results in the lower frame of figure
818 gezelter 3067 \ref{fig:t5peDynamics} show a much greater difference between the SF
819     results and the Ewald results. At all temperatures shown, TIP5P-E
820 chrisfen 3064 relaxes faster than experiment with the Ewald sum while tracking
821 gezelter 3067 experiment fairly well when using the SF technique, independent of the
822     choice of damping constant. Their are several possible reasons for
823     this deviation between techniques. The Ewald results were calculated
824     using shorter (10ps) trajectories than the SF results (25ps). A quick
825     calculation from a 10~ps trajectory with SF with an $\alpha$ of
826     0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in $\tau_2^y$,
827     placing the result more in line with that obtained using the Ewald
828     sum. This example supports this explanation; however, recomputing the
829     results to meet a lower statistical standard is
830 chrisfen 3064 counter-productive. Assuming the Ewald results are not entirely the
831     product of poor statistics, differences in techniques to integrate the
832     orientational motion could also play a role. {\sc shake} is the most
833     commonly used technique for approximating rigid-body orientational
834     motion,\cite{Ryckaert77} whereas in {\sc oopse}, we maintain and
835     integrate the entire rotation matrix using the {\sc dlm}
836     method.\cite{Meineke05} Since {\sc shake} is an iterative constraint
837     technique, if the convergence tolerances are raised for increased
838     performance, error will accumulate in the orientational
839     motion. Finally, the Ewald results were calculated using the $NVT$
840 gezelter 3067 ensemble, while the $NVE$ ensemble was used for SF calculations. The
841     additional mode of motion due to the thermostat will alter the
842     dynamics, resulting in differences between $NVT$ and $NVE$
843 chrisfen 3064 results. These differences are increasingly noticeable as the
844     thermostat time constant decreases.
845    
846 gezelter 3067 \subsection{Comparison of Reaction Field and Damped Electrostatics for
847     SSD/RF}
848 chrisfen 3064
849     In section \ref{sec:dampingMultipoles}, we described a method for
850 gezelter 3066 applying the damped SF technique to systems containing point
851 chrisfen 3064 multipoles. The soft sticky dipole (SSD) family of water models is the
852     perfect test case because of the dipole-dipole (and
853     charge-dipole/quadrupole) interactions that are present. As alluded to
854     in the name, soft sticky dipole water models are single point
855     molecules that consist of a ``soft'' Lennard-Jones sphere, a
856     point-dipole, and a tetrahedral function for capturing the hydrogen
857     bonding nature of water - a spherical harmonic term for water-water
858     tetrahedral interactions and a point-quadrupole for interactions with
859     surrounding charges. Detailed descriptions of these models can be
860     found in other studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
861    
862     In deciding which version of the SSD model to use, we need only
863 gezelter 3066 consider that the SF technique was presented as a pairwise
864 chrisfen 3064 replacement for the Ewald summation. It has been suggested that models
865     parametrized for the Ewald summation (like TIP5P-E) would be
866     appropriate for use with a reaction field and vice versa.\cite{Rick04}
867     Therefore, we decided to test the SSD/RF water model, which was
868     parametrized for use with a reaction field, with this damped
869     electrostatic technique to see how the calculated properties change.
870    
871    
872     \begin{table}
873     \caption{Properties of SSD/RF when using different electrostatic
874     correction methods.}
875     \footnotesize
876     \centering
877     \begin{tabular}{ llccc }
878     \toprule
879     \toprule
880 gezelter 3067 & & Reaction Field [Ref. \citen{Fennell04}] & Damped Electrostatics &
881     Experiment [Ref. FIXME] \\
882 chrisfen 3064 & & $\epsilon = 80$ & $R_\textrm{c} = 12$\AA ; $\alpha = 0.2125$~\AA$^{-1}$ & \\
883     \midrule
884     $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 [\citen{CRC80}]\\
885     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 [\citen{Wagner02}] \\
886     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 [\citen{Mills73}]\\
887     $n_C$ & & 4.4 & 4.2 & 4.7 [\citen{Hura00}]\\
888     $n_H$ & & 3.7 & 3.7 & 3.5 [\citen{Soper86}]\\
889     $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 [\citen{Eisenberg69}]\\
890     $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 [\citen{Krynicki66}]\\
891     \bottomrule
892     \end{tabular}
893     \label{tab:dampedSSDRF}
894     \end{table}
895     The properties shown in table \ref{tab:dampedSSDRF} compare
896     quite well. The average density shows a modest increase when
897     using damped electrostatics in place of the reaction field. This comes
898     about because we neglect the pressure effect due to the surroundings
899     outside of the cutoff, instead relying on screening effects to
900     neutralize electrostatic interactions at long distances. The $C_p$
901     also shows a slight increase, indicating greater fluctuation in the
902     enthalpy at constant pressure. The only other differences between the
903     damped and reaction field results are the dipole reorientational time
904     constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
905     the water molecules relax more quickly and exhibit behavior very
906     similar to the experimental values. These results indicate that not
907     only is it reasonable to use damped electrostatics with SSD/RF, it is
908     recommended if capturing realistic dynamics is of primary
909     importance. This is an encouraging result because the damping methods
910     are more generally applicable than reaction field. Using damping,
911     SSD/RF can be used effectively with mixed systems, such as dissolved
912     ions, dissolved organic molecules, or even proteins.
913    
914 gezelter 3067 \subsection{Predictions of Ice Polymorph Stability}
915 chrisfen 3064
916     In an earlier work, we performed a series of free energy calculations
917     on several ice polymorphs which are stable or metastable at low
918     pressures, one of which (Ice-$i$) we observed in spontaneous
919 gezelter 3067 crystallizations of an SSD single point water model.\cite{Fennell05}
920     In this study, a distinct dependence of the free energies on the
921     interaction cutoff and correction technique was observed. Since the
922     SF technique can be used as a simple and efficient replacement for the
923     Ewald summation, our final test of this method is to see if it is
924     capable of addressing the spurious stability of the Ice-$i$ phases
925     with the various common water models. To this end, we have performed
926     thermodynamic integrations of all of the previously discussed ice
927     polymorphs using the SF technique with a cutoff radius of 12~\AA\ and
928     an $\alpha$ of 0.2125~\AA . These calculations were performed on
929     TIP5P-E and TIP4P-Ew (variants of the TIP5P and TIP4P models optimized
930     for the Ewald summation) as well as SPC/E and SSD/RF.
931 chrisfen 3064
932     \begin{table}
933     \centering
934     \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
935 gezelter 3066 using the damped SF electrostatic correction method with a
936 chrisfen 3064 variety of water models.}
937     \begin{tabular}{ lccccc }
938     \toprule
939     \toprule
940     Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
941     \cmidrule(lr){2-6}
942     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
943     \midrule
944     TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
945     TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
946     SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
947     SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
948     \bottomrule
949     \end{tabular}
950     \label{tab:dampedFreeEnergy}
951     \end{table}
952     The results of these calculations in table \ref{tab:dampedFreeEnergy}
953 gezelter 3067 show similar behavior to the Ewald results in the previous
954     study.\cite{Fennell05} The Helmholtz free energies of the ice
955     polymorphs with SSD/RF order in the same fashion, with Ice-$i$ having
956     the lowest free energy; however, the Ice-$i$ and ice B free energies
957     are quite a bit closer (nearly isoenergetic). The SPC/E results show
958     the different polymorphs to be nearly isoenergetic when using the
959     Ewald summation.\cite{Fennell05} Ice B has the lowest Helmholtz free
960     energy; however, all the polymorph results overlap within error.
961 chrisfen 3064
962     The most interesting results from these calculations come from the
963     more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
964     optimized for use with an electrostatic correction and are
965 gezelter 3067 geometrically arranged to mimic water using drastically different
966     charge distributions. In TIP5P-E, the primary location for the
967     negative charge in the molecule is assigned to the lone-pairs of the
968     oxygen, while TIP4P-Ew places the negative charge near the
969     center-of-mass along the H-O-H bisector. There is some debate as to
970     which is the proper choice for the negative charge location, and this
971     has in part led to a six-site water model that balances both of these
972     options.\cite{Vega05,Nada03} The limited results in table
973     \ref{tab:dampedFreeEnergy} support the results of Vega {\it et al.},
974     which indicate the TIP4P charge location geometry is more physically
975     valid.\cite{Vega05} With the TIP4P-Ew water model, the experimentally
976     observed polymorph (ice I$_\textrm{h}$) is the preferred form with ice
977     I$_\textrm{c}$ slightly higher in energy, though overlapping within
978     error. Additionally, the spurious ice B and Ice-$i^\prime$ structures
979     are destabilized relative to these polymorphs. TIP5P-E shows similar
980     behavior to SPC/E, where there is no real free energy distinction
981     between the various polymorphs, because many overlap within
982     error. While ice B is close in free energy to the other polymorphs,
983     these results fail to support the findings of other researchers
984     indicating the preferred form of TIP5P at 1~atm is a structure similar
985     to ice B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we
986     are looking at TIP5P-E rather than TIP5P, and the differences in the
987 chrisfen 3064 Lennard-Jones parameters could be a reason for this dissimilarity.
988     Overall, these results indicate that TIP4P-Ew is a better mimic of
989 gezelter 3067 the solid forms of water than some of the other models.
990 chrisfen 3064
991     \section{Conclusions}
992    
993 gezelter 3067 This investigation of pairwise electrostatic summation techniques
994     shows that there is a viable and computationally efficient alternative
995     to the Ewald summation. The SF method (eq. \ref{eq:DSFPot}) has
996     proven itself capable of reproducing structural, thermodynamic, and
997     dynamic quantities that are nearly quantitative matches to results
998     from far more expensive methods. Additionally, we have now extended
999     the damping formalism to electrostatic multipoles, so the damped SF
1000     potential can be used in systems that contain mixtures of charges and
1001     point multipoles.
1002    
1003     We have also provided a simple linear prescription for choosing
1004     optimal damping parameters given a choice of cutoff radius. The
1005     damping parameters were chosen to obtain a static dielectric constant
1006     as close as possible to the experimental value, which should be useful
1007     for simulating the electrostatic screening properties of liquid water
1008     accurately. The linear formula for optimal damping was the same for
1009     a complicated multipoint model as it was for a simple point-dipolar
1010     model.
1011    
1012     As in all purely pairwise cutoff methods, the damped SF method is
1013     expected to scale approximately {\it linearly} with system size, and
1014     is easily parallelizable. This should result in substantial
1015     reductions in the computational cost of performing large simulations.
1016     With the proper use of pre-computation and spline interpolation, the
1017     damped SF method is exactly the same cost as a simple real-space
1018     cutoff.
1019    
1020     We are not suggesting that there is any flaw with the Ewald sum; in
1021     fact, it is the standard by which the damped SF method has been
1022     judged. However, these results provide further evidence that in the
1023     typical simulations performed today, the Ewald summation may no longer
1024     be required to obtain the level of accuracy most researchers have come
1025     to expect.
1026    
1027 chrisfen 3064 \section{Acknowledgments}
1028     Support for this project was provided by the National Science
1029     Foundation under grant CHE-0134881. Computation time was provided by
1030 gezelter 3067 the Notre Dame Center for Research Computing. The authors would like
1031     to thank Steve Corcelli and Ed Maginn for helpful discussions and
1032     comments.
1033 chrisfen 3064
1034     \newpage
1035    
1036     \bibliographystyle{achemso}
1037     \bibliography{multipoleSFPaper}
1038    
1039    
1040     \end{document}