ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipoleSFPaper/multipoleSFPaper.tex
Revision: 3068
Committed: Tue Oct 24 21:29:44 2006 UTC (17 years, 10 months ago) by gezelter
Content type: application/x-tex
File size: 55908 byte(s)
Log Message:
THe one day paper editing session.

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 gezelter 3068 %\usepackage{endfloat}
5 chrisfen 3064 \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 3068 %\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 gezelter 3068 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
37     Department of Chemistry and Biochemistry\\
38     University of Notre Dame\\
39 chrisfen 3064 Notre Dame, Indiana 46556}
40    
41     \date{\today}
42    
43     \maketitle
44     %\doublespacing
45    
46     \begin{abstract}
47 gezelter 3066 The damped, shifted-force electrostatic potential has been shown to
48     give nearly quantitative agreement with smooth particle mesh Ewald for
49     energy differences between configurations as well as for atomic force
50     and molecular torque vectors. In this paper, we extend this technique
51     to handle interactions between electrostatic multipoles. We also
52     investigate the effects of damped and shifted electrostatics on the
53     static, thermodynamic, and dynamic properties of liquid water and
54     various polymorphs of ice. We provide a way of choosing the optimal
55     damping strength for a given cutoff radius that reproduces the static
56     dielectric constant in a variety of water models.
57 chrisfen 3064 \end{abstract}
58    
59 gezelter 3068 \newpage
60    
61 chrisfen 3064 %\narrowtext
62    
63     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64     % BODY OF TEXT
65     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66    
67     \section{Introduction}
68    
69 gezelter 3066 Over the past several years, there has been increasing interest in
70     pairwise methods for correcting electrostatic interactions in computer
71 chrisfen 3064 simulations of condensed molecular
72 gezelter 3066 systems.\cite{Wolf99,Zahn02,Kast03,Beck05,Ma05,Fennell06} These
73     techniques were developed from the observations and efforts of Wolf
74     {\it et al.} and their work towards an $\mathcal{O}(N)$ Coulombic
75     sum.\cite{Wolf99} Wolf's method of cutoff neutralization is able to
76     obtain excellent agreement with Madelung energies in ionic
77     crystals.\cite{Wolf99}
78 chrisfen 3064
79 gezelter 3066 In a recent paper, we showed that simple modifications to Wolf's
80     method could give nearly quantitative agreement with smooth particle
81     mesh Ewald (SPME) for quantities of interest in Monte Carlo
82     (i.e. configurational energy differences) and Molecular Dynamics
83     (i.e. atomic force and molecular torque vectors).\cite{Fennell06} We
84     described the undamped and damped shifted potential (SP) and shifted
85 gezelter 3067 force (SF) techniques. The potential for the damped form of the SP
86     method is given by
87 chrisfen 3064 \begin{equation}
88 gezelter 3067 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
89     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
90     \quad r\leqslant R_\textrm{c},
91     \label{eq:DSPPot}
92     \end{equation}
93     while the damped form of the SF method is given by
94     \begin{equation}
95 chrisfen 3064 \begin{split}
96     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
97     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
98     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
99     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
100     + \frac{2\alpha}{\pi^{1/2}}
101     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
102     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
103     \quad r\leqslant R_\textrm{c}.
104     \label{eq:DSFPot}
105     \end{split}
106     \end{equation}
107 gezelter 3067 (In these potentials and in all electrostatic quantities that follow,
108 gezelter 3066 the standard $4 \pi \epsilon_{0}$ has been omitted for clarity.)
109 chrisfen 3064
110 gezelter 3066 The damped SF method is an improvement over the SP method because
111     there is no discontinuity in the forces as particles move out of the
112     cutoff radius ($R_\textrm{c}$). This is accomplished by shifting the
113     forces (and potential) to zero at $R_\textrm{c}$. This is analogous to
114     neutralizing the charge as well as the force effect of the charges
115     within $R_\textrm{c}$.
116    
117     To complete the charge neutralization procedure, a self-neutralization
118 gezelter 3067 term is included in the potential. This term is constant (as long as
119     the charges and cutoff radius do not change), and exists outside the
120     normal pair-loop. It can be thought of as a contribution from a
121     charge opposite in sign, but equal in magnitude, to the central
122 gezelter 3066 charge, but which has been spread out over the surface of the cutoff
123     sphere. This term is calculated via a single loop over all charges in
124     the system. For the undamped case, the self term can be written as
125 chrisfen 3064 \begin{equation}
126 gezelter 3066 V_\textrm{self} = \frac{1}{2 R_\textrm{c}} \sum_{i=1}^N q_i^{2},
127 chrisfen 3064 \label{eq:selfTerm}
128     \end{equation}
129     while for the damped case it can be written as
130     \begin{equation}
131 gezelter 3066 V_\textrm{self} = \left(\frac{\alpha}{\sqrt{\pi}}
132     + \frac{\textrm{erfc}(\alpha
133     R_\textrm{c})}{2R_\textrm{c}}\right) \sum_{i=1}^N q_i^{2}.
134 chrisfen 3064 \label{eq:dampSelfTerm}
135     \end{equation}
136     The first term within the parentheses of equation
137 gezelter 3066 (\ref{eq:dampSelfTerm}) is identical to the self term in the Ewald
138 chrisfen 3064 summation, and comes from the utilization of the complimentary error
139     function for electrostatic damping.\cite{deLeeuw80,Wolf99}
140    
141 gezelter 3066 The SF, SP, and Wolf methods operate by neutralizing the total charge
142     contained within the cutoff sphere surrounding each particle. This is
143     accomplished by creating image charges on the surface of the cutoff
144     sphere for each pair interaction computed within the sphere. The
145     damping function applied to the potential is also an important method
146     for accelerating convergence. In the case of systems involving
147     electrostatic distributions of higher order than point charges, the
148     question remains: How will the shifting and damping need to be
149     modified in order to accommodate point multipoles?
150 chrisfen 3064
151 gezelter 3066 \section{Electrostatic Damping for Point
152     Multipoles}\label{sec:dampingMultipoles}
153 chrisfen 3064
154 gezelter 3066 To apply the SF method for systems involving point multipoles, we
155     consider separately the two techniques (shifting and damping) which
156     contribute to the effectiveness of the DSF potential.
157 chrisfen 3064
158 gezelter 3066 As noted above, shifting the potential and forces is employed to
159     neutralize the total charge contained within each cutoff sphere;
160     however, in a system composed purely of point multipoles, each cutoff
161     sphere is already neutral, so shifting becomes unnecessary.
162    
163     In a mixed system of charges and multipoles, the undamped SF potential
164     needs only to shift the force terms between charges and smoothly
165     truncate the multipolar interactions with a switching function. The
166     switching function is required for energy conservation, because a
167     discontinuity will exist in both the potential and forces at
168     $R_\textrm{c}$ in the absence of shifting terms.
169    
170     To damp the SF potential for point multipoles, we need to incorporate
171     the complimentary error function term into the standard forms of the
172     multipolar potentials. We can determine the necessary damping
173     functions by replacing $1/r$ with $\mathrm{erfc}(\alpha r)/r$ in the
174     multipole expansion. This procedure quickly becomes quite complex
175     with ``two-center'' systems, like the dipole-dipole potential, and is
176     typically approached using spherical harmonics.\cite{Hirschfelder67} A
177     simpler method for determining damped multipolar interaction
178     potentials arises when we adopt the tensor formalism described by
179     Stone.\cite{Stone02}
180    
181     The tensor formalism for electrostatic interactions involves obtaining
182     the multipole interactions from successive gradients of the monopole
183     potential. Thus, tensors of rank zero through two are
184 chrisfen 3064 \begin{equation}
185 gezelter 3066 T = \frac{1}{r_{ij}},
186     \label{eq:tensorRank1}
187 chrisfen 3064 \end{equation}
188     \begin{equation}
189 gezelter 3066 T_\alpha = \nabla_\alpha \frac{1}{r_{ij}},
190     \label{eq:tensorRank2}
191 chrisfen 3064 \end{equation}
192 gezelter 3066 \begin{equation}
193     T_{\alpha\beta} = \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
194     \label{eq:tensorRank3}
195     \end{equation}
196     where the form of the first tensor is the charge-charge potential, the
197     second gives the charge-dipole potential, and the third gives the
198     charge-quadrupole and dipole-dipole potentials.\cite{Stone02} Since
199     the force is $-\nabla V$, the forces for each potential come from the
200     next higher tensor.
201 chrisfen 3064
202 gezelter 3066 As one would do with the multipolar expansion, we can replace $r^{-1}$
203 gezelter 3067 with $\mathrm{erfc}(\alpha r)/r$ to obtain damped forms of the
204 gezelter 3066 electrostatic potentials. Equation \ref{eq:tensorRank2} generates a
205     damped charge-dipole potential,
206 chrisfen 3064 \begin{equation}
207 gezelter 3068 V_\textrm{Dcd} = -q_i\frac{\mathbf{r}_{ij}\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
208     c_1(r_{ij}),
209 chrisfen 3064 \label{eq:dChargeDipole}
210     \end{equation}
211     where $c_1(r_{ij})$ is
212     \begin{equation}
213     c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
214     + \textrm{erfc}(\alpha r_{ij}).
215     \label{eq:c1Func}
216     \end{equation}
217    
218 gezelter 3066 Equation \ref{eq:tensorRank3} generates a damped dipole-dipole potential,
219 chrisfen 3064 \begin{equation}
220 gezelter 3066 V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
221     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
222     c_2(r_{ij}) -
223     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
224     c_1(r_{ij}),
225     \label{eq:dampDipoleDipole}
226 chrisfen 3064 \end{equation}
227 gezelter 3066 where
228 chrisfen 3064 \begin{equation}
229     c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
230     + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
231     + \textrm{erfc}(\alpha r_{ij}).
232     \end{equation}
233     Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
234     term. Continuing with higher rank tensors, we can obtain the damping
235     functions for higher multipole potentials and forces. Each subsequent
236     damping function includes one additional term, and we can simplify the
237     procedure for obtaining these terms by writing out the following
238     generating function,
239     \begin{equation}
240     c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
241     {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
242     \label{eq:dampingGeneratingFunc}
243     \end{equation}
244     where,
245     \begin{equation}
246     m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
247     m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
248     m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
249     1 & m = -1\textrm{ or }0,
250     \end{array}\right.
251     \label{eq:doubleFactorial}
252     \end{equation}
253     and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
254 gezelter 3066 is similar in form to those obtained by Smith and Aguado and Madden
255     for the application of the Ewald sum to
256 chrisfen 3064 multipoles.\cite{Smith82,Smith98,Aguado03}
257    
258     Returning to the dipole-dipole example, the potential consists of a
259     portion dependent upon $r^{-5}$ and another dependent upon
260 gezelter 3066 $r^{-3}$.
261 chrisfen 3064 $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
262 gezelter 3066 respectively. The forces for the damped dipole-dipole interaction, are
263     obtained from the next higher tensor, $T_{\alpha \beta \gamma}$,
264 chrisfen 3064 \begin{equation}
265     \begin{split}
266     F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
267     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
268     c_3(r_{ij})\\ &-
269     3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
270     \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
271     \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
272     {r^5_{ij}} c_2(r_{ij}),
273     \end{split}
274     \label{eq:dampDipoleDipoleForces}
275     \end{equation}
276 gezelter 3066 Using the tensor formalism, we can dampen higher order multipolar
277     interactions using the same effective damping function that we use for
278 gezelter 3068 charge-charge interactions. This allows us to include multipoles in
279     simulations involving damped electrostatic interactions. In general,
280     if the multipolar potentials are left in $\mathbf{r}_{ij}/r_{ij}$
281     form, instead of reducing them to the more common angular forms which
282     use $\hat{r}_{ij}$ (or the resultant angles), one may simply replace
283     any $1/r_{ij}^{2n+1}$ dependence with $c_n(r_{ij}) / r_{ij}^{2n+1}$ to
284     obtain the damped version of that multipolar potential.
285 chrisfen 3064
286 gezelter 3068 As a practical consideration, we note that the evaluation of the
287     complementary error function inside a pair loop can become quite
288     costly. In practice, we pre-compute the $c_n(r)$ functions over a
289     grid of $r$ values and use cubic spline interpolation to obtain
290     estimates of these functions when necessary. Using this procedure,
291     the computational cost of damped electrostatics is only marginally
292     more costly than the undamped case.
293 chrisfen 3064
294 gezelter 3068 \section{Applications of Damped Shifted-Force Electrostatics}
295 gezelter 3067
296 gezelter 3066 Our earlier work on the SF method showed that it can give nearly
297     quantitive agreement with SPME-derived configurational energy
298     differences. The force and torque vectors in identical configurations
299     are also nearly equivalent under the damped SF potential and
300     SPME.\cite{Fennell06} Although these measures bode well for the
301     performance of the SF method in both Monte Carlo and Molecular
302     Dynamics simulations, it would be helpful to have direct comparisons
303     of structural, thermodynamic, and dynamic quantities. To address
304     this, we performed a detailed analysis of a group of simulations
305     involving water models (both point charge and multipolar) under a
306     number of different simulation conditions.
307 chrisfen 3064
308 gezelter 3066 To provide the most difficult test for the damped SF method, we have
309     chosen a model that has been optimized for use with Ewald sum, and
310     have compared the simulated properties to those computed via Ewald.
311     It is well known that water models parametrized for use with the Ewald
312     sum give calculated properties that are in better agreement with
313     experiment.\cite{vanderSpoel98,Horn04,Rick04} For these reasons, we
314     chose the TIP5P-E water model for our comparisons involving point
315     charges.\cite{Rick04}
316 chrisfen 3064
317 gezelter 3066 The soft sticky dipole (SSD) family of water models is the perfect
318     test case for the point-multipolar extension of damped electrostatics.
319     SSD water models are single point molecules that consist of a ``soft''
320     Lennard-Jones sphere, a point-dipole, and a tetrahedral function for
321     capturing the hydrogen bonding nature of water - a spherical harmonic
322     term for water-water tetrahedral interactions and a point-quadrupole
323     for interactions with surrounding charges. Detailed descriptions of
324     these models can be found in other
325     studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
326    
327     In deciding which version of the SSD model to use, we need only
328     consider that the SF technique was presented as a pairwise replacement
329     for the Ewald summation. It has been suggested that models
330     parametrized for the Ewald summation (like TIP5P-E) would be
331     appropriate for use with a reaction field and vice versa.\cite{Rick04}
332     Therefore, we decided to test the SSD/RF water model, which was
333     parametrized for use with a reaction field, with the damped
334     electrostatic technique to see how the calculated properties change.
335    
336 chrisfen 3064 The TIP5P-E water model is a variant of Mahoney and Jorgensen's
337     five-point transferable intermolecular potential (TIP5P) model for
338     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
339 gezelter 3066 maximum in liquid water near 4$^\circ$C. As with many previous point
340     charge water models (such as ST2, TIP3P, TIP4P, SPC, and SPC/E), TIP5P
341     was parametrized using a simple cutoff with no long-range
342     electrostatic
343 chrisfen 3064 correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
344     Without this correction, the pressure term on the central particle
345     from the surroundings is missing. When this correction is included,
346     systems of these particles expand to compensate for this added
347     pressure term and under-predict the density of water under standard
348 gezelter 3066 conditions. In developing TIP5P-E, Rick preserved the geometry and
349     point charge magnitudes in TIP5P and focused on altering the
350     Lennard-Jones parameters to correct the density at 298~K. With the
351 chrisfen 3064 density corrected, he compared common water properties for TIP5P-E
352     using the Ewald sum with TIP5P using a 9~\AA\ cutoff.
353    
354 gezelter 3066 In the following sections, we compare these same properties calculated
355     from TIP5P-E using the Ewald sum with TIP5P-E using the damped SF
356     technique. Our comparisons include the SF technique with a 12~\AA\
357     cutoff and an $\alpha$ = 0.0, 0.1, and 0.2~\AA$^{-1}$, as well as a
358     9~\AA\ cutoff with an $\alpha$ = 0.2~\AA$^{-1}$.
359 chrisfen 3064
360 gezelter 3066 \subsection{The Density Maximum of TIP5P-E}\label{sec:t5peDensity}
361 chrisfen 3064
362     To compare densities, $NPT$ simulations were performed with the same
363     temperatures as those selected by Rick in his Ewald summation
364     simulations.\cite{Rick04} In order to improve statistics around the
365     density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
366     25$^\circ$C, while 2~ns trajectories were obtained at all other
367     temperatures. The average densities were calculated from the later
368     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
369     method for accumulating statistics, these sequences were spliced into
370     200 segments, each providing an average density. These 200 density
371     values were used to calculate the average and standard deviation of
372     the density at each temperature.\cite{Mahoney00}
373    
374     \begin{figure}
375     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
376     \caption{Density versus temperature for the TIP5P-E water model when
377 gezelter 3066 using the Ewald summation (Ref. \citen{Rick04}) and the SF method with
378     varying cutoff radii and damping coefficients. The pressure term from
379     the image-charge shell is larger than that provided by the
380     reciprocal-space portion of the Ewald summation, leading to slightly
381     lower densities. This effect is more visible with the 9~\AA\ cutoff,
382     where the image charges exert a greater force on the central
383     particle. The error bars for the SF methods show the average one-sigma
384     uncertainty of the density measurement, and this uncertainty is the
385     same for all the SF curves.}
386 chrisfen 3064 \label{fig:t5peDensities}
387     \end{figure}
388     Figure \ref{fig:t5peDensities} shows the densities calculated for
389 gezelter 3066 TIP5P-E using differing electrostatic corrections overlaid with the
390     experimental values.\cite{CRC80} The densities when using the SF
391     technique are close to, but typically lower than, those calculated
392 chrisfen 3064 using the Ewald summation. These slightly reduced densities indicate
393     that the pressure component from the image charges at R$_\textrm{c}$
394     is larger than that exerted by the reciprocal-space portion of the
395 gezelter 3066 Ewald summation. Bringing the image charges closer to the central
396 chrisfen 3064 particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
397     preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
398     charge interactions on the central particle and results in a further
399     reduction of the densities.
400    
401     Because the strength of the image charge interactions has a noticeable
402     effect on the density, we would expect the use of electrostatic
403     damping to also play a role in these calculations. Larger values of
404     $\alpha$ weaken the pair-interactions; and since electrostatic damping
405     is distance-dependent, force components from the image charges will be
406     reduced more than those from particles close the the central
407     charge. This effect is visible in figure \ref{fig:t5peDensities} with
408 gezelter 3066 the damped SF sums showing slightly higher densities; however, it is
409     clear that the choice of cutoff radius plays a much more important
410     role in the resulting densities.
411 chrisfen 3064
412 gezelter 3066 All of the above density calculations were performed with systems of
413     512 water molecules. Rick observed a system size dependence of the
414     computed densities when using the Ewald summation, most likely due to
415     his tying of the convergence parameter to the box
416 chrisfen 3064 dimensions.\cite{Rick04} For systems of 256 water molecules, the
417     calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
418     system size of 256 molecules would force the use of a shorter
419 gezelter 3066 R$_\textrm{c}$ when using the SF technique, and this would also lower
420     the densities. Moving to larger systems, as long as the R$_\textrm{c}$
421     remains at a fixed value, we would expect the densities to remain
422     constant.
423 chrisfen 3064
424 gezelter 3066 \subsection{Liquid Structure of TIP5P-E}\label{sec:t5peLiqStructure}
425 chrisfen 3064
426 gezelter 3068 The experimentally-determined oxygen-oxygen pair correlation function
427     ($g_\textrm{OO}(r)$) for liquid water has been compared in great
428     detail with predictions of the various common water models, and TIP5P
429     was found to be in better agreement than other rigid, non-polarizable
430     models.\cite{Sorenson00} This excellent agreement with experiment was
431     maintained when Rick developed TIP5P-E.\cite{Rick04} To check whether
432     the choice of using the Ewald summation or the SF technique alters the
433     liquid structure, we calculated this correlation function at 298~K and
434     1~atm for the parameters used in the previous section.
435 chrisfen 3064
436     \begin{figure}
437     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
438 gezelter 3068 \caption{The oxygen-oxygen pair correlation functions calculated for
439     TIP5P-E at 298~K and 1atm while using the Ewald summation
440     (Ref. \cite{Rick04}) and the SF technique with varying
441     parameters. Even with the lower densities obtained using the SF
442     technique, the correlation functions are essentially identical.}
443 chrisfen 3064 \label{fig:t5peGofRs}
444     \end{figure}
445 gezelter 3068 The pair correlation functions calculated for TIP5P-E while using the
446     SF technique with various parameters are overlaid on the same function
447     obtained while using the Ewald summation in
448 chrisfen 3064 figure~\ref{fig:t5peGofRs}. The differences in density do not appear
449 gezelter 3068 to have any effect on the liquid structure as the correlation
450     functions are indistinguishable. These results do indicate that
451 chrisfen 3064 $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
452     correction.
453    
454 gezelter 3066 \subsection{Thermodynamic Properties of TIP5P-E}\label{sec:t5peThermo}
455 chrisfen 3064
456 gezelter 3066 In addition to the density and structual features of the liquid, there
457     are a variety of thermodynamic quantities that can be calculated for
458     water and compared directly to experimental values. Some of these
459     additional quantities include the latent heat of vaporization ($\Delta
460     H_\textrm{vap}$), the constant pressure heat capacity ($C_p$), the
461     isothermal compressibility ($\kappa_T$), the thermal expansivity
462     ($\alpha_p$), and the static dielectric constant ($\epsilon$). All of
463     these properties were calculated for TIP5P-E with the Ewald summation,
464 gezelter 3068 so they provide a good set of reference data for comparisons involving
465     the SF technique.
466 chrisfen 3064
467     The $\Delta H_\textrm{vap}$ is the enthalpy change required to
468     transform one mole of substance from the liquid phase to the gas
469     phase.\cite{Berry00} In molecular simulations, this quantity can be
470     determined via
471     \begin{equation}
472     \begin{split}
473 gezelter 3066 \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq} \\
474     &= E_\textrm{gas} - E_\textrm{liq}
475     + P(V_\textrm{gas} - V_\textrm{liq}) \\
476     &\approx -\frac{\langle U_\textrm{liq}\rangle}{N}+ RT,
477 chrisfen 3064 \end{split}
478     \label{eq:DeltaHVap}
479     \end{equation}
480 gezelter 3066 where $E$ is the total energy, $U$ is the potential energy, $P$ is the
481 chrisfen 3064 pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
482     the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
483     seen in the last line of equation (\ref{eq:DeltaHVap}), we can
484     approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
485     state. This allows us to cancel the kinetic energy terms, leaving only
486 gezelter 3068 the $U_\textrm{liq}$ term. Additionally, the $PV$ term for the gas is
487 chrisfen 3064 several orders of magnitude larger than that of the liquid, so we can
488 gezelter 3068 neglect the liquid $PV$ term.
489 chrisfen 3064
490     The remaining thermodynamic properties can all be calculated from
491     fluctuations of the enthalpy, volume, and system dipole
492     moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
493     enthalpy in constant pressure simulations via
494     \begin{equation}
495     \begin{split}
496 gezelter 3066 C_p = \left(\frac{\partial H}{\partial T}\right)_{N,P}
497 chrisfen 3064 = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
498     \end{split}
499     \label{eq:Cp}
500     \end{equation}
501     where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
502     \begin{equation}
503     \begin{split}
504     \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
505 gezelter 3068 = \frac{(\langle V^2\rangle_{NPT} - \langle V\rangle^{2}_{NPT})}
506 gezelter 3066 {k_BT\langle V\rangle_{NPT}},
507 chrisfen 3064 \end{split}
508     \label{eq:kappa}
509     \end{equation}
510     and $\alpha_p$ can be calculated via
511     \begin{equation}
512     \begin{split}
513     \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
514 gezelter 3066 = \frac{(\langle VH\rangle_{NPT}
515     - \langle V\rangle_{NPT}\langle H\rangle_{NPT})}
516     {k_BT^2\langle V\rangle_{NPT}}.
517 chrisfen 3064 \end{split}
518     \label{eq:alpha}
519     \end{equation}
520     Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
521     be calculated for systems of non-polarizable substances via
522     \begin{equation}
523     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
524     \label{eq:staticDielectric}
525     \end{equation}
526     where $\epsilon_0$ is the permittivity of free space and $\langle
527     M^2\rangle$ is the fluctuation of the system dipole
528     moment.\cite{Allen87} The numerator in the fractional term in equation
529     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
530     dipole moment, identical to the quantity calculated in the
531     finite-system Kirkwood $g$ factor ($G_k$):
532     \begin{equation}
533     G_k = \frac{\langle M^2\rangle}{N\mu^2},
534     \label{eq:KirkwoodFactor}
535     \end{equation}
536     where $\mu$ is the dipole moment of a single molecule of the
537     homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
538     fluctuation term in both equation (\ref{eq:staticDielectric}) and
539     \ref{eq:KirkwoodFactor} is calculated as follows,
540     \begin{equation}
541     \begin{split}
542     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
543     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
544     &= \langle M_x^2+M_y^2+M_z^2\rangle
545     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
546     + \langle M_x\rangle^2).
547     \end{split}
548     \label{eq:fluctBoxDipole}
549     \end{equation}
550     This fluctuation term can be accumulated during the simulation;
551     however, it converges rather slowly, thus requiring multi-nanosecond
552     simulation times.\cite{Horn04} In the case of tin-foil boundary
553     conditions, the dielectric/surface term of the Ewald summation is
554 gezelter 3066 equal to zero. Since the SF method also lacks this
555 chrisfen 3064 dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
556     valid for determining static dielectric constants.
557    
558     All of the above properties were calculated from the same trajectories
559     used to determine the densities in section \ref{sec:t5peDensity}
560     except for the static dielectric constants. The $\epsilon$ values were
561     accumulated from 2~ns $NVE$ ensemble trajectories with system densities
562     fixed at the average values from the $NPT$ simulations at each of the
563     temperatures. The resulting values are displayed in figure
564     \ref{fig:t5peThermo}.
565     \begin{figure}
566     \centering
567 gezelter 3067 \includegraphics[width=5.8in]{./figures/t5peThermo.pdf}
568 chrisfen 3064 \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
569 gezelter 3066 and the SF techniques along with the experimental values. Units
570 chrisfen 3064 for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
571     cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
572     and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
573     reference \cite{Rick04}. Experimental values for $\Delta
574     H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
575     \cite{Kell75}. Experimental values for $C_p$ are from reference
576     \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
577     \cite{Malmberg56}.}
578     \label{fig:t5peThermo}
579     \end{figure}
580    
581 gezelter 3066 For all of the properties computed, the trends with temperature
582     obtained when using the Ewald summation are reproduced with the SF
583     technique. One noticeable difference between the properties calculated
584     using the two methods are the lower $\Delta H_\textrm{vap}$ values
585     when using SF. This is to be expected due to the direct weakening of
586     the electrostatic interaction through forced neutralization. This
587     results in an increase of the intermolecular potential producing lower
588     values from equation (\ref{eq:DeltaHVap}). The slopes of these values
589     with temperature are similar to that seen using the Ewald summation;
590     however, they are both steeper than the experimental trend, indirectly
591     resulting in the inflated $C_p$ values at all temperatures.
592 chrisfen 3064
593     Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
594     all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
595 gezelter 3066 and $C_p$ results, the deviations between experiment and simulation in
596     this region are not the fault of the electrostatic summation methods
597     but are due to the geometry and parameters of the TIP5P class of water
598     models. Like most rigid, non-polarizable, point-charge water models,
599     the density decreases with temperature at a much faster rate than
600     experiment (see figure \ref{fig:t5peDensities}). This reduced density
601     leads to the inflated compressibility and expansivity values at higher
602     temperatures seen here in figure \ref{fig:t5peThermo}. Incorporation
603     of polarizability and many-body effects are required in order for
604     water models to overcome differences between simulation-based and
605     experimentally determined densities at these higher
606 chrisfen 3064 temperatures.\cite{Laasonen93,Donchev06}
607    
608     At temperatures below the freezing point for experimental water, the
609 gezelter 3066 differences between SF and the Ewald summation results are more
610 chrisfen 3064 apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
611     indicate a more pronounced transition in the supercooled regime,
612 gezelter 3068 particularly in the case of SF without damping. This points to the
613     onset of a more frustrated or glassy behavior for the undamped and
614     weakly-damped SF simulations of TIP5P-E at temperatures below 250~K
615     than is seen from the Ewald sum. Undamped SF electrostatics has a
616     stronger contribution from nearby charges. Damping these local
617     interactions or using a reciprocal-space method makes the water less
618     sensitive to ordering on a shorter length scale. We can recover
619     nearly quantitative agreement with the Ewald results by increasing the
620     damping constant.
621 chrisfen 3064
622     The final thermodynamic property displayed in figure
623     \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
624 gezelter 3066 between the Ewald and SF methods (and with experiment). It is known
625     that the dielectric constant is dependent upon and is quite sensitive
626     to the imposed boundary conditions.\cite{Neumann80,Neumann83} This is
627     readily apparent in the converged $\epsilon$ values accumulated for
628     the SF simulations. Lack of a damping function results in dielectric
629 chrisfen 3064 constants significantly smaller than those obtained using the Ewald
630     sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
631     agreement considerably. It should be noted that the choice of the
632 gezelter 3066 ``Ewald coefficient'' ($\kappa$) and real-space cutoff values also
633     have a significant effect on the calculated static dielectric constant
634     when using the Ewald summation. In the simulations of TIP5P-E with the
635     Ewald sum, this screening parameter was tethered to the simulation box
636     size (as was the $R_\textrm{c}$).\cite{Rick04} In general, systems
637     with larger screening parameters reported larger dielectric constant
638     values, the same behavior we see here with {\sc sf}; however, the
639     choice of cutoff radius also plays an important role.
640 chrisfen 3064
641 gezelter 3068 \subsection{Optimal Damping Coefficients for Damped
642     Electrostatics}\label{sec:t5peDielectric}
643 chrisfen 3064
644 gezelter 3066 In the previous section, we observed that the choice of damping
645     coefficient plays a major role in the calculated dielectric constant
646     for the SF method. Similar damping parameter behavior was observed in
647     the long-time correlated motions of the NaCl crystal.\cite{Fennell06}
648     The static dielectric constant is calculated from the long-time
649     fluctuations of the system's accumulated dipole moment
650     (Eq. (\ref{eq:staticDielectric})), so it is quite sensitive to the
651     choice of damping parameter. Since $\alpha$ is an adjustable
652     parameter, it would be best to choose optimal damping constants such
653     that any arbitrary choice of cutoff radius will properly capture the
654     dielectric behavior of the liquid.
655 chrisfen 3064
656     In order to find these optimal values, we mapped out the static
657     dielectric constant as a function of both the damping parameter and
658 gezelter 3066 cutoff radius for TIP5P-E and for a point-dipolar water model
659     (SSD/RF). To calculate the static dielectric constant, we performed
660     5~ns $NPT$ calculations on systems of 512 water molecules and averaged
661     over the converged region (typically the later 2.5~ns) of equation
662     (\ref{eq:staticDielectric}). The selected cutoff radii ranged from 9,
663     10, 11, and 12~\AA , and the damping parameter values ranged from 0.1
664     to 0.35~\AA$^{-1}$.
665 chrisfen 3064
666     \begin{table}
667     \centering
668 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
669     cutoff radius $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
670 chrisfen 3064 \vspace{6pt}
671 gezelter 3066 \begin{tabular}{ lccccccccc }
672 chrisfen 3064 \toprule
673     \toprule
674 gezelter 3066 & \multicolumn{4}{c}{TIP5P-E} & & \multicolumn{4}{c}{SSD/RF} \\
675     & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} & & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
676     \cmidrule(lr){2-5} \cmidrule(lr){7-10}
677     $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 & & 9 & 10 & 11 & 12 \\
678 chrisfen 3064 \midrule
679 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 \\
680     & \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} \\
681     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 \\
682     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 \\
683     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 \\
684     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 \\
685     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 \\
686     & \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} \\
687     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 \\
688     & \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} \\
689     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 \\
690 chrisfen 3064 \bottomrule
691     \end{tabular}
692 gezelter 3066 \label{tab:DielectricMap}
693 chrisfen 3064 \end{table}
694 gezelter 3066
695 chrisfen 3064 The results of these calculations are displayed in table
696 gezelter 3066 \ref{tab:DielectricMap}. The dielectric constants for both models
697     decrease linearly with increasing cutoff radii ($R_\textrm{c}$) and
698     increase linearly with increasing damping ($\alpha$). Another point
699     to note is that choosing $\alpha$ and $R_\textrm{c}$ identical to
700     those used with the Ewald summation results in the same calculated
701     dielectric constant. As an example, in the paper outlining the
702     development of TIP5P-E, the real-space cutoff and Ewald coefficient
703     were tethered to the system size, and for a 512 molecule system are
704     approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04}
705     These parameters resulted in a dielectric constant of 92$\pm$14, while
706     with SF these parameters give a dielectric constant of
707 chrisfen 3064 90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
708     $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
709     0.35~\AA$^{-1}$, and these parameters resulted in a dielectric
710 gezelter 3066 constant equal to 63$\pm$1.\cite{Horn04} Calculations using SF with
711     these parameters and this water model give a dielectric constant of
712     61$\pm$1. Since the dielectric constant is dependent on $\alpha$ and
713     $R_\textrm{c}$ with the SF technique, it might be interesting to
714 gezelter 3068 investigate the dependence of the static dielectric constant on the
715     choice of convergence parameters ($R_\textrm{c}$ and $\kappa$)
716     utilized in most implementations of the Ewald sum.
717 chrisfen 3064
718 gezelter 3068 It is also apparent from this table that electrostatic damping has a
719     more pronounced effect on the dipolar interactions of SSD/RF than the
720     monopolar interactions of TIP5P-E. The dielectric constant covers a
721     much wider range and has a steeper slope with increasing damping
722     parameter.
723 gezelter 3066
724     Although it is tempting to choose damping parameters equivalent to the
725 gezelter 3068 Ewald examples to obtain quantitative agreement, the results of our
726     previous work indicate that values this high are destructive to both
727     the energetics and dynamics.\cite{Fennell06} Ideally, $\alpha$ should
728     not exceed 0.3~\AA$^{-1}$ for any of the cutoff values in this
729     range. If the optimal damping parameter is chosen to be midway between
730     0.275 and 0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff,
731     then the linear trend with $R_\textrm{c}$ will always keep $\alpha$
732     below 0.3~\AA$^{-1}$ for the studied cutoff radii. This linear
733     progression would give values of 0.2875, 0.2625, 0.2375, and
734     0.2125~\AA$^{-1}$ for cutoff radii of 9, 10, 11, and 12~\AA. Setting
735     this to be the default behavior for the damped SF technique will
736     result in consistent dielectric behavior for these and other condensed
737     molecular systems, regardless of the chosen cutoff radius. The static
738     dielectric constants for TIP5P-E simulations with 9 and 12\AA\
739     $R_\textrm{c}$ values using their respective damping parameters are
740     76$\pm$1 and 75$\pm$2. These values are lower than the values reported
741     for TIP5P-E with the Ewald sum; however, they are more in line with
742     the values reported by Mahoney {\it et al.} for TIP5P while using a
743     reaction field (RF) with an infinite RF dielectric constant
744     (81.5$\pm$1.6).\cite{Mahoney00}
745 gezelter 3066
746 gezelter 3068 Using the same linear relationship utilized with TIP5P-E above, the
747     static dielectric constants for SSD/RF with $R_\textrm{c}$ values of 9
748     and 12~\AA\ are 88$\pm$8 and 82.6$\pm$0.6. These values compare
749     favorably with the experimental value of 78.3.\cite{Malmberg56} These
750     results are also not surprising given that early studies of the SSD
751     model indicated a static dielectric constant around 81.\cite{Liu96}
752 gezelter 3066
753 gezelter 3068 As a final note on optimal damping parameters, aside from a slight
754 chrisfen 3064 lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
755 gezelter 3068 not alter any of the other thermodynamic properties.
756 chrisfen 3064
757 gezelter 3067 \subsection{Dynamic Properties of TIP5P-E}\label{sec:t5peDynamics}
758 chrisfen 3064
759 gezelter 3068 To look at the dynamic properties of TIP5P-E when using the SF method,
760     200~ps $NVE$ simulations were performed for each temperature at the
761     average density obtained from the $NPT$ simulations. $R_\textrm{c}$
762     values of 9 and 12~\AA\ and the ideal $\alpha$ values determined above
763     (0.2875 and 0.2125~\AA$^{-1}$) were used for the damped
764     electrostatics. The self-diffusion constants (D) were calculated from
765     linear fits to the long-time portion of the mean square displacement
766     ($\langle r^{2}(t) \rangle$).\cite{Allen87}
767 chrisfen 3064
768     In addition to translational diffusion, orientational relaxation times
769     were calculated for comparisons with the Ewald simulations and with
770     experiments. These values were determined from the same 200~ps $NVE$
771     trajectories used for translational diffusion by calculating the
772     orientational time correlation function,
773     \begin{equation}
774     C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
775     \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
776     \label{eq:OrientCorr}
777     \end{equation}
778     where $P_l$ is the Legendre polynomial of order $l$ and
779     $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
780 gezelter 3067 axis $\alpha$. The body-fixed reference frame used for our
781     orientational correlation functions has the $z$-axis running along the
782     HOH bisector, and the $y$-axis connecting the two hydrogen atoms.
783     $C_l^y$ is therefore calculated from the time evolution of a vector of
784     unit length pointing between the two hydrogen atoms.
785 chrisfen 3064
786     From the orientation autocorrelation functions, we can obtain time
787 gezelter 3067 constants for rotational relaxation. The relatively short time
788 chrisfen 3064 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 gezelter 3067 \includegraphics[width=5.8in]{./figures/t5peDynamics.pdf}
801 chrisfen 3064 \caption{Diffusion constants ({\it upper}) and reorientational time
802 gezelter 3066 constants ({\it lower}) for TIP5P-E using the Ewald sum and SF
803 chrisfen 3064 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 gezelter 3066 the Ewald sum are reproduced with the SF technique. The enhanced
812 gezelter 3067 diffusion (relative to experiment) at high temperatures are again the
813     product of the lower simulated densities and do not provide any
814     special insight into differences between the electrostatic summation
815 gezelter 3068 techniques. Though not apparent in this figure, SF values at the
816     lowest temperature are approximately twice as slow as $D$ values
817     obtained using the Ewald sum. These values support the observation
818     from section \ref{sec:t5peThermo} that the SF simulations result in a
819     slightly more viscous supercooled region than is obtained using the
820     Ewald sum.
821 chrisfen 3064
822     The $\tau_2^y$ results in the lower frame of figure
823 gezelter 3067 \ref{fig:t5peDynamics} show a much greater difference between the SF
824     results and the Ewald results. At all temperatures shown, TIP5P-E
825 chrisfen 3064 relaxes faster than experiment with the Ewald sum while tracking
826 gezelter 3067 experiment fairly well when using the SF technique, independent of the
827 gezelter 3068 choice of damping constant. There are several possible reasons for
828 gezelter 3067 this deviation between techniques. The Ewald results were calculated
829 gezelter 3068 using shorter (10~ps) trajectories than the SF results (200~ps).
830     Calculation of these SF values from a 10~ps trajectory (with
831     subsequently lower accuracy) showed a 0.4~ps drop in $\tau_2^y$,
832 gezelter 3067 placing the result more in line with that obtained using the Ewald
833 gezelter 3068 sum. Recomputing correlation times to meet a lower statistical
834     standard is counter-productive, however. Assuming the Ewald results
835     are not entirely the product of poor statistics, differences in
836     techniques to integrate the orientational motion could also play a
837     role. {\sc shake} is the most commonly used technique for
838     approximating rigid-body orientational motion,\cite{Ryckaert77}
839     whereas in {\sc oopse}, we maintain and integrate the entire rotation
840     matrix using the {\sc dlm} method.\cite{Meineke05} Since {\sc shake}
841     is an iterative constraint technique, if the convergence tolerances
842     are raised for increased performance, error will accumulate in the
843     orientational motion. Finally, the Ewald results were calculated using
844     the $NVT$ ensemble, while the $NVE$ ensemble was used for SF
845     calculations. The motion due to the extended variable (the thermostat)
846     will always alter the dynamics, resulting in differences between $NVT$
847     and $NVE$ results. These differences are increasingly noticeable as
848     the time constant for the thermostat decreases.
849 chrisfen 3064
850 gezelter 3067 \subsection{Comparison of Reaction Field and Damped Electrostatics for
851     SSD/RF}
852 chrisfen 3064
853 gezelter 3068 SSD/RF was parametrized for use with a reaction field, which is a
854     common and relatively inexpensive way of handling long-range
855     electrostatic corrections in dipolar systems.\cite{Onsager36}
856     Although there is no reason to expect that damped electrostatics will
857     behave in a similar fashion to the reaction field, it is well known
858     that model that are parametrized for use with Ewald do better than
859     unoptimized models under the influence of a reaction
860     field.\cite{Rick04} We compared a number of properties of SSD/RF that
861     had previously been computed using a reaction field with those same
862     values under damped electrostatics.
863 chrisfen 3064
864 gezelter 3068 The properties shown in table \ref{tab:dampedSSDRF} show that damped
865     electrostatics produces even better agreement with experiment than is
866     obtained via reaction field. The average density shows a modest
867     increase when using damped electrostatics in place of the reaction
868     field. This comes about because we neglect the pressure effect due to
869     the surroundings outside of the cutoff, instead relying on screening
870     effects to neutralize electrostatic interactions at long
871     distances. The $C_p$ also shows a slight increase, indicating greater
872     fluctuation in the enthalpy at constant pressure. The only other
873     differences between the damped electrostatics and the reaction field
874     results are the dipole reorientational time constants, $\tau_1$ and
875     $\tau_2$. When using damped electrostatics, the water molecules relax
876     more quickly and exhibit behavior closer to the experimental
877     values. These results indicate that since there is no need to specify
878     an external dielectric constant with the damped electrostatics, it is
879     almost certainly a better choice for dipolar simulations than the
880     reaction field method. Using damped electrostatics, SSD/RF can be
881     used effectively with mixed charge / dipolar systems, such as
882     dissolved ions, dissolved organic molecules, or even proteins.
883 chrisfen 3064
884     \begin{table}
885 gezelter 3068 \caption{Properties of SSD/RF when using reaction field or damped
886     electrostatics. Simulations were carried out at 298~K, 1~atm, and
887     with 512 molecules.}
888 chrisfen 3064 \footnotesize
889     \centering
890     \begin{tabular}{ llccc }
891     \toprule
892     \toprule
893 gezelter 3068 & & Reaction Field (Ref. \citen{Fennell04}) & Damped Electrostatics &
894     Experiment [Ref.] \\
895 chrisfen 3064 & & $\epsilon = 80$ & $R_\textrm{c} = 12$\AA ; $\alpha = 0.2125$~\AA$^{-1}$ & \\
896     \midrule
897     $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 [\citen{CRC80}]\\
898     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 [\citen{Wagner02}] \\
899     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 [\citen{Mills73}]\\
900     $n_C$ & & 4.4 & 4.2 & 4.7 [\citen{Hura00}]\\
901     $n_H$ & & 3.7 & 3.7 & 3.5 [\citen{Soper86}]\\
902     $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 [\citen{Eisenberg69}]\\
903     $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 [\citen{Krynicki66}]\\
904     \bottomrule
905     \end{tabular}
906     \label{tab:dampedSSDRF}
907     \end{table}
908    
909 gezelter 3067 \subsection{Predictions of Ice Polymorph Stability}
910 chrisfen 3064
911 gezelter 3068 In an earlier paper, we performed a series of free energy calculations
912 chrisfen 3064 on several ice polymorphs which are stable or metastable at low
913     pressures, one of which (Ice-$i$) we observed in spontaneous
914 gezelter 3068 crystallizations of an early version of the SSD/RF water
915     model.\cite{Fennell05} In this study, a distinct dependence of the
916     computed free energies on the cutoff radius and electrostatic
917     summation method was observed. Since the SF technique can be used as
918     a simple and efficient replacement for the Ewald summation, our final
919     test of this method is to see if it is capable of addressing the
920     spurious stability of the Ice-$i$ phases with the various common water
921     models. To this end, we have performed thermodynamic integrations of
922     all of the previously discussed ice polymorphs using the SF technique
923     with a cutoff radius of 12~\AA\ and an $\alpha$ of 0.2125~\AA . These
924     calculations were performed on TIP5P-E and TIP4P-Ew (variants of the
925     TIP5P and TIP4P models optimized for the Ewald summation) as well as
926     SPC/E and SSD/RF.
927 chrisfen 3064
928     \begin{table}
929     \centering
930     \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
931 gezelter 3066 using the damped SF electrostatic correction method with a
932 chrisfen 3064 variety of water models.}
933     \begin{tabular}{ lccccc }
934     \toprule
935     \toprule
936     Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
937     \cmidrule(lr){2-6}
938     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
939     \midrule
940     TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
941     TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
942     SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
943     SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
944     \bottomrule
945     \end{tabular}
946     \label{tab:dampedFreeEnergy}
947     \end{table}
948     The results of these calculations in table \ref{tab:dampedFreeEnergy}
949 gezelter 3067 show similar behavior to the Ewald results in the previous
950     study.\cite{Fennell05} The Helmholtz free energies of the ice
951     polymorphs with SSD/RF order in the same fashion, with Ice-$i$ having
952     the lowest free energy; however, the Ice-$i$ and ice B free energies
953     are quite a bit closer (nearly isoenergetic). The SPC/E results show
954 gezelter 3068 the different polymorphs to be nearly isoenergetic. This is the same
955     behavior observed using an Ewald correction.\cite{Fennell05} Ice B has
956     the lowest Helmholtz free energy for SPC/E; however, all the polymorph
957     results overlap within the error estimates.
958 chrisfen 3064
959     The most interesting results from these calculations come from the
960     more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
961     optimized for use with an electrostatic correction and are
962 gezelter 3067 geometrically arranged to mimic water using drastically different
963 gezelter 3068 charge distributions. In TIP5P-E, the primary location for the
964 gezelter 3067 negative charge in the molecule is assigned to the lone-pairs of the
965     oxygen, while TIP4P-Ew places the negative charge near the
966     center-of-mass along the H-O-H bisector. There is some debate as to
967     which is the proper choice for the negative charge location, and this
968     has in part led to a six-site water model that balances both of these
969     options.\cite{Vega05,Nada03} The limited results in table
970     \ref{tab:dampedFreeEnergy} support the results of Vega {\it et al.},
971 gezelter 3068 which indicate the TIP4P charge location geometry performs better
972     under some circumstances.\cite{Vega05} With the TIP4P-Ew water model,
973     the experimentally observed polymorph (ice I$_\textrm{h}$) is the
974     preferred form with ice I$_\textrm{c}$ slightly higher in energy,
975     though overlapping within error. Additionally, the spurious ice B and
976     Ice-$i^\prime$ structures are destabilized relative to these
977     polymorphs. TIP5P-E shows similar behavior to SPC/E, where there is no
978     real free energy distinction between the various polymorphs. While ice
979     B is close in free energy to the other polymorphs, these results fail
980     to support the findings of other researchers indicating the preferred
981     form of TIP5P at 1~atm is a structure similar to ice
982     B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we were
983     looking at TIP5P-E rather than TIP5P, and the differences in the
984     Lennard-Jones parameters could cause this discrepancy. Overall, these
985     results indicate that TIP4P-Ew is a better mimic of the solid forms of
986     water than some of the other models.
987 chrisfen 3064
988     \section{Conclusions}
989    
990 gezelter 3067 This investigation of pairwise electrostatic summation techniques
991     shows that there is a viable and computationally efficient alternative
992     to the Ewald summation. The SF method (eq. \ref{eq:DSFPot}) has
993     proven itself capable of reproducing structural, thermodynamic, and
994     dynamic quantities that are nearly quantitative matches to results
995     from far more expensive methods. Additionally, we have now extended
996     the damping formalism to electrostatic multipoles, so the damped SF
997     potential can be used in systems that contain mixtures of charges and
998     point multipoles.
999    
1000     We have also provided a simple linear prescription for choosing
1001     optimal damping parameters given a choice of cutoff radius. The
1002     damping parameters were chosen to obtain a static dielectric constant
1003     as close as possible to the experimental value, which should be useful
1004     for simulating the electrostatic screening properties of liquid water
1005     accurately. The linear formula for optimal damping was the same for
1006     a complicated multipoint model as it was for a simple point-dipolar
1007     model.
1008    
1009     As in all purely pairwise cutoff methods, the damped SF method is
1010     expected to scale approximately {\it linearly} with system size, and
1011     is easily parallelizable. This should result in substantial
1012     reductions in the computational cost of performing large simulations.
1013     With the proper use of pre-computation and spline interpolation, the
1014 gezelter 3068 damped SF method is essentially the same cost as a simple real-space
1015 gezelter 3067 cutoff.
1016    
1017     We are not suggesting that there is any flaw with the Ewald sum; in
1018     fact, it is the standard by which the damped SF method has been
1019     judged. However, these results provide further evidence that in the
1020     typical simulations performed today, the Ewald summation may no longer
1021     be required to obtain the level of accuracy most researchers have come
1022     to expect.
1023    
1024 chrisfen 3064 \section{Acknowledgments}
1025     Support for this project was provided by the National Science
1026     Foundation under grant CHE-0134881. Computation time was provided by
1027 gezelter 3067 the Notre Dame Center for Research Computing. The authors would like
1028     to thank Steve Corcelli and Ed Maginn for helpful discussions and
1029     comments.
1030 chrisfen 3064
1031     \newpage
1032    
1033     \bibliographystyle{achemso}
1034     \bibliography{multipoleSFPaper}
1035    
1036    
1037     \end{document}