ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipoleSFPaper/multipoleSFPaper.tex
Revision: 3065
Committed: Tue Oct 24 06:27:31 2006 UTC (18 years, 9 months ago) by chrisfen
Content type: application/x-tex
File size: 56789 byte(s)
Log Message:
fixed the bib file and some references

File Contents

# User Rev Content
1 chrisfen 3064 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2     \documentclass[12pt]{article}
3     \usepackage{tabls}
4     \usepackage{endfloat}
5     \usepackage[tbtags]{amsmath}
6     \usepackage{amsmath,bm}
7     \usepackage{amssymb}
8     \usepackage{mathrsfs}
9     \usepackage{setspace}
10     \usepackage{tabularx}
11     \usepackage{graphicx}
12     \usepackage{booktabs}
13     \usepackage{colortbl}
14     \usepackage[ref]{overcite}
15     \pagestyle{plain}
16     \pagenumbering{arabic}
17     \oddsidemargin 0.0cm \evensidemargin 0.0cm
18     \topmargin -21pt \headsep 10pt
19     \textheight 9.0in \textwidth 6.5in
20     \brokenpenalty=10000
21     \renewcommand{\baselinestretch}{1.2}
22     \renewcommand\citemid{\ } % no comma in optional reference note
23    
24     \begin{document}
25    
26     \title{Pairwise electrostatic corrections for monopolar and multipolar systems}
27    
28     \author{Christopher J. Fennell and J. Daniel Gezelter \\
29     Department of Chemistry and Biochemistry\\ University of Notre Dame\\
30     Notre Dame, Indiana 46556}
31    
32     \date{\today}
33    
34     \maketitle
35     %\doublespacing
36    
37     \begin{abstract}
38     \end{abstract}
39    
40     %\narrowtext
41    
42     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43     % BODY OF TEXT
44     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45    
46     \section{Introduction}
47    
48     Over the past several years, there has been increased interest in
49     pairwise methods for correcting electrostatic interaction in computer
50     simulations of condensed molecular
51 chrisfen 3065 systems.\cite{Wolf99,Zahn02,Kast03,Fennell06}
52 chrisfen 3064
53     \section{Shifted Force Pairwise Electrostatic Correction}
54    
55     In our previous look at pairwise electrostatic correction methods, we
56     described the undamped and damped shifted potential {\sc sp} and
57     shifted force {\sc sf} techniques.\cite{Fennell06} These techniques
58     were developed from the observations and efforts of Wolf {\it et al.}
59     and their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
60     The potential for the damped form of the {\sc sp} method is given by
61     \begin{equation}
62     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
63     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
64     \quad r\leqslant R_\textrm{c},
65     \label{eq:DSPPot}
66     \end{equation}
67     while the damped form of the {\sc sf} method is given by
68     \begin{equation}
69     \begin{split}
70     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
71     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
72     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
73     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
74     + \frac{2\alpha}{\pi^{1/2}}
75     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
76     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
77     \quad r\leqslant R_\textrm{c}.
78     \label{eq:DSFPot}
79     \end{split}
80     \end{equation}
81     Overall, the damped {\sc sf} method is considered an improvement over
82     the {\sc sp} method because there is no discontinuity in the forces as
83     particles move out of the cutoff radius ($R_\textrm{c}$). This is
84     accomplished by shifting the forces (and potential) to zero at
85     $R_\textrm{c}$. This is analogous to neutralizing the charge and the
86     force effect of the charges within $R_\textrm{c}$.
87    
88     As we mentioned in the previous study, to complete the charge
89     neutralization procedure, a self-neutralization term needs to be
90     included in the potential. This term exists outside the pair-loop, and
91     can be thought of as a charge opposite in sign and equal in magnitude
92     to the central particle, spread out over the entire cutoff
93     sphere. This term is calculated via a single loop over all charges in
94     the system. For the undamped case, the self-term can be written as
95     \begin{equation}
96     V_\textrm{self} = \sum_{i=1}^N\frac{q_iq_i}{2R_\textrm{c}},
97     \label{eq:selfTerm}
98     \end{equation}
99     while for the damped case it can be written as
100     \begin{equation}
101     V_\textrm{Dself} = \sum_{i=1}^Nq_iq_i\left(\frac{\alpha}{\sqrt{\pi}}
102     + \frac{\textrm{erfc}(\alpha R_\textrm{c})}{2R_\textrm{c}}\right).
103     \label{eq:dampSelfTerm}
104     \end{equation}
105     The first term within the parentheses of equation
106     (\ref{eq:dampSelfTerm}) is identical to the self-term in the Ewald
107     summation, and comes from the utilization of the complimentary error
108     function for electrostatic damping.\cite{deLeeuw80,Wolf99}
109    
110     \section{Multipolar Electrostatic Damping}\label{sec:dampingMultipoles}
111    
112     As stated above, the {\sc sf} method operates by neutralizing the
113     charge pair force interactions (and potential) within the cutoff
114     sphere with shifting and by damping the electrostatic
115     interactions. Now we would like to consider an extension of these
116     techniques to include point multipole interactions. How will the
117     shifting and damping need to be modified in order to accommodate point
118     multipoles?
119    
120     Of the two component techniques, the easiest to adapt is
121     shifting. Shifting is employed to neutralize the cutoff sphere;
122     however, in a system composed purely of point multipoles, the cutoff
123     sphere is already neutralized. This means that shifting is not
124     necessary between point multipoles. In a mixed system of monopoles and
125     multipoles, the undamped {\sc sf} potential needs only to shift the
126     force terms of the monopole and smoothly truncate the multipole
127     interactions with a switching function. The switching function is
128     required in order to conserve energy, because a discontinuity will
129     exist in both the potential and forces at $R_\textrm{c}$ in the
130     absence of shifting terms.
131    
132     If we want to dampen the {\sc sf} potential, then we need to
133     incorporate the complimentary error function term into the multipole
134     potentials. The most direct way to do this is by replacing $r^{-1}$
135     with erfc$(\alpha r)\cdot r^{-1}$ in the multipole expansion. In the
136     multipole expansion, rather than considering only the interactions
137     between single point charges, the electrostatic interaction is
138     reformulated such that it describes the interaction between charge
139     distributions about central sites of the respective sets of
140     charges.\cite{Hirschfelder67} This procedure is what leads to the
141     familiar charge-dipole,
142     \begin{equation}
143     V_\textrm{cd} = -q_i\frac{\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}{r^3_{ij}}
144     = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}},
145     \label{eq:chargeDipole}
146     \end{equation}
147     and dipole-dipole,
148     \begin{equation}
149     V_\textrm{dd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
150     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} -
151     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}},
152     \label{eq:dipoleDipole}
153     \end{equation}
154     interaction potentials.
155    
156     Using the charge-dipole interaction as an example, if we insert
157     erfc$(\alpha r)\cdot r^{-1}$ in place of $r^{-1}$, a damped
158     charge-dipole results,
159     \begin{equation}
160     V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
161     \label{eq:dChargeDipole}
162     \end{equation}
163     where $c_1(r_{ij})$ is
164     \begin{equation}
165     c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
166     + \textrm{erfc}(\alpha r_{ij}).
167     \label{eq:c1Func}
168     \end{equation}
169     Thus, $c_1(r_{ij})$ is the resulting damping term that modifies the
170     standard charge-dipole potential (Eq. (\ref{eq:chargeDipole})). Note
171     that this damping term is dependent upon distance and not upon
172     orientation, and that it is acting on what was originally an
173     $r^{-3}$ function. By writing the damped form in this manner, we
174     can collect the damping into one function and apply it to the original
175     potential when damping is desired. This works well for potentials that
176     have only one $r^{-n}$ term (where $n$ is an odd positive integer);
177     but in the case of the dipole-dipole potential, there is one part
178     dependent on $r^{-3}$ and another dependent on $r^{-5}$. In order to
179     properly damping this potential, each of these parts is dampened with
180     separate damping functions. We can determine the necessary damping
181     functions by continuing with the multipole expansion; however, it
182     quickly becomes more complex with ``two-center'' systems, like the
183     dipole-dipole potential, and is typically approached with a spherical
184     harmonic formalism.\cite{Hirschfelder67} A simpler method for
185     determining these functions arises from adopting the tensor formalism
186     for expressing the electrostatic interactions.\cite{Stone02}
187    
188     The tensor formalism for electrostatic interactions involves obtaining
189     the multipole interactions from successive gradients of the monopole
190     potential. Thus, tensors of rank one through three are
191     \begin{equation}
192     T = \frac{1}{4\pi\epsilon_0r_{ij}},
193     \label{eq:tensorRank1}
194     \end{equation}
195     \begin{equation}
196     T_\alpha = \frac{1}{4\pi\epsilon_0}\nabla_\alpha \frac{1}{r_{ij}},
197     \label{eq:tensorRank2}
198     \end{equation}
199     \begin{equation}
200     T_{\alpha\beta} = \frac{1}{4\pi\epsilon_0}
201     \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
202     \label{eq:tensorRank3}
203     \end{equation}
204     where the form of the first tensor gives the monopole-monopole
205     potential, the second gives the monopole-dipole potential, and the
206     third gives the monopole-quadrupole and dipole-dipole
207     potentials.\cite{Stone02} Since the force is $-\nabla V$, the forces
208     for each potential come from the next higher tensor.
209    
210     To obtain the damped electrostatic forms, we replace $r^{-1}$ with
211     erfc$(\alpha r)\cdot r^{-1}$. Equation \ref{eq:tensorRank2} generates
212     $c_1(r_{ij})$, just like the multipole expansion, while equation
213     \ref{eq:tensorRank3} generates $c_2(r_{ij})$, where
214     \begin{equation}
215     c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
216     + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
217     + \textrm{erfc}(\alpha r_{ij}).
218     \end{equation}
219     Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
220     term. Continuing with higher rank tensors, we can obtain the damping
221     functions for higher multipole potentials and forces. Each subsequent
222     damping function includes one additional term, and we can simplify the
223     procedure for obtaining these terms by writing out the following
224     generating function,
225     \begin{equation}
226     c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
227     {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
228     \label{eq:dampingGeneratingFunc}
229     \end{equation}
230     where,
231     \begin{equation}
232     m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
233     m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
234     m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
235     1 & m = -1\textrm{ or }0,
236     \end{array}\right.
237     \label{eq:doubleFactorial}
238     \end{equation}
239     and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
240     is similar in form to those obtained by researchers for the
241     application of the Ewald sum to
242     multipoles.\cite{Smith82,Smith98,Aguado03}
243    
244     Returning to the dipole-dipole example, the potential consists of a
245     portion dependent upon $r^{-5}$ and another dependent upon
246     $r^{-3}$. In the damped dipole-dipole potential,
247     \begin{equation}
248     V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
249     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
250     c_2(r_{ij}) -
251     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
252     c_1(r_{ij}),
253     \label{eq:dampDipoleDipole}
254     \end{equation}
255     $c_2(r_{ij})$ and $c_1(r_{ij})$ dampen these two parts
256     respectively. The forces for the damped dipole-dipole interaction,
257     \begin{equation}
258     \begin{split}
259     F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
260     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
261     c_3(r_{ij})\\ &-
262     3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
263     \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
264     \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
265     {r^5_{ij}} c_2(r_{ij}),
266     \end{split}
267     \label{eq:dampDipoleDipoleForces}
268     \end{equation}
269     rely on higher order damping functions because we perform another
270     gradient operation. In this manner, we can dampen higher order
271     multipolar interactions along with the monopole interactions, allowing
272     us to include multipoles in simulations involving damped electrostatic
273     interactions.
274    
275     \section{Applications of Pairwise Electrostatic Corrections: Water}
276    
277     In our earlier work, we performed a statistical comparison of both
278     energetics and dynamics of the {\sc sf} and {\sc sp} methods, showing
279     that these pairwise correction techniques are nearly equivalent to the
280     Ewald summation in the simulation of general condensed
281     systems.\cite{Fennell06} It would be beneficial to have some specific
282     examples to better illustrate the similarities and differences of the
283     pairwise (specifically {\sc sf}) and Ewald techniques. To address
284     this, we decided to perform a detailed analysis involving water
285     simulations. Water is a good test in that there is a great deal of
286     detailed experimental information as well as an extensive library of
287     results from a variety of theoretical models. In choosing a model for
288     comparing these correction techniques, it is important to consider a
289     model that has been optimized for use with corrected electrostatics so
290     that the calculated properties are in better agreement with
291     experiment.\cite{vanderSpoel98,Horn04,Rick04} For this reason, we
292     chose the TIP5P-E water model.\cite{Rick04}
293    
294     \subsection{TIP5P-E}
295    
296     The TIP5P-E water model is a variant of Mahoney and Jorgensen's
297     five-point transferable intermolecular potential (TIP5P) model for
298     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
299     maximum anomaly present in liquid water near 4$^\circ$C. As with many
300     previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
301     and SPC/E), TIP5P was parametrized using a simple cutoff with no
302     long-range electrostatic
303     correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
304     Without this correction, the pressure term on the central particle
305     from the surroundings is missing. When this correction is included,
306     systems of these particles expand to compensate for this added
307     pressure term and under-predict the density of water under standard
308     conditions. When using any form of long-range electrostatic
309     correction, it has become common practice to develop or utilize a
310     reparametrized water model that corrects for this
311     effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
312     this practice and was optimized for use with the Ewald
313     summation.\cite{Rick04} In his publication, Rick preserved the
314     geometry and point charge magnitudes in TIP5P and focused on altering
315     the Lennard-Jones parameters to correct the density at 298~K. With the
316     density corrected, he compared common water properties for TIP5P-E
317     using the Ewald sum with TIP5P using a 9~\AA\ cutoff.
318    
319     In the following sections, we compared these same water properties
320     calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
321     {\sc sf} technique. In the above evaluation of the pairwise
322     techniques, we observed some flexibility in the choice of parameters.
323     Because of this, the following comparisons include the {\sc sf}
324     technique with a 12~\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
325     0.2~\AA$^{-1}$, as well as a 9~\AA\ cutoff with an $\alpha$ =
326     0.2~\AA$^{-1}$.
327    
328     \subsubsection{Density}\label{sec:t5peDensity}
329    
330     As stated previously, the property that prompted the development of
331     TIP5P-E was the density at 1 atm. The density depends upon the
332     internal pressure of the system in the $NPT$ ensemble, and the
333     calculation of the pressure includes a components from both the
334     kinetic energy and the virial. More specifically, the instantaneous
335     molecular pressure ($p(t)$) is given by
336     \begin{equation}
337     p(t) = \frac{1}{\textrm{d}V}\sum_\mu
338     \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
339     + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
340     \label{eq:MolecularPressure}
341     \end{equation}
342     where d is the dimensionality of the system, $V$ is the volume,
343     $\mathbf{P}_{\mu}$ is the momentum of molecule $\mu$, $\mathbf{R}_\mu$
344     is the position of the center of mass ($M_\mu$) of molecule $\mu$, and
345     $\mathbf{F}_{\mu i}$ is the force on atom $i$ of molecule
346     $\mu$.\cite{Melchionna93} The virial term (the right term in the
347     brackets of equation \ref{eq:MolecularPressure}) is directly dependent
348     on the interatomic forces. Since the {\sc sp} method does not modify
349     the forces, the pressure using {\sc sp} will be identical to that
350     obtained without an electrostatic correction. The {\sc sf} method
351     does alter the virial component and, by way of the modified pressures,
352     should provide densities more in line with those obtained using the
353     Ewald summation.
354    
355     To compare densities, $NPT$ simulations were performed with the same
356     temperatures as those selected by Rick in his Ewald summation
357     simulations.\cite{Rick04} In order to improve statistics around the
358     density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
359     25$^\circ$C, while 2~ns trajectories were obtained at all other
360     temperatures. The average densities were calculated from the later
361     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
362     method for accumulating statistics, these sequences were spliced into
363     200 segments, each providing an average density. These 200 density
364     values were used to calculate the average and standard deviation of
365     the density at each temperature.\cite{Mahoney00}
366    
367     \begin{figure}
368     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
369     \caption{Density versus temperature for the TIP5P-E water model when
370     using the Ewald summation (Ref. \citen{Rick04}) and the {\sc sf} method
371     with various parameters. The pressure term from the image-charge shell
372     is larger than that provided by the reciprocal-space portion of the
373     Ewald summation, leading to slightly lower densities. This effect is
374     more visible with the 9~\AA\ cutoff, where the image charges exert a
375     greater force on the central particle. The error bars for the {\sc sf}
376     methods show the average one-sigma uncertainty of the density
377     measurement, and this uncertainty is the same for all the {\sc sf}
378     curves.}
379     \label{fig:t5peDensities}
380     \end{figure}
381     Figure \ref{fig:t5peDensities} shows the densities calculated for
382     TIP5P-E using differing electrostatic corrections overlaid on the
383     experimental values.\cite{CRC80} The densities when using the {\sc sf}
384     technique are close to, though typically lower than, those calculated
385     using the Ewald summation. These slightly reduced densities indicate
386     that the pressure component from the image charges at R$_\textrm{c}$
387     is larger than that exerted by the reciprocal-space portion of the
388     Ewald summation. Bringing the image charges closer to the central
389     particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than the
390     preferred 12~\AA\ R$_\textrm{c}$) increases the strength of the image
391     charge interactions on the central particle and results in a further
392     reduction of the densities.
393    
394     Because the strength of the image charge interactions has a noticeable
395     effect on the density, we would expect the use of electrostatic
396     damping to also play a role in these calculations. Larger values of
397     $\alpha$ weaken the pair-interactions; and since electrostatic damping
398     is distance-dependent, force components from the image charges will be
399     reduced more than those from particles close the the central
400     charge. This effect is visible in figure \ref{fig:t5peDensities} with
401     the damped {\sc sf} sums showing slightly higher densities; however,
402     it is apparent that the choice of cutoff radius plays a much more
403     important role in the resulting densities.
404    
405     As a final note, all of the above density calculations were performed
406     with systems of 512 water molecules. Rick observed a system size
407     dependence of the computed densities when using the Ewald summation,
408     most likely due to his tying of the convergence parameter to the box
409     dimensions.\cite{Rick04} For systems of 256 water molecules, the
410     calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
411     system size of 256 molecules would force the use of a shorter
412     R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
413     lower the densities. Moving to larger systems, as long as the
414     R$_\textrm{c}$ remains at a fixed value, we would expect the densities
415     to remain constant.
416    
417     \subsubsection{Liquid Structure}\label{sec:t5peLiqStructure}
418    
419     A common function considered when developing and comparing water
420     models is the oxygen-oxygen radial distribution function
421     ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
422     finding a pair of oxygen atoms some distance ($r$) apart relative to a
423     random distribution at the same density.\cite{Allen87} It is
424     calculated via
425     \begin{equation}
426     g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
427     \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
428     \label{eq:GOOofR}
429     \end{equation}
430     where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
431     atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
432     neutron scattering experiments through the oxygen-oxygen structure
433     factor ($S_\textrm{OO}(k)$) by the following relationship:
434     \begin{equation}
435     S_\textrm{OO}(k) = 1 + 4\pi\rho
436     \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
437     \label{eq:SOOofK}
438     \end{equation}
439     Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
440     of $g_\textrm{OO}(r)$.
441    
442     The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
443     been compared in great detail with the various common water models,
444     and TIP5P was found to be in better agreement than other rigid,
445     non-polarizable models.\cite{Sorenson00} This excellent agreement with
446     experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
447     check whether the choice of using the Ewald summation or the {\sc sf}
448     technique alters the liquid structure, the $g_\textrm{OO}(r)$s at
449     298~K and 1~atm were determined for the systems compared in the
450     previous section.
451    
452     \begin{figure}
453     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
454     \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and
455     1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
456     sf} technique with varying parameters. Even with the reduced densities
457     using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
458     identical.}
459     \label{fig:t5peGofRs}
460     \end{figure}
461     The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
462     sf} technique with a various parameters are overlaid on the
463     $g_\textrm{OO}(r)$ while using the Ewald summation in
464     figure~\ref{fig:t5peGofRs}. The differences in density do not appear
465     to have any effect on the liquid structure as the $g_\textrm{OO}(r)$s
466     are indistinguishable. These results indicate that the
467     $g_\textrm{OO}(r)$ is insensitive to the choice of electrostatic
468     correction.
469    
470     \subsubsection{Thermodynamic Properties}\label{sec:t5peThermo}
471    
472     In addition to the density, there are a variety of thermodynamic
473     quantities that can be calculated for water and compared directly to
474     experimental values. Some of these additional quantities include the
475     latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
476     pressure heat capacity ($C_p$), the isothermal compressibility
477     ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
478     dielectric constant ($\epsilon$). All of these properties were
479     calculated for TIP5P-E with the Ewald summation, so they provide a
480     good set for comparisons involving the {\sc sf} technique.
481    
482     The $\Delta H_\textrm{vap}$ is the enthalpy change required to
483     transform one mole of substance from the liquid phase to the gas
484     phase.\cite{Berry00} In molecular simulations, this quantity can be
485     determined via
486     \begin{equation}
487     \begin{split}
488     \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
489     &= E_\textrm{gas} - E_\textrm{liq.}
490     + p(V_\textrm{gas} - V_\textrm{liq.}) \\
491     &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
492     \end{split}
493     \label{eq:DeltaHVap}
494     \end{equation}
495     where $E$ is the total energy, $U$ is the potential energy, $p$ is the
496     pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
497     the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
498     seen in the last line of equation (\ref{eq:DeltaHVap}), we can
499     approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
500     state. This allows us to cancel the kinetic energy terms, leaving only
501     the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
502     several orders of magnitude larger than that of the liquid, so we can
503     neglect the liquid $pV$ term.
504    
505     The remaining thermodynamic properties can all be calculated from
506     fluctuations of the enthalpy, volume, and system dipole
507     moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
508     enthalpy in constant pressure simulations via
509     \begin{equation}
510     \begin{split}
511     C_p = \left(\frac{\partial H}{\partial T}\right)_{N,p}
512     = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
513     \end{split}
514     \label{eq:Cp}
515     \end{equation}
516     where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
517     \begin{equation}
518     \begin{split}
519     \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
520     = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
521     {k_BT\langle V\rangle_{N,P,T}},
522     \end{split}
523     \label{eq:kappa}
524     \end{equation}
525     and $\alpha_p$ can be calculated via
526     \begin{equation}
527     \begin{split}
528     \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
529     = \frac{(\langle VH\rangle_{N,P,T}
530     - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
531     {k_BT^2\langle V\rangle_{N,P,T}}.
532     \end{split}
533     \label{eq:alpha}
534     \end{equation}
535     Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
536     be calculated for systems of non-polarizable substances via
537     \begin{equation}
538     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
539     \label{eq:staticDielectric}
540     \end{equation}
541     where $\epsilon_0$ is the permittivity of free space and $\langle
542     M^2\rangle$ is the fluctuation of the system dipole
543     moment.\cite{Allen87} The numerator in the fractional term in equation
544     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
545     dipole moment, identical to the quantity calculated in the
546     finite-system Kirkwood $g$ factor ($G_k$):
547     \begin{equation}
548     G_k = \frac{\langle M^2\rangle}{N\mu^2},
549     \label{eq:KirkwoodFactor}
550     \end{equation}
551     where $\mu$ is the dipole moment of a single molecule of the
552     homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
553     fluctuation term in both equation (\ref{eq:staticDielectric}) and
554     \ref{eq:KirkwoodFactor} is calculated as follows,
555     \begin{equation}
556     \begin{split}
557     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
558     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
559     &= \langle M_x^2+M_y^2+M_z^2\rangle
560     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
561     + \langle M_x\rangle^2).
562     \end{split}
563     \label{eq:fluctBoxDipole}
564     \end{equation}
565     This fluctuation term can be accumulated during the simulation;
566     however, it converges rather slowly, thus requiring multi-nanosecond
567     simulation times.\cite{Horn04} In the case of tin-foil boundary
568     conditions, the dielectric/surface term of the Ewald summation is
569     equal to zero. Since the {\sc sf} method also lacks this
570     dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
571     valid for determining static dielectric constants.
572    
573     All of the above properties were calculated from the same trajectories
574     used to determine the densities in section \ref{sec:t5peDensity}
575     except for the static dielectric constants. The $\epsilon$ values were
576     accumulated from 2~ns $NVE$ ensemble trajectories with system densities
577     fixed at the average values from the $NPT$ simulations at each of the
578     temperatures. The resulting values are displayed in figure
579     \ref{fig:t5peThermo}.
580     \begin{figure}
581     \centering
582     \includegraphics[width=4.5in]{./figures/t5peThermo.pdf}
583     \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
584     and the {\sc sf} techniques along with the experimental values. Units
585     for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
586     cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
587     and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
588     reference \cite{Rick04}. Experimental values for $\Delta
589     H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
590     \cite{Kell75}. Experimental values for $C_p$ are from reference
591     \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
592     \cite{Malmberg56}.}
593     \label{fig:t5peThermo}
594     \end{figure}
595    
596     As observed for the density in section \ref{sec:t5peDensity}, the
597     property trends with temperature seen when using the Ewald summation
598     are reproduced with the {\sc sf} technique. One noticeable difference
599     between the properties calculated using the two methods are the lower
600     $\Delta H_\textrm{vap}$ values when using {\sc sf}. This is to be
601     expected due to the direct weakening of the electrostatic interaction
602     through forced neutralization. This results in an increase of the
603     intermolecular potential producing lower values from equation
604     (\ref{eq:DeltaHVap}). The slopes of these values with temperature are
605     similar to that seen using the Ewald summation; however, they are both
606     steeper than the experimental trend, indirectly resulting in the
607     inflated $C_p$ values at all temperatures.
608    
609     Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$ values
610     all overlap within error. As indicated for the $\Delta H_\textrm{vap}$
611     and $C_p$ results discussed in the previous paragraph, the deviations
612     between experiment and simulation in this region are not the fault of
613     the electrostatic summation methods but are due to the geometry and
614     parameters of the TIP5P class of water models. Like most rigid,
615     non-polarizable, point-charge water models, the density decreases with
616     temperature at a much faster rate than experiment (see figure
617     \ref{fig:t5peDensities}). This reduced density leads to the inflated
618     compressibility and expansivity values at higher temperatures seen
619     here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
620     and many-body effects are required in order for water models to
621     overcome differences between simulation-based and experimentally
622     determined densities at these higher
623     temperatures.\cite{Laasonen93,Donchev06}
624    
625     At temperatures below the freezing point for experimental water, the
626     differences between {\sc sf} and the Ewald summation results are more
627     apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
628     indicate a more pronounced transition in the supercooled regime,
629     particularly in the case of {\sc sf} without damping. This points to
630     the onset of a more frustrated or glassy behavior for TIP5P-E at
631     temperatures below 250~K in the {\sc sf} simulations, indicating that
632     disorder in the reciprocal-space term of the Ewald summation might act
633     to loosen up the local structure more than the image-charges in {\sc
634     sf}. The damped {\sc sf} actually makes a better comparison with
635     experiment in this region, particularly for the $\alpha_p$ values. The
636     local interactions in the undamped {\sc sf} technique appear to be too
637     strong since the property change is much more dramatic than the damped
638     forms, while the Ewald summation appears to weight the
639     reciprocal-space interactions at the expense the local interactions,
640     disagreeing with the experimental results. This observation is
641     explored further in section \ref{sec:t5peDynamics}.
642    
643     The final thermodynamic property displayed in figure
644     \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
645     between the Ewald summation and the {\sc sf} technique (and experiment
646     for that matter). It is known that the dielectric constant is
647     dependent upon and quite sensitive to the imposed boundary
648     conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
649     converged $\epsilon$ values accumulated for the {\sc sf}
650     simulations. Lack of a damping function results in dielectric
651     constants significantly smaller than those obtained using the Ewald
652     sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
653     agreement considerably. It should be noted that the choice of the
654     ``Ewald coefficient'' value also has a significant effect on the
655     calculated value when using the Ewald summation. In the simulations of
656     TIP5P-E with the Ewald sum, this screening parameter was tethered to
657     the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04} In
658     general, systems with larger screening parameters reported larger
659     dielectric constant values, the same behavior we see here with {\sc
660     sf}; however, the choice of cutoff radius also plays an important
661     role. This connection is further explored below as optimal damping
662     coefficients for different choices of $R_\textrm{c}$ are determined
663     for {\sc sf} in order to best capture the dielectric behavior.
664    
665     \subsubsection{Dielectric Constant}\label{sec:t5peDielectric}
666    
667     In the previous, we observed that the choice of damping coefficient
668     plays a major role in the calculated dielectric constant. This is not
669     too surprising given the results for damping parameter influence on
670     the long-time correlated motions of the NaCl crystal in our previous
671     study.\cite{Fennell06} The static dielectric constant is calculated
672     from the long-time fluctuations of the system's accumulated dipole
673     moment (Eq. (\ref{eq:staticDielectric})), so it is going to be quite
674     sensitive to the choice of damping parameter. We would like to choose
675     optimal damping constants such that any arbitrary choice of cutoff
676     radius will properly capture the dielectric behavior of the liquid.
677    
678     In order to find these optimal values, we mapped out the static
679     dielectric constant as a function of both the damping parameter and
680     cutoff radius. To calculate the static dielectric constant, we
681     performed 5~ns $NPT$ calculations on systems of 512 TIP5P-E water
682     molecules and averaged over the converged region (typically the later
683     2.5~ns) of equation (\ref{eq:staticDielectric}). The selected cutoff
684     radii ranged from 9, 10, 11, and 12~\AA , and the damping parameter
685     values ranged from 0.1 to 0.35~\AA$^{-1}$.
686    
687     \begin{table}
688     \centering
689     \caption{TIP5P-E dielectric constant as a function of $\alpha$ and
690     $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
691     \vspace{6pt}
692     \begin{tabular}{ lcccc }
693     \toprule
694     \toprule
695     & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
696     \cmidrule(lr){2-5}
697     $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 \\
698     \midrule
699     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 \\
700     & \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} \\
701     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 \\
702     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 \\
703     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 \\
704     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 \\
705     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 \\
706     & \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} \\
707     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 \\
708     & \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} \\
709     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 \\
710     \bottomrule
711     \end{tabular}
712     \label{tab:tip5peDielectricMap}
713     \end{table}
714     The results of these calculations are displayed in table
715     \ref{tab:tip5peDielectricMap}. Coloring of the table is based on the
716     numerical values within the cells and was added to better facilitate
717     interpretation of the displayed data and highlight trends in the
718     dielectric constant behavior. This makes it easy to see that the
719     dielectric constant is linear with respect to $\alpha$ and
720     $R_\textrm{c}$ in the low to moderate damping regions, and the slope
721     is 0.025~\AA$^{-1}\cdot R_\textrm{c}^{-1}$. Another point to note is
722     that choosing $\alpha$ and $R_\textrm{c}$ identical to those used with
723     the Ewald summation results in the same calculated dielectric
724     constant. As an example, in the paper outlining the development of
725     TIP5P-E, the real-space cutoff and Ewald coefficient were tethered to
726     the system size, and for a 512 molecule system are approximately
727     12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04} These
728     parameters resulted in a dielectric constant of 92$\pm$14, while with
729     {\sc sf} these parameters give a dielectric constant of
730     90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
731     $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
732     0.35~\AA$^{-1}$, and these parameters resulted in a dielectric
733     constant equal to 63$\pm$1.\cite{Horn04} Calculations using {\sc sf}
734     with these parameters and this water model give a dielectric constant
735     of 61$\pm$1. Since the dielectric constant is dependent on $\alpha$
736     and $R_\textrm{c}$ with the {\sc sf} technique, it might be
737     interesting to investigate the dielectric dependence of the real-space
738     Ewald parameters.
739    
740     Although it is tempting to choose damping parameters equivalent to
741     these Ewald examples, the results of our previous work indicate that
742     values this high are destructive to both the energetics and
743     dynamics.\cite{Fennell06} Ideally, $\alpha$ should not exceed
744     0.3~\AA$^{-1}$ for any of the cutoff values in this range. If the
745     optimal damping parameter is chosen to be midway between 0.275 and
746     0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff, then the
747     linear trend with $R_\textrm{c}$ will always keep $\alpha$ below
748     0.3~\AA$^{-1}$ for the studied cutoff radii. This linear progression
749     would give values of 0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for
750     cutoff radii of 9, 10, 11, and 12~\AA. Setting this to be the default
751     behavior for the damped {\sc sf} technique will result in consistent
752     dielectric behavior for these and other condensed molecular systems,
753     regardless of the chosen cutoff radius. The static dielectric
754     constants for TIP5P-E simulations with 9 and 12\AA\ $R_\textrm{c}$
755     values using their respective damping parameters are 76$\pm$1 and
756     75$\pm$2. These values are lower than the values reported for TIP5P-E
757     with the Ewald sum; however, they are more in line with the values
758     reported by Mahoney {\it et al.} for TIP5P while using a reaction
759     field (RF) with an infinite RF dielectric constant
760     (81.5$\pm$1.6).\cite{Mahoney00} As a final note, aside from a slight
761     lowering of $\Delta H_\textrm{vap}$, using these $\alpha$ values does
762     not alter the other other thermodynamic properties.
763    
764     \subsubsection{Dynamic Properties}\label{sec:t5peDynamics}
765    
766     To look at the dynamic properties of TIP5P-E when using the {\sc sf}
767     method, 200~ps $NVE$ simulations were performed for each temperature
768     at the average density reported by the $NPT$ simulations using 9 and
769     12~\AA\ $R_\textrm{c}$ values using the ideal $\alpha$ values
770     determined above (0.2875 and 0.2125~\AA$^{-1}$). The self-diffusion
771     constants ($D$) were calculated using the mean square displacement
772     (MSD) form of the Einstein relation,
773     \begin{equation}
774     D = \lim_{t\rightarrow\infty}
775     \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
776     \label{eq:MSD}
777     \end{equation}
778     where $t$ is time, and $\mathbf{r}_i$ is the position of particle
779     $i$.\cite{Allen87}
780    
781     \begin{figure}
782     \centering
783     \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
784     \caption{Body-fixed coordinate frame for a water molecule. The
785     respective molecular principle axes point in the direction of the
786     labeled frame axes.}
787     \label{fig:waterFrame}
788     \end{figure}
789     In addition to translational diffusion, orientational relaxation times
790     were calculated for comparisons with the Ewald simulations and with
791     experiments. These values were determined from the same 200~ps $NVE$
792     trajectories used for translational diffusion by calculating the
793     orientational time correlation function,
794     \begin{equation}
795     C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
796     \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
797     \label{eq:OrientCorr}
798     \end{equation}
799     where $P_l$ is the Legendre polynomial of order $l$ and
800     $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
801     principle axis $\alpha$. The principle axis frame for these water
802     molecules is shown in figure \ref{fig:waterFrame}. As an example,
803     $C_l^y$ is calculated from the time evolution of the unit vector
804     connecting the two hydrogen atoms.
805    
806     \begin{figure}
807     \centering
808     \includegraphics[width=3.5in]{./figures/exampleOrientCorr.pdf}
809     \caption{Example plots of the orientational autocorrelation functions
810     for the first and second Legendre polynomials. These curves show the
811     time decay of the unit vector along the $y$ principle axis.}
812     \label{fig:OrientCorr}
813     \end{figure}
814     From the orientation autocorrelation functions, we can obtain time
815     constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
816     some example plots of orientational autocorrelation functions for the
817     first and second Legendre polynomials. The relatively short time
818     portions (between 1 and 3~ps for water) of these curves can be fit to
819     an exponential decay to obtain these constants, and they are directly
820     comparable to water orientational relaxation times from nuclear
821     magnetic resonance (NMR). The relaxation constant obtained from
822     $C_2^y(t)$ is of particular interest because it describes the
823     relaxation of the principle axis connecting the hydrogen atoms. Thus,
824     $C_2^y(t)$ can be compared to the intermolecular portion of the
825     dipole-dipole relaxation from a proton NMR signal and should provide
826     the best estimate of the NMR relaxation time constant.\cite{Impey82}
827    
828     \begin{figure}
829     \centering
830     \includegraphics[width=3.5in]{./figures/t5peDynamics.pdf}
831     \caption{Diffusion constants ({\it upper}) and reorientational time
832     constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
833     technique compared with experiment. Data at temperatures less than
834     0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
835     easier comparisons in the more relevant temperature regime.}
836     \label{fig:t5peDynamics}
837     \end{figure}
838     Results for the diffusion constants and orientational relaxation times
839     are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
840     apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
841     the Ewald sum are reproduced with the {\sc sf} technique. The enhanced
842     diffusion at high temperatures are again the product of the lower
843     densities in comparison with experiment and do not provide any special
844     insight into differences between the electrostatic summation
845     techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
846     diffuse a little faster than with the Ewald sum; however, use of light
847     to moderate damping results in indistinguishable $D$ values. Though
848     not apparent in this figure, {\sc sf} values at the lowest temperature
849     are approximately twice as slow as $D$ values obtained using the Ewald
850     sum. These values support the observation from section
851     \ref{sec:t5peThermo} that there appeared to be a change to a more
852     glassy-like phase with the {\sc sf} technique at these lower
853     temperatures, though this change seems to be more prominent with the
854     {\it undamped} {\sc sf} method, which has stronger local pairwise
855     electrostatic interactions.
856    
857     The $\tau_2^y$ results in the lower frame of figure
858     \ref{fig:t5peDynamics} show a much greater difference between the {\sc
859     sf} results and the Ewald results. At all temperatures shown, TIP5P-E
860     relaxes faster than experiment with the Ewald sum while tracking
861     experiment fairly well when using the {\sc sf} technique, independent
862     of the choice of damping constant. Their are several possible reasons
863     for this deviation between techniques. The Ewald results were
864     calculated using shorter (10ps) trajectories than the {\sc sf} results
865     (25ps). A quick calculation from a 10~ps trajectory with {\sc sf} with
866     an $\alpha$ of 0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in
867     $\tau_2^y$, placing the result more in line with that obtained using
868     the Ewald sum. This example supports this explanation; however,
869     recomputing the results to meet a poorer statistical standard is
870     counter-productive. Assuming the Ewald results are not entirely the
871     product of poor statistics, differences in techniques to integrate the
872     orientational motion could also play a role. {\sc shake} is the most
873     commonly used technique for approximating rigid-body orientational
874     motion,\cite{Ryckaert77} whereas in {\sc oopse}, we maintain and
875     integrate the entire rotation matrix using the {\sc dlm}
876     method.\cite{Meineke05} Since {\sc shake} is an iterative constraint
877     technique, if the convergence tolerances are raised for increased
878     performance, error will accumulate in the orientational
879     motion. Finally, the Ewald results were calculated using the $NVT$
880     ensemble, while the $NVE$ ensemble was used for {\sc sf}
881     calculations. The additional mode of motion due to the thermostat will
882     alter the dynamics, resulting in differences between $NVT$ and $NVE$
883     results. These differences are increasingly noticeable as the
884     thermostat time constant decreases.
885    
886    
887     \subsection{SSD/RF}
888    
889     In section \ref{sec:dampingMultipoles}, we described a method for
890     applying the damped {\sc sf} technique to systems containing point
891     multipoles. The soft sticky dipole (SSD) family of water models is the
892     perfect test case because of the dipole-dipole (and
893     charge-dipole/quadrupole) interactions that are present. As alluded to
894     in the name, soft sticky dipole water models are single point
895     molecules that consist of a ``soft'' Lennard-Jones sphere, a
896     point-dipole, and a tetrahedral function for capturing the hydrogen
897     bonding nature of water - a spherical harmonic term for water-water
898     tetrahedral interactions and a point-quadrupole for interactions with
899     surrounding charges. Detailed descriptions of these models can be
900     found in other studies.\cite{Liu96b,Chandra99,Tan03,Fennell04}
901    
902     In deciding which version of the SSD model to use, we need only
903     consider that the {\sc sf} technique was presented as a pairwise
904     replacement for the Ewald summation. It has been suggested that models
905     parametrized for the Ewald summation (like TIP5P-E) would be
906     appropriate for use with a reaction field and vice versa.\cite{Rick04}
907     Therefore, we decided to test the SSD/RF water model, which was
908     parametrized for use with a reaction field, with this damped
909     electrostatic technique to see how the calculated properties change.
910    
911     \subsubsection{Dipolar Damping}
912    
913     \begin{table}
914     \caption{Properties of SSD/RF when using different electrostatic
915     correction methods.}
916     \footnotesize
917     \centering
918     \begin{tabular}{ llccc }
919     \toprule
920     \toprule
921     & & Reaction Field [Ref. \citen{Fennell04}] & Damped Electrostatics & Experiment [Ref.] \\
922     & & $\epsilon = 80$ & $R_\textrm{c} = 12$\AA ; $\alpha = 0.2125$~\AA$^{-1}$ & \\
923     \midrule
924     $\rho$ & (g cm$^{-3}$) & 0.997(1) & 1.004(4) & 0.997 [\citen{CRC80}]\\
925     $C_p$ & (cal mol$^{-1}$ K$^{-1}$) & 23.8(2) & 27(1) & 18.005 [\citen{Wagner02}] \\
926     $D$ & ($10^{-5}$ cm$^2$ s$^{-1}$) & 2.32(6) & 2.33(2) & 2.299 [\citen{Mills73}]\\
927     $n_C$ & & 4.4 & 4.2 & 4.7 [\citen{Hura00}]\\
928     $n_H$ & & 3.7 & 3.7 & 3.5 [\citen{Soper86}]\\
929     $\tau_1$ & (ps) & 7.2(4) & 5.86(8) & 5.7 [\citen{Eisenberg69}]\\
930     $\tau_2$ & (ps) & 3.2(2) & 2.45(7) & 2.3 [\citen{Krynicki66}]\\
931     \bottomrule
932     \end{tabular}
933     \label{tab:dampedSSDRF}
934     \end{table}
935     The properties shown in table \ref{tab:dampedSSDRF} compare
936     quite well. The average density shows a modest increase when
937     using damped electrostatics in place of the reaction field. This comes
938     about because we neglect the pressure effect due to the surroundings
939     outside of the cutoff, instead relying on screening effects to
940     neutralize electrostatic interactions at long distances. The $C_p$
941     also shows a slight increase, indicating greater fluctuation in the
942     enthalpy at constant pressure. The only other differences between the
943     damped and reaction field results are the dipole reorientational time
944     constants, $\tau_1$ and $\tau_2$. When using damped electrostatics,
945     the water molecules relax more quickly and exhibit behavior very
946     similar to the experimental values. These results indicate that not
947     only is it reasonable to use damped electrostatics with SSD/RF, it is
948     recommended if capturing realistic dynamics is of primary
949     importance. This is an encouraging result because the damping methods
950     are more generally applicable than reaction field. Using damping,
951     SSD/RF can be used effectively with mixed systems, such as dissolved
952     ions, dissolved organic molecules, or even proteins.
953    
954     \subsubsection{Dielectric Constant}
955    
956     \begin{table}
957     \centering
958     \caption{SSD/RF dielectric constant as a function of $\alpha$ and $R_\textrm{c}$. The color scale ranges from blue (10) to red (100).}
959     \vspace{6pt}
960     \begin{tabular}{ lcccc }
961     \toprule
962     \toprule
963     & \multicolumn{4}{c}{$R_\textrm{c}$ (\AA )} \\
964     \cmidrule(lr){2-5}
965     $\alpha$ (\AA$^{-1}$) & 9 & 10 & 11 & 12 \\
966     \midrule
967     0.35 & \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 \\
968     & \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} \\
969     0.3 & \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 \\
970     0.275 & \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 \\
971     0.25 & \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 \\
972     0.225 & \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 \\
973     0.2 & \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 \\
974     & \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} \\
975     0.15 & \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 \\
976     & \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} \\
977     0.1 & \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 \\
978     \bottomrule
979     \end{tabular}
980     \label{tab:ssdrfDielectricMap}
981     \end{table}
982     In addition to the properties tabulated in table
983     \ref{tab:dampedSSDRF}, we mapped out the static dielectric constant of
984     SSD/RF as a function of $R_\textrm{c}$ and $\alpha$ in the same
985     fashion as in section \ref{sec:t5peDielectric} with TIP5P-E. It is
986     apparent from this table that electrostatic damping has a more
987     pronounced effect on the dipolar interactions of SSD/RF than the
988     monopolar interactions of TIP5P-E. The dielectric constant covers a
989     much wider range and has a steeper slope with increasing damping
990     parameter. We can use the same trend of $\alpha$ with $R_\textrm{c}$
991     used with TIP5P-E, and for a 12~\AA\ $R_\textrm{c}$, the resulting
992     dielectric constant is 82.6$\pm$0.6. This value compares very
993     favorably with the experimental value of 78.3.\cite{Malmberg56} This
994     is not surprising given that early studies of the SSD model indicate a
995     static dielectric constant around 81.\cite{Liu96}
996    
997     \section{Application of Pairwise Electrostatic Corrections: Imaginary Ice}
998    
999     In an earlier work, we performed a series of free energy calculations
1000     on several ice polymorphs which are stable or metastable at low
1001     pressures, one of which (Ice-$i$) we observed in spontaneous
1002     crystallizations of an SSD type single point water
1003     model.\cite{Fennell05} In this study, a distinct dependence of the
1004     free energies on the interaction cutoff and correction technique was
1005     observed. Being that the {\sc sf} technique can be used as a simple
1006     and efficient replacement for the Ewald summation, it would be
1007     interesting to apply it in addressing the question of the free
1008     energies of these ice polymorphs with varying water models. To this
1009     end, we set up thermodynamic integrations of all of the previously
1010     discussed ice polymorphs using the {\sc sf} technique with a cutoff
1011     radius of 12~\AA\ and an $\alpha$ of 0.2125~\AA . These calculations
1012     were performed on TIP5P-E and TIP4P-Ew (variants of the root models
1013     optimized for the Ewald summation) as well as SPC/E, and SSD/RF.
1014    
1015     \begin{table}
1016     \centering
1017     \caption{Helmholtz free energies of ice polymorphs at 1~atm and 200~K
1018     using the damped {\sc sf} electrostatic correction method with a
1019     variety of water models.}
1020     \begin{tabular}{ lccccc }
1021     \toprule
1022     \toprule
1023     Model & I$_\textrm{h}$ & I$_\textrm{c}$ & B & Ice-$i$ & Ice-$i^\prime$ \\
1024     \cmidrule(lr){2-6}
1025     & \multicolumn{5}{c}{(kcal mol$^{-1}$)} \\
1026     \midrule
1027     TIP5P-E & -11.98(4) & -11.96(4) & -11.87(3) & - & -11.95(3) \\
1028     TIP4P-Ew & -13.11(3) & -13.09(3) & -12.97(3) & - & -12.98(3) \\
1029     SPC/E & -12.99(3) & -13.00(3) & -13.03(3) & - & -12.99(3) \\
1030     SSD/RF & -11.83(3) & -11.66(4) & -12.32(3) & -12.39(3) & - \\
1031     \bottomrule
1032     \end{tabular}
1033     \label{tab:dampedFreeEnergy}
1034     \end{table}
1035     The results of these calculations in table \ref{tab:dampedFreeEnergy}
1036     show similar behavior to the Ewald results in the previous study, at
1037     least for SSD/RF and SPC/E which are present in both.\cite{Fennell05}
1038     The Helmholtz free energies of the ice polymorphs with SSD/RF order in
1039     the same fashion, with Ice-$i$ having the lowest free energy; however,
1040     the Ice-$i$ and ice B free energies are quite a bit closer (nearly
1041     isoenergetic). The SPC/E results show the near isoenergetic behavior
1042     when using the Ewald summation.\cite{Fennell05} Ice B has the lowest
1043     Helmholtz free energy; however, all the polymorph results overlap
1044     within error.
1045    
1046     The most interesting results from these calculations come from the
1047     more expensive TIP4P-Ew and TIP5P-E results. Both of these models were
1048     optimized for use with an electrostatic correction and are
1049     geometrically arranged to mimic water following two different
1050     ideas. In TIP5P-E, the primary location for the negative charge in the
1051     molecule is assigned to the lone-pairs of the oxygen, while TIP4P-Ew
1052     places the negative charge near the center-of-mass along the H-O-H
1053     bisector. There is some debate as to which is the proper choice for
1054     the negative charge location, and this has in part led to a six-site
1055     water model that balances both of these options.\cite{Vega05,Nada03}
1056     The limited results in table \ref{tab:dampedFreeEnergy} support the
1057     results of Vega {\it et al.}, which indicate the TIP4P charge location
1058     geometry is more physically valid.\cite{Vega05} With the TIP4P-Ew
1059     water model, the experimentally observed polymorph (ice
1060     I$_\textrm{h}$) is the preferred form with ice I$_\textrm{c}$ slightly
1061     higher in energy, though overlapping within error. Additionally, the
1062     less realistic ice B and Ice-$i^\prime$ structures are destabilized
1063     relative to these polymorphs. TIP5P-E shows similar behavior to SPC/E,
1064     where there is no real free energy distinction between the various
1065     polymorphs, because many overlap within error. While ice B is close in
1066     free energy to the other polymorphs, these results fail to support the
1067     findings of other researchers indicating the preferred form of TIP5P
1068     at 1~atm is a structure similar to ice
1069     B.\cite{Yamada02,Vega05,Abascal05} It should be noted that we are
1070     looking at TIP5P-E rather than TIP5P, and the differences in the
1071     Lennard-Jones parameters could be a reason for this dissimilarity.
1072     Overall, these results indicate that TIP4P-Ew is a better mimic of
1073     real water than these other models when studying crystallization and
1074     solid forms of water.
1075    
1076     \section{Conclusions}
1077    
1078     \section{Acknowledgments}
1079     Support for this project was provided by the National Science
1080     Foundation under grant CHE-0134881. Computation time was provided by
1081     the Notre Dame Center for Research Computing.
1082    
1083     \newpage
1084    
1085     \bibliographystyle{achemso}
1086     \bibliography{multipoleSFPaper}
1087    
1088    
1089     \end{document}