ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2619
Committed: Wed Mar 15 00:21:39 2006 UTC (18 years, 6 months ago) by chrisfen
Content type: application/x-tex
File size: 46711 byte(s)
Log Message:
Updated some plots and captions.  Added a Makefile.

File Contents

# User Rev Content
1 chrisfen 2575 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 2617 %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4 chrisfen 2575 \usepackage{endfloat}
5     \usepackage{amsmath}
6 chrisfen 2594 \usepackage{amssymb}
7 chrisfen 2575 \usepackage{epsf}
8     \usepackage{times}
9 gezelter 2617 \usepackage{mathptmx}
10 chrisfen 2575 \usepackage{setspace}
11     \usepackage{tabularx}
12     \usepackage{graphicx}
13 chrisfen 2595 \usepackage{booktabs}
14 chrisfen 2605 \usepackage{bibentry}
15     \usepackage{mathrsfs}
16 chrisfen 2575 \usepackage[ref]{overcite}
17     \pagestyle{plain}
18     \pagenumbering{arabic}
19     \oddsidemargin 0.0cm \evensidemargin 0.0cm
20     \topmargin -21pt \headsep 10pt
21     \textheight 9.0in \textwidth 6.5in
22     \brokenpenalty=10000
23     \renewcommand{\baselinestretch}{1.2}
24     \renewcommand\citemid{\ } % no comma in optional reference note
25    
26     \begin{document}
27    
28 gezelter 2617 \title{Is the Ewald Summation necessary? Pairwise alternatives to the accepted standard for long-range electrostatics}
29 chrisfen 2575
30 gezelter 2617 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
31     gezelter@nd.edu} \\
32 chrisfen 2575 Department of Chemistry and Biochemistry\\
33     University of Notre Dame\\
34     Notre Dame, Indiana 46556}
35    
36     \date{\today}
37    
38     \maketitle
39 gezelter 2617 \doublespacing
40    
41 chrisfen 2605 \nobibliography{}
42 chrisfen 2575 \begin{abstract}
43 gezelter 2617 A new method for accumulating electrostatic interactions was derived
44     from the previous efforts described in \bibentry{Wolf99} and
45     \bibentry{Zahn02} as a possible replacement for lattice sum methods in
46     molecular simulations. Comparisons were performed with this and other
47     pairwise electrostatic summation techniques against the smooth
48     particle mesh Ewald (SPME) summation to see how well they reproduce
49     the energetics and dynamics of a variety of simulation types. The
50     newly derived Shifted-Force technique shows a remarkable ability to
51     reproduce the behavior exhibited in simulations using SPME with an
52     $\mathscr{O}(N)$ computational cost, equivalent to merely the
53     real-space portion of the lattice summation.
54 chrisfen 2619
55 chrisfen 2575 \end{abstract}
56    
57 gezelter 2617 \newpage
58    
59 chrisfen 2575 %\narrowtext
60    
61 chrisfen 2601 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 chrisfen 2575 % BODY OF TEXT
63 chrisfen 2601 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64 chrisfen 2575
65     \section{Introduction}
66    
67 gezelter 2617 In molecular simulations, proper accumulation of the electrostatic
68     interactions is considered one of the most essential and
69     computationally demanding tasks.
70 chrisfen 2604
71 chrisfen 2608 \subsection{The Ewald Sum}
72 chrisfen 2604 blah blah blah Ewald Sum Important blah blah blah
73    
74 chrisfen 2610 \begin{figure}
75     \centering
76 gezelter 2617 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
77     \caption{How the application of the Ewald summation has changed with
78     the increase in computer power. Initially, only small numbers of
79     particles could be studied, and the Ewald sum acted to replicate the
80     unit cell charge distribution out to convergence. Now, much larger
81     systems of charges are investigated with fixed distance cutoffs. The
82     calculated structure factor is used to sum out to great distance, and
83     a surrounding dielectric term is included.}
84 chrisfen 2610 \label{fig:ewaldTime}
85     \end{figure}
86    
87 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
88 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
89     for an accurate accumulation of electrostatic interactions in an
90     efficient pairwise fashion.\cite{Wolf99} Wolf \textit{et al.} observed
91     that the electrostatic interaction is effectively short-ranged in
92     condensed phase systems and that neutralization of the charge
93     contained within the cutoff radius is crucial for potential
94     stability. They devised a pairwise summation method that ensures
95     charge neutrality and gives results similar to those obtained with
96     the Ewald summation. The resulting shifted Coulomb potential
97     (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
98     placement on the cutoff sphere and a distance-dependent damping
99     function (identical to that seen in the real-space portion of the
100     Ewald sum) to aid convergence
101 chrisfen 2601 \begin{equation}
102 chrisfen 2610 V^{\textrm{Wolf}}(r_{ij})= \frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
103 chrisfen 2601 \label{eq:WolfPot}
104     \end{equation}
105 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
106     potential. However, neutralizing the charge contained within each
107     cutoff sphere requires the placement of a self-image charge on the
108     surface of the cutoff sphere. This additional self-term in the total
109     potential enables Wolf {\it et al.} to obtain excellent estimates of
110     Madelung energies for many crystals.
111    
112     In order to use their charge-neutralized potential in molecular
113     dynamics simulations, Wolf \textit{et al.} suggested taking the
114     derivative of this potential prior to evaluation of the limit. This
115     procedure gives an expression for the forces,
116 chrisfen 2601 \begin{equation}
117 gezelter 2617 F^{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{-\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right]+\left[\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right]\right\},
118 chrisfen 2601 \label{eq:WolfForces}
119     \end{equation}
120 gezelter 2617 that incorporates both image charges and damping of the electrostatic
121     interaction.
122    
123     More recently, Zahn \textit{et al.} investigated these potential and
124     force expressions for use in simulations involving water.\cite{Zahn02}
125     In their work, they pointed out that the method that the forces and
126     derivative of the potential are not commensurate. Attempts to use
127     both Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will
128     lead to poor energy conservation. They correctly observed that taking
129     the limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating
130     the derivatives is mathematically invalid.
131    
132     Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
133     method'' as a way to use this technique in Molecular Dynamics
134     simulations. Taking the integral of the forces shown in equation
135     \ref{eq:WolfForces}, they proposed a new damped Coulomb
136     potential,
137 chrisfen 2601 \begin{equation}
138 chrisfen 2610 V^{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\}.
139 chrisfen 2601 \label{eq:ZahnPot}
140     \end{equation}
141 gezelter 2617 They showed that this potential does fairly well at capturing the
142     structural and dynamic properties of water compared the same
143     properties obtained using the Ewald sum.
144 chrisfen 2601
145 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
146    
147 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
148     al.} are constructed using two different (and separable) computational
149     tricks: \begin{itemize}
150     \item shifting through the use of image charges, and
151     \item damping the electrostatic interaction.
152     \end{itemize} Wolf \textit{et al.} treated the
153     development of their summation method as a progressive application of
154     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
155     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
156     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
157     both techniques. It is possible, however, to separate these
158     tricks and study their effects independently.
159    
160     Starting with the original observation that the effective range of the
161     electrostatic interaction in condensed phases is considerably less
162     than $r^{-1}$, either the cutoff sphere neutralization or the
163     distance-dependent damping technique could be used as a foundation for
164     a new pairwise summation method. Wolf \textit{et al.} made the
165     observation that charge neutralization within the cutoff sphere plays
166     a significant role in energy convergence; therefore we will begin our
167     analysis with the various shifted forms that maintain this charge
168     neutralization. We can evaluate the methods of Wolf
169     \textit{et al.} and Zahn \textit{et al.} by considering the standard
170     shifted potential,
171 chrisfen 2601 \begin{equation}
172 gezelter 2617 v^\textrm{SP}(r) = \begin{cases}
173     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
174     R_\textrm{c}
175     \end{cases},
176     \label{eq:shiftingPotForm}
177     \end{equation}
178     and shifted force,
179     \begin{equation}
180     v^\textrm{SF}(r) = \begin{cases}
181     v(r)-v_\textrm{c}-\left(\frac{\textrm{d}v(r)}{\textrm{d}r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
182     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
183 chrisfen 2601 \end{cases},
184 chrisfen 2612 \label{eq:shiftingForm}
185 chrisfen 2601 \end{equation}
186 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
187     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
188     that both the potential and the forces goes to zero at the cutoff
189     radius, while the Shifted Potential ({\sc sp}) form only ensures the
190     potential is smooth at the cutoff radius
191     ($R_\textrm{c}$).\cite{Allen87}
192    
193    
194    
195    
196     If the derivative term is taken to be zero, we are left with the shifted Coulomb potential devised by Wolf \textit{et al.},\cite{Wolf99}
197 chrisfen 2601 \begin{equation}
198 chrisfen 2612 V^\textrm{WSP}(r_{ij}) = \begin{cases} q_iq_j\left(\frac{1}{r_{ij}}-\frac{1}{R_\textrm{c}}\right) &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
199     \end{cases}.
200     \label{eq:WolfSP}
201     \end{equation}
202     The forces associated with this potential are obtained by taking the derivative, resulting in the following,
203     \begin{equation}
204     F^\textrm{WSP}(r_{ij}) = \begin{cases} q_iq_j\left(-\frac{1}{r_{ij}^2}\right) &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
205     \end{cases}.
206     \label{eq:FWolfSP}
207     \end{equation}
208     These forces are identical to the forces of the standard electrostatic interaction, and this was addressed by Wolf \textit{et al.} as undesirable. They pointed out that the effect of the image charges is neglected in the forces when they would expect there to be some pressure exerted due to their presence.\cite{Wolf99} As a consequence the forces, though mathematically valid, may not be physically correct due to this missing component. Additionally, there is a discontinuity in the forces. This can be remedied with the use of a switching function to zero the potential and forces smoothly as particles near $R_\textrm{c}$.
209    
210     If the derivative term in equation \ref{eq:shiftingForm} is evaluated, we obtain an hitherto undiscussed shifted force Coulomb potential,
211     \begin{equation}
212     V^\textrm{SF}(r_{ij}) = \begin{cases} q_iq_j\left\{\frac{1}{r_{ij}}-\frac{1}{R_\textrm{c}}+\left[\frac{1}{R_\textrm{c}^2}\right](r_{ij}-R_\textrm{c})\right\} &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
213     \end{cases}.
214     \label{eq:SFPot}
215     \end{equation}
216     Taking the derivative of this shifted force potential gives the following forces,
217     \begin{equation}
218     F^\textrm{SF}(r_{ij}) = \begin{cases} q_iq_j\left(-\frac{1}{r_{ij}^2}+\frac{1}{R_\textrm{c}^2}\right) &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
219     \end{cases}.
220     \label{eq:SFForces}
221     \end{equation}
222     Using this formulation rather than the simple shifted potential (Eq. \ref{eq:WolfSP}) means that there are no discontinuities in the forces in addition to the potential. This form also has the benefit that the image charges have a force presence, addressing the concerns about a missing physical component. One side effect of this treatment is a slight alteration in the shape of the potential that comes about from the derivative term. Thus, a degree of clarity about the original formulation of the potential is lost in order to gain functionality in dynamics simulations.
223    
224     Wolf \textit{et al.} originally discussed the energetics of the shifted Coulomb potential (Eq. \ref{eq:WolfSP}), and they found that it was still insufficient for accurate determination of the energy. The energy would fluctuate around the expected value with increasing cutoff radius, but the oscillations appeared to be converging toward the correct value.\cite{Wolf99} A damping function was incorporated to accelerate convergence; and though alternative functional forms could be used,\cite{Jones56,Heyes81} the complimentary error function was chosen to draw parallels to the Ewald summation. Incorporating damping into the simple Coulomb potential,
225     \begin{equation}
226 chrisfen 2601 v(r_{ij}) = \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}},
227     \label{eq:dampCoulomb}
228     \end{equation}
229 chrisfen 2612 the shifted potential (Eq. \ref{eq:WolfSP}) can be rederived \textit{via} equation \ref{eq:shiftingForm},
230 chrisfen 2601 \begin{equation}
231 chrisfen 2612 V^{\textrm{DSP}}(r_{ij}) = \begin{cases} q_iq_j\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\frac{\textrm{erfc}(\alpha R_\textrm{c})}{R_\textrm{c}}\right] &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
232     \end{cases}.
233     \label{eq:DSPPot}
234 chrisfen 2601 \end{equation}
235 chrisfen 2612 The derivative of this Shifted-Potential can be taken to obtain forces for use in MD,
236     \begin{equation}
237     F^{\textrm{DSP}}(r_{ij}) = \begin{cases} q_iq_j\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right] &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
238     \end{cases}.
239     \label{eq:DSPForces}
240     \end{equation}
241     Again, this Shifted-Potential suffers from a discontinuity in the forces, and a lack of an image-charge component in the forces. To remedy these concerns, a Shifted-Force variant is obtained by inclusion of the derivative term in equation \ref{eq:shiftingForm} to give,
242     \begin{equation}
243     V^\mathrm{DSF}(r_{ij}) = \begin{cases} q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}}\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\} &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
244     \end{cases}.
245     \label{eq:DSFPot}
246     \end{equation}
247     The derivative of the above potential gives the following forces,
248     \begin{equation}
249     F^\mathrm{DSF}(r_{ij}) = \begin{cases} q_iq_j\left\{-\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right]+\left[\frac{\textrm{erfc}(\alpha R_{\textrm{c}})}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2R_{\textrm{c}}^2)}}{R_{\textrm{c}}}\right]\right\} &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
250     \end{cases}.
251     \label{eq:DSFForces}
252     \end{equation}
253 chrisfen 2601
254 chrisfen 2612 This new Shifted-Force potential is similar to equation \ref{eq:ZahnPot} derived by Zahn \textit{et al.}; however, there are two important differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from equation \ref{eq:shiftingForm} is equal to equation \ref{eq:dampCoulomb} with $R_\textrm{c}$ supplied for $r_{ij}$. This term is not present in the Zahn potential, resulting in a discontinuity as particles cross $R_\textrm{c}$. Second, the sign of the derivative portion is different. The constant $v_\textrm{c}$ term is not going to have a presence in the forces after performing the derivative, but the negative sign does effect the derivative. In fact, it introduces a discontinuity in the forces at the cutoff, because the force function is shifted in the wrong direction and doesn't cross zero at $R_\textrm{c}$. Thus, these alterations make for an electrostatic summation method that is continuous in both the potential and forces and incorporates the pairwise sum considerations stressed by Wolf \textit{et al.}\cite{Wolf99}
255 chrisfen 2602
256 chrisfen 2608 \section{Methods}
257    
258 chrisfen 2609 \subsection{What Qualities are Important?}\label{sec:Qualities}
259     In classical molecular mechanics simulations, there are two primary techniques utilized to obtain information about the system of interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these techniques utilize pairwise summations of interactions between particle sites, but they use these summations in different ways.
260 chrisfen 2608
261 chrisfen 2609 In MC, the potential energy difference between two subsequent configurations dictates the progression of MC sampling. Going back to the origins of this method, the Canonical ensemble acceptance criteria laid out by Metropolis \textit{et al.} states that a subsequent configuration is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and 1.\cite{Metropolis53} Maintaining a consistent $\Delta E$ when using an alternate method for handling the long-range electrostatics ensures proper sampling within the ensemble.
262 chrisfen 2608
263 chrisfen 2609 In MD, the derivative of the potential directs how the system will progress in time. Consequently, the force and torque vectors on each body in the system dictate how it develops as a whole. If the magnitude and direction of these vectors are similar when using alternate electrostatic summation techniques, the dynamics in the near term will be indistinguishable. Because error in MD calculations is cumulative, one should expect greater deviation in the long term trajectories with greater differences in these vectors between configurations using different long-range electrostatics.
264 chrisfen 2608
265 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
266     Evaluation of the pairwise summation techniques (outlined in section \ref{sec:ESMethods}) for use in MC simulations was performed through study of the energy differences between conformations. Considering the SPME results to be the correct or desired behavior, ideal performance of a tested method was taken to be agreement between the energy differences calculated. Linear least squares regression of the $\Delta E$ values between configurations using SPME against $\Delta E$ values using tested methods provides a quantitative comparison of this agreement. Unitary results for both the correlation and correlation coefficient for these regressions indicate equivalent energetic results between the methods. The correlation is the slope of the plotted data while the correlation coefficient ($R^2$) is a measure of the of the data scatter around the fitted line and tells about the quality of the fit (Fig. \ref{fig:linearFit}).
267 chrisfen 2608
268 chrisfen 2609 \begin{figure}
269     \centering
270 chrisfen 2619 \includegraphics[width = \linewidth]{./dualLinear.pdf}
271     \caption{Example least squares regressions of the configuration energy differences for SPC/E water systems. The upper plot shows a data set with a poor correlation coefficient ($R^2$), while the lower plot shows a data set with a good correlation coefficient.}
272 chrisfen 2609 \label{fig:linearFit}
273     \end{figure}
274    
275 chrisfen 2619 Each system type (detailed in section \ref{sec:RepSims}) studied
276     consisted of 500 independent configurations, each equilibrated from
277     higher temperature trajectories. Thus, 124,750 $\Delta E$ data points
278     are used in a regression of a single system type. Results and
279     discussion for the individual analysis of each of the system types
280     appear in the supporting information, while the cumulative results
281     over all the investigated systems appears below in section
282     \ref{sec:EnergyResults}.
283 chrisfen 2609
284     \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
285     Evaluation of the pairwise methods (outlined in section \ref{sec:ESMethods}) for use in MD simulations was performed through comparison of the force and torque vectors obtained with those from SPME. Both the magnitude and the direction of these vectors on each of the bodies in the system were analyzed. For the magnitude of these vectors, linear least squares regression analysis can be performed as described previously for comparing $\Delta E$ values. Instead of a single value between two system configurations, there is a value for each particle in each configuration. For a system of 1000 water molecules and 40 ions, there are 1040 force vectors and 1000 torque vectors. With 500 configurations, this results in 520,000 force and 500,000 torque vector comparisons samples for each system type.
286    
287 chrisfen 2610 The force and torque vector directions were investigated through measurement of the angle ($\theta$) formed between those from the particular method and those from SPME
288     \begin{equation}
289     \theta_F = \frac{\vec{F}_\textrm{SPME}}{|\vec{F}_\textrm{SPME}|}\cdot\frac{\vec{F}_\textrm{Method}}{|\vec{F}_\textrm{Method}|}.
290     \end{equation}
291     Each of these $\theta$ values was accumulated in a distribution function, weighted by the area on the unit sphere. Non-linear fits were used to measure the shape of the resulting distributions.
292 chrisfen 2609
293     \begin{figure}
294     \centering
295 gezelter 2617 \includegraphics[width = \linewidth]{./gaussFit.pdf}
296 chrisfen 2609 \caption{Sample fit of the angular distribution of the force vectors over all of the studied systems. Gaussian fits were used to obtain values for the variance in force and torque vectors used in the following figure.}
297     \label{fig:gaussian}
298     \end{figure}
299    
300     Figure \ref{fig:gaussian} shows an example distribution with applied non-linear fits. The solid line is a Gaussian profile, while the dotted line is a Voigt profile, a convolution of a Gaussian and a Lorentzian. Since this distribution is a measure of angular error between two different electrostatic summation methods, there is particular reason for the profile to adhere to a specific shape. Because of this and the Gaussian profile's more statistically meaningful properties, Gaussian fits was used to compare all the tested methods. The variance ($\sigma^2$) was extracted from each of these fits and was used to compare distribution widths. Values of $\sigma^2$ near zero indicate vector directions indistinguishable from those calculated when using SPME.
301    
302     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
303     Evaluation of the long-time dynamics of charged systems was performed by considering the NaCl crystal system while using a subset of the best performing pairwise methods. The NaCl crystal was chosen to avoid possible complications involving the propagation techniques of orientational motion in molecular systems. To enhance the atomic motion, these crystals were equilibrated at 1000 K, near the experimental $T_m$ for NaCl. Simulations were performed under the microcanonical ensemble, and velocity autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each of the trajectories,
304     \begin{equation}
305     C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
306     \label{eq:vCorr}
307     \end{equation}
308     Velocity autocorrelation functions require detailed short time data and long trajectories for good statistics, thus velocity information was saved every 5 fs over 100 ps trajectories. The power spectrum ($I(\omega)$) is obtained via Fourier transform of the autocorrelation function
309     \begin{equation}
310     I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
311     \label{eq:powerSpec}
312     \end{equation}
313     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$.
314    
315     \subsection{Representative Simulations}\label{sec:RepSims}
316     A variety of common and representative simulations were analyzed to determine the relative effectiveness of the pairwise summation techniques in reproducing the energetics and dynamics exhibited by SPME. The studied systems were as follows:
317 chrisfen 2599 \begin{enumerate}
318 chrisfen 2586 \item Liquid Water
319     \item Crystalline Water (Ice I$_\textrm{c}$)
320 chrisfen 2595 \item NaCl Crystal
321     \item NaCl Melt
322 chrisfen 2599 \item Low Ionic Strength Solution of NaCl in Water
323     \item High Ionic Strength Solution of NaCl in Water
324 chrisfen 2586 \item 6 \AA\ Radius Sphere of Argon in Water
325 chrisfen 2599 \end{enumerate}
326 chrisfen 2609 By utilizing the pairwise techniques (outlined in section \ref{sec:ESMethods}) in systems composed entirely of neutral groups, charged particles, and mixtures of the two, we can comment on possible system dependence and/or universal applicability of the techniques.
327 chrisfen 2586
328 chrisfen 2601 Generation of the system configurations was dependent on the system type. For the solid and liquid water configurations, configuration snapshots were taken at regular intervals from higher temperature 1000 SPC/E water molecule trajectories and each equilibrated individually. The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl- ions and were selected and equilibrated in the same fashion as the water systems. For the low and high ionic strength NaCl solutions, 4 and 40 ions were first solvated in a 1000 water molecule boxes respectively. Ion and water positions were then randomly swapped, and the resulting configurations were again equilibrated individually. Finally, for the Argon/Water "charge void" systems, the identities of all the SPC/E waters within 6 \AA\ of the center of the equilibrated water configurations were converted to argon (Fig. \ref{fig:argonSlice}).
329 chrisfen 2586
330     \begin{figure}
331     \centering
332 gezelter 2617 \includegraphics[width = \linewidth]{./slice.pdf}
333 chrisfen 2586 \caption{A slice from the center of a water box used in a charge void simulation. The darkened region represents the boundary sphere within which the water molecules were converted to argon atoms.}
334 chrisfen 2601 \label{fig:argonSlice}
335 chrisfen 2586 \end{figure}
336    
337 chrisfen 2609 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
338     Electrostatic summation method comparisons were performed using SPME, the Shifted-Potential and Shifted-Force methods - both with damping parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak, moderate, and strong damping respectively), reaction field with an infinite dielectric constant, and an unmodified cutoff. Group-based cutoffs with a fifth-order polynomial switching function were necessary for the reaction field simulations and were utilized in the SP, SF, and pure cutoff methods for comparison to the standard lack of group-based cutoffs with a hard truncation. The SPME calculations were performed using the TINKER implementation of SPME,\cite{Ponder87} while all other method calculations were performed using the OOPSE molecular mechanics package.\cite{Meineke05}
339 chrisfen 2586
340 chrisfen 2609 These methods were additionally evaluated with three different cutoff radii (9, 12, and 15 \AA) to investigate possible cutoff radius dependence. It should be noted that the damping parameter chosen in SPME, or so called ``Ewald Coefficient", has a significant effect on the energies and forces calculated. Typical molecular mechanics packages default this to a value dependent on the cutoff radius and a tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically associated with increased accuracy in the real-space portion of the summation.\cite{Essmann95} The default TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
341    
342 chrisfen 2575 \section{Results and Discussion}
343    
344 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
345     In order to evaluate the performance of the pairwise electrostatic summation methods for Monte Carlo simulations, the energy differences between configurations were compared to the values obtained when using SPME. The results for the subsequent regression analysis are shown in figure \ref{fig:delE}.
346 chrisfen 2590
347     \begin{figure}
348     \centering
349 gezelter 2617 \includegraphics[width=5.5in]{./delEplot.pdf}
350 chrisfen 2608 \caption{Statistical analysis of the quality of configurational energy differences for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate $\Delta E$ values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
351 chrisfen 2601 \label{fig:delE}
352 chrisfen 2594 \end{figure}
353    
354 chrisfen 2610 In this figure, it is apparent that it is unreasonable to expect realistic results using an unmodified cutoff. This is not all that surprising since this results in large energy fluctuations as atoms move in and out of the cutoff radius. These fluctuations can be alleviated to some degree by using group based cutoffs with a switching function.\cite{Steinbach94} The Group Switch Cutoff row doesn't show a significant improvement in this plot because the salt and salt solution systems contain non-neutral groups, see the accompanying supporting information for a comparison where all groups are neutral.
355 chrisfen 2594
356 chrisfen 2610 Correcting the resulting charged cutoff sphere is one of the purposes of the damped Coulomb summation proposed by Wolf \textit{et al.},\cite{Wolf99} and this correction indeed improves the results as seen in the Shifted-Potental rows. While the undamped case of this method is a significant improvement over the pure cutoff, it still doesn't correlate that well with SPME. Inclusion of potential damping improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows an excellent correlation and quality of fit with the SPME results, particularly with a cutoff radius greater than 12 \AA . Use of a larger damping parameter is more helpful for the shortest cutoff shown, but it has a detrimental effect on simulations with larger cutoffs. In the Shifted-Force sets, increasing damping results in progressively poorer correlation. Overall, the undamped case is the best performing set, as the correlation and quality of fits are consistently superior regardless of the cutoff distance. This result is beneficial in that the undamped case is less computationally prohibitive do to the lack of complimentary error function calculation when performing the electrostatic pair interaction. The reaction field results illustrates some of that method's limitations, primarily that it was developed for use in homogenous systems; although it does provide results that are an improvement over those from an unmodified cutoff.
357 chrisfen 2609
358 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
359 chrisfen 2599
360 chrisfen 2610 Evaluation of pairwise methods for use in Molecular Dynamics simulations requires consideration of effects on the forces and torques. Investigation of the force and torque vector magnitudes provides a measure of the strength of these values relative to SPME. Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the force and torque vector magnitude regression results for the accumulated analysis over all the system types.
361 chrisfen 2594
362     \begin{figure}
363     \centering
364 gezelter 2617 \includegraphics[width=5.5in]{./frcMagplot.pdf}
365 chrisfen 2608 \caption{Statistical analysis of the quality of the force vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate force magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
366 chrisfen 2601 \label{fig:frcMag}
367 chrisfen 2594 \end{figure}
368    
369 chrisfen 2610 Figure \ref{fig:frcMag}, for the most part, parallels the results seen in the previous $\Delta E$ section. The unmodified cutoff results are poor, but using group based cutoffs and a switching function provides a improvement much more significant than what was seen with $\Delta E$. Looking at the Shifted-Potential sets, the slope and $R^2$ improve with the use of damping to an optimal result of 0.2 \AA $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping, while beneficial for simulations with a cutoff radius of 9 \AA\ , is detrimental to simulations with larger cutoff radii. The undamped Shifted-Force method gives forces in line with those obtained using SPME, and use of a damping function results in minor improvement. The reaction field results are surprisingly good, considering the poor quality of the fits for the $\Delta E$ results. There is still a considerable degree of scatter in the data, but it correlates well in general. To be fair, we again note that the reaction field calculations do not encompass NaCl crystal and melt systems, so these results are partly biased towards conditions in which the method performs more favorably.
370 chrisfen 2594
371     \begin{figure}
372     \centering
373 gezelter 2617 \includegraphics[width=5.5in]{./trqMagplot.pdf}
374 chrisfen 2608 \caption{Statistical analysis of the quality of the torque vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate torque magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
375 chrisfen 2601 \label{fig:trqMag}
376 chrisfen 2594 \end{figure}
377    
378 chrisfen 2610 To evaluate the torque vector magnitudes, the data set from which values are drawn is limited to rigid molecules in the systems (i.e. water molecules). In spite of this smaller sampling pool, the torque vector magnitude results in figure \ref{fig:trqMag} are still similar to those seen for the forces; however, they more clearly show the improved behavior that comes with increasing the cutoff radius. Moderate damping is beneficial to the Shifted-Potential and helpful yet possibly unnecessary with the Shifted-Force method, and they also show that over-damping adversely effects all cutoff radii rather than showing an improvement for systems with short cutoffs. The reaction field method performs well when calculating the torques, better than the Shifted Force method over this limited data set.
379 chrisfen 2594
380 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
381 chrisfen 2599
382 chrisfen 2610 Having force and torque vectors with magnitudes that are well correlated to SPME is good, but if they are not pointing in the proper direction the results will be incorrect. These vector directions were investigated through measurement of the angle formed between them and those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared through the variance ($\sigma^2$) of the Gaussian fits of the angle error distributions of the combined set over all system types.
383 chrisfen 2594
384     \begin{figure}
385     \centering
386 gezelter 2617 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
387 chrisfen 2608 \caption{Statistical analysis of the quality of the Gaussian fit of the force and torque vector angular distributions for a given electrostatic method compared with the reference Ewald sum. Results with a variance ($\sigma^2$) equal to zero (dashed line) indicate force and torque directions indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
388 chrisfen 2601 \label{fig:frcTrqAng}
389 chrisfen 2594 \end{figure}
390    
391 chrisfen 2610 Both the force and torque $\sigma^2$ results from the analysis of the total accumulated system data are tabulated in figure \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case show the improvement afforded by choosing a longer simulation cutoff. Increasing the cutoff from 9 to 12 \AA\ typically results in a halving of the distribution widths, with a similar improvement going from 12 to 15 \AA . The undamped Shifted-Force, Group Based Cutoff, and Reaction Field methods all do equivalently well at capturing the direction of both the force and torque vectors. Using damping improves the angular behavior significantly for the Shifted-Potential and moderately for the Shifted-Force methods. Increasing the damping too far is destructive for both methods, particularly to the torque vectors. Again it is important to recognize that the force vectors cover all particles in the systems, while torque vectors are only available for neutral molecular groups. Damping appears to have a more beneficial effect on non-neutral bodies, and this observation is investigated further in the accompanying supporting information.
392 chrisfen 2594
393 chrisfen 2595 \begin{table}[htbp]
394     \centering
395     \caption{Variance ($\sigma^2$) of the force (top set) and torque (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods. Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function. The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}
396 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
397 chrisfen 2595 \\
398     \toprule
399     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
400     \cmidrule(lr){3-6}
401     \cmidrule(l){7-10}
402 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
403 chrisfen 2595 \midrule
404 chrisfen 2599
405     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
406     & \textbf{Y} & \textbf{2.486} & \textbf{2.160} & \textbf{0.667} & \textbf{0.608} & \textbf{1.768} & \textbf{1.766} & \textbf{0.676} & \textbf{0.609} \\
407     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
408     & \textbf{Y} & \textbf{0.515} & \textbf{0.288} & \textbf{0.127} & \textbf{0.586} & \textbf{0.308} & \textbf{0.249} & \textbf{0.127} & \textbf{0.586} \\
409     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
410     & \textbf{Y} & \textbf{0.228} & \textbf{0.099} & \textbf{0.121} & \textbf{0.598} & \textbf{0.144} & \textbf{0.090} & \textbf{0.121} & \textbf{0.598} \\
411 chrisfen 2594
412 chrisfen 2595 \midrule
413    
414 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
415     & \textbf{Y} & \textbf{2.115} & \textbf{1.914} & \textbf{1.878} & \textbf{5.142} & \textbf{2.076} & \textbf{2.039} & \textbf{1.972} & \textbf{5.146} \\
416     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
417     & \textbf{Y} & \textbf{0.810} & \textbf{0.685} & \textbf{1.352} & \textbf{5.082} & \textbf{0.765} & \textbf{0.714} & \textbf{1.360} & \textbf{5.082} \\
418     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
419     & \textbf{Y} & \textbf{0.282} & \textbf{0.294} & \textbf{1.272} & \textbf{4.999} & \textbf{0.324} & \textbf{0.318} & \textbf{1.272} & \textbf{4.999} \\
420 chrisfen 2595
421     \bottomrule
422     \end{tabular}
423 chrisfen 2601 \label{tab:groupAngle}
424 chrisfen 2595 \end{table}
425    
426 chrisfen 2610 Although not discussed previously, group based cutoffs can be applied to both the Shifted-Potential and Shifted-Force methods. Use off a switching function corrects for the discontinuities that arise when atoms of a group exit the cutoff before the group's center of mass. Though there are no significant benefit or drawbacks observed in $\Delta E$ and vector magnitude results when doing this, there is a measurable improvement in the vector angle results. Table \ref{tab:groupAngle} shows the angular variance values obtained using group based cutoffs and a switching function alongside the standard results seen in figure \ref{fig:frcTrqAng} for comparison purposes. The Shifted-Potential shows much narrower angular distributions for both the force and torque vectors when using an $\alpha$ of 0.2 \AA$^{-1}$ or less, while Shifted-Force shows improvements in the undamped and lightly damped cases. Thus, by calculating the electrostatic interactions in terms of molecular pairs rather than atomic pairs, the direction of the force and torque vectors are determined more accurately.
427 chrisfen 2595
428 chrisfen 2610 One additional trend to recognize in table \ref{tab:groupAngle} is that the $\sigma^2$ values for both Shifted-Potential and Shifted-Force converge as $\alpha$ increases, something that is easier to see when using group based cutoffs. Looking back on figures \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this behavior clearly at large $\alpha$ and cutoff values. The reason for this is that the complimentary error function inserted into the potential weakens the electrostatic interaction as $\alpha$ increases. Thus, at larger values of $\alpha$, both the summation method types progress toward non-interacting functions, so care is required in choosing large damping functions lest one generate an undesirable loss in the pair interaction. Kast \textit{et al.} developed a method for choosing appropriate $\alpha$ values for these types of electrostatic summation methods by fitting to $g(r)$ data, and their methods indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear to be reasonable choices to obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on these findings, choices this high would introduce error in the molecular torques, particularly for the shorter cutoffs. Based on the above findings, empirical damping up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably unnecessary when using the Shifted-Force method.
429 chrisfen 2595
430 chrisfen 2608 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
431 chrisfen 2601
432     In the previous studies using a Shifted-Force variant of the damped Wolf coulomb potential, the structure and dynamics of water were investigated rather extensively.\cite{Zahn02,Kast03} Their results indicated that the damped Shifted-Force method results in properties very similar to those obtained when using the Ewald summation. Considering the statistical results shown above, the good performance of this method is not that surprising. Rather than consider the same systems and simply recapitulate their results, we decided to look at the solid state dynamical behavior obtained using the best performing summation methods from the above results.
433    
434     \begin{figure}
435     \centering
436 gezelter 2617 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
437 chrisfen 2616 \caption{Power spectra obtained from the velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, Shifted-Force ($\alpha$ = 0, 0.1, \& 0.2), and Shifted-Potential ($\alpha$ = 0.2). Apodization of the correlation functions via a cubic switching function between 40 and 50 ps was used to clear up the spectral noise resulting from data truncation, and had no noticeable effect on peak location or magnitude. The inset shows the frequency region below 100 cm$^{-1}$ to highlight where the spectra begin to differ.}
438 chrisfen 2610 \label{fig:methodPS}
439 chrisfen 2601 \end{figure}
440    
441 chrisfen 2610 Figure \ref{fig:methodPS} shows the power spectra for the NaCl crystals (from averaged Na and Cl ion velocity autocorrelation functions) using the stated electrostatic summation methods. While high frequency peaks of all the spectra overlap, showing the same general features, the low frequency region shows how the summation methods differ. Considering the low-frequency inset (expanded in the upper frame of figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the correlated motions are blue-shifted when using undamped or weakly damped Shifted-Force. When using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the Shifted-Force and Shifted-Potential methods give near identical correlated motion behavior as the Ewald method (which has a damping value of 0.3119). The damping acts as a distance dependent Gaussian screening of the point charges for the pairwise summation methods. This weakening of the electrostatic interaction with distance explains why the long-ranged correlated motions are at lower frequencies for the moderately damped methods than for undamped or weakly damped methods. To see this effect more clearly, we show how damping strength affects a simple real-space electrostatic potential,
442 chrisfen 2601 \begin{equation}
443     V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r_{ij}})}{r_{ij}}\right]S(r),
444     \end{equation}
445 chrisfen 2610 where $S(r)$ is a switching function that smoothly zeroes the potential at the cutoff radius. Figure \ref{fig:dampInc} shows how the low frequency motions are dependent on the damping used in the direct electrostatic sum. As the damping increases, the peaks drop to lower frequencies. Incidentally, use of an $\alpha$ of 0.25 \AA$^{-1}$ on a simple electrostatic summation results in low frequency correlated dynamics equivalent to a simulation using SPME. When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks shift to higher frequency in exponential fashion. Though not shown, the spectrum for the simple undamped electrostatic potential is blue-shifted such that the lowest frequency peak resides near 325 cm$^{-1}$. In light of these results, the undamped Shifted-Force method producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite respectable; however, it appears as though moderate damping is required for accurate reproduction of crystal dynamics.
446 chrisfen 2601 \begin{figure}
447     \centering
448 gezelter 2617 \includegraphics[width = \linewidth]{./comboSquare.pdf}
449 chrisfen 2615 \caption{Upper: Zoomed inset from figure \ref{fig:methodPS}. As the damping value for the Shifted-Force potential increases, the low-frequency peaks red-shift. Lower: Low-frequency correlated motions for NaCl crystals at 1000 K when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$. As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME. The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
450 chrisfen 2601 \label{fig:dampInc}
451     \end{figure}
452    
453 chrisfen 2575 \section{Conclusions}
454    
455 chrisfen 2608 This investigation of pairwise electrostatic summation techniques shows that there are viable and more computationally efficient electrostatic summation techniques than the Ewald summation, chiefly methods derived from the damped Coulombic sum originally proposed by Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the Shifted-Force method, reformulated above as equation \ref{eq:SFPot}, shows a remarkable ability to reproduce the energetic and dynamic characteristics exhibited by simulations employing lattice summation techniques. The cumulative energy difference results showed the undamped Shifted-Force and moderately damped Shifted-Potential methods produced results nearly identical to SPME. Similarly for the dynamic features, the undamped or moderately damped Shifted-Force and moderately damped Shifted-Potential methods produce force and torque vector magnitude and directions very similar to the expected values. These results translate into long-time dynamic behavior equivalent to that produced in simulations using SPME.
456 chrisfen 2604
457 chrisfen 2608 Aside from the computational cost benefit, these techniques have applicability in situations where the use of the Ewald sum can prove problematic. Primary among them is their use in interfacial systems, where the unmodified lattice sum techniques artificially accentuate the periodicity of the system in an undesirable manner. There have been alterations to the standard Ewald techniques, via corrections and reformulations, to compensate for these systems; but the pairwise techniques discussed here require no modifications, making them natural tools to tackle these problems. Additionally, this transferability gives them benefits over other pairwise methods, like reaction field, because estimations of physical properties (e.g. the dielectric constant) are unnecessary.
458 chrisfen 2605
459 chrisfen 2608 We are not suggesting any flaw with the Ewald sum; in fact, it is the standard by which these simple pairwise sums are judged. However, these results do suggest that in the typical simulations performed today, the Ewald summation may no longer be required to obtain the level of accuracy most researcher have come to expect
460 chrisfen 2605
461 chrisfen 2575 \section{Acknowledgments}
462 chrisfen 2594 \newpage
463    
464 gezelter 2617 \bibliographystyle{jcp2}
465 chrisfen 2575 \bibliography{electrostaticMethods}
466    
467    
468     \end{document}