ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2605
Committed: Wed Mar 8 15:14:20 2006 UTC (18 years, 5 months ago) by chrisfen
Content type: application/x-tex
File size: 36974 byte(s)
Log Message:
Abstract and conclusion added

File Contents

# User Rev Content
1 chrisfen 2575 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2     \documentclass[12pt]{article}
3     \usepackage{endfloat}
4     \usepackage{amsmath}
5 chrisfen 2594 \usepackage{amssymb}
6 chrisfen 2599 %\usepackage{ifsym}
7 chrisfen 2575 \usepackage{epsf}
8     \usepackage{times}
9     \usepackage{mathptm}
10     \usepackage{setspace}
11     \usepackage{tabularx}
12     \usepackage{graphicx}
13 chrisfen 2595 \usepackage{booktabs}
14 chrisfen 2605 \usepackage{bibentry}
15     \usepackage{mathrsfs}
16 chrisfen 2586 %\usepackage{berkeley}
17 chrisfen 2575 \usepackage[ref]{overcite}
18     \pagestyle{plain}
19     \pagenumbering{arabic}
20     \oddsidemargin 0.0cm \evensidemargin 0.0cm
21     \topmargin -21pt \headsep 10pt
22     \textheight 9.0in \textwidth 6.5in
23     \brokenpenalty=10000
24     \renewcommand{\baselinestretch}{1.2}
25     \renewcommand\citemid{\ } % no comma in optional reference note
26    
27     \begin{document}
28    
29 chrisfen 2599 \title{Is the Ewald Summation necessary in typical molecular simulations: Alternatives to the accepted standard of cutoff policies}
30 chrisfen 2575
31     \author{Christopher J. Fennell and J. Daniel Gezelter \\
32     Department of Chemistry and Biochemistry\\
33     University of Notre Dame\\
34     Notre Dame, Indiana 46556}
35    
36     \date{\today}
37    
38     \maketitle
39     %\doublespacing
40 chrisfen 2605 \nobibliography{}
41 chrisfen 2575 \begin{abstract}
42 chrisfen 2605 A new method for accumulating electrostatic interactions was derived from the previous efforts described in \bibentry{Wolf99} and \bibentry{Zahn02} as a possible replacement for lattice sum methods in molecular simulations. Comparisons were performed with this and other pairwise electrostatic summation techniques against the smooth particle mesh Ewald (SPME) summation to see how well they reproduce the energetics and dynamics of a variety of simulation types. The newly derived Shifted-Force technique shows a remarkable ability to reproduce the behavior exhibited in simulations using SPME with an $\mathscr{O}(N)$ computational cost, equivalent to merely the real-space portion of the lattice summation.
43 chrisfen 2575 \end{abstract}
44    
45     %\narrowtext
46    
47 chrisfen 2601 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48 chrisfen 2575 % BODY OF TEXT
49 chrisfen 2601 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50 chrisfen 2575
51     \section{Introduction}
52    
53 chrisfen 2605 In molecular simulations, proper accumulation of the electrostatic interactions is considered one of the most essential and computationally demanding tasks.
54 chrisfen 2604
55     blah blah blah Ewald Sum Important blah blah blah
56    
57 chrisfen 2601 In a recent paper by Wolf \textit{et al.}, a procedure was outlined for accumulation of electrostatic interactions in a simple pairwise fashion.\cite{Wolf99} They took the observation that the electrostatic interaction is short-ranged in systems of charges and that charge neutralization within the cutoff spheres is crucial for potential stability, and they devised a pairwise summation method that ensures charge neutrality and gives results similar to those obtained using the Ewald summation. The resulting shifted Coulomb potential (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through placement on the cutoff sphere and a distance-dependent damping function (identical to that seen in the real-space portion of the Ewald sum) to hasten energetic convergence
58     \begin{equation}
59     V(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\}.
60     \label{eq:WolfPot}
61     \end{equation}
62     In order to use this potential in molecular dynamics simulations, the derivative of this potential was taken, followed by evaluation of the limit to give the following forces,
63     \begin{equation}
64     F(r_{ij}) = 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{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right].
65     \label{eq:WolfForces}
66     \end{equation}
67     More recently, Zahn \textit{et al.} investigated this electrostatic summation method for use in simulations involving water.\cite{Zahn02} In their work, they point out that the method as proposed is problematic for use in Molecular Dynamics simulations, because the integral of the forces and derivative of the potential are not equivalent. This comes about from the procedure of taking the limit shown in equation \ref{eq:WolfPot} after calculating the derivatives.\cite{Wolf99} Zahn \textit{et al.} proposed a shifted force adaptation of this "Wolf summation method" as a way to use this technique in Molecular Dynamics simulations. Taking the integral of the forces shown in equation \ref{eq:WolfForces}, they obtained a new shifted damped Coulomb potential
68     \begin{equation}
69     V\left(r_{ij}\right)=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\}.
70     \label{eq:ZahnPot}
71     \end{equation}
72     They showed that this new potential does well in capturing the structural and dynamic properties of water in their simulations.
73    
74     While implementing these methods for use in our own work, we discovered the potential presented in equation \ref{eq:ZahnPot} is still not entirely correct. The derivative of this equation leads to a sign error in the forces, resulting in erroneous dynamics. We can apply the standard shifted force potential form for Lennard-Jones potentials,
75     \begin{equation}
76     V^\textrm{SF}(r_{ij}) = \begin{cases} v(r_{ij})-v_\textrm{c}-\left(\frac{\textrm{d}v(r_{ij})}{\textrm{d}r_{ij}}\right)_{r_{ij}=R_\textrm{c}}(r_{ij}-R_\textrm{c}) &\quad r_{ij}\leqslant R_\textrm{c} \\ 0 &\quad r_{ij}>R_\textrm{c}
77     \end{cases},
78     \end{equation}
79     where $v(r_{ij})$ unshifted form of the potential, and $v_c$ is a constant term that insures the potential goes to zero at the cutoff radius.\cite{Allen87} Using the simple damped Coulomb potential as the starting point,
80     \begin{equation}
81     v(r_{ij}) = \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}},
82     \label{eq:dampCoulomb}
83     \end{equation}
84     the resulting shifted force potential is
85     \begin{equation}
86     V^\mathrm{SF}\left(r_{ij}\right)=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\}.
87     \label{eq:SFPot}
88     \end{equation}
89 chrisfen 2602 Equation \ref{eq:SFPot} 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 is simply equation \ref{eq:dampCoulomb} with $R_\textrm{c}$ supplied for $r_{ij}$. This term is not present in equation \ref{eq:ZahnPot}, resulting in a discontinuity in the potential 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}
90 chrisfen 2601
91 chrisfen 2604 It is important to note that shifted force techniques have a drawback in that they alter the shape of the original potential. We thereby lose a degree of clarity about the original formulation of the potential in order to gain functionality in dynamics simulations. An alternative direction would be use the derivatives of the original potential for the forces. This was addressed by Wolf \textit{et al.} as undesirable, because 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. In Monte Carlo simulations, this argument is mute, because forces are not evaluated. We decided to consider both the Shifted-Force technique described above and this Shifted-Potential technique to determine their usability in the evaluation of both energetic and dynamic results in simulations with electrostatics.
92 chrisfen 2602
93 chrisfen 2604 A variety of simulation situations were assembled and analyzed to determine the relative effectiveness of the adapted Wolf spherical truncation schemes at reproducing the results obtained using a smooth particle mesh Ewald (SPME) summation technique.\cite{Essmann95} In addition to the Shifted-Potential and Shifted-Force adapted Wolf methods, both reaction field and uncorrected cutoff methods were included for comparison purposes. The general usability of these methods in both Monte Carlo and Molecular Dynamics calculations was assessed through statistical analysis over the combined results from all of the following studied systems:
94 chrisfen 2599 \begin{enumerate}
95 chrisfen 2586 \item Liquid Water
96     \item Crystalline Water (Ice I$_\textrm{c}$)
97 chrisfen 2595 \item NaCl Crystal
98     \item NaCl Melt
99 chrisfen 2599 \item Low Ionic Strength Solution of NaCl in Water
100     \item High Ionic Strength Solution of NaCl in Water
101 chrisfen 2586 \item 6 \AA\ Radius Sphere of Argon in Water
102 chrisfen 2599 \end{enumerate}
103 chrisfen 2604 By studying these methods in systems composed entirely of neutral groups, charged particles, and mixtures of the two, we can either comment on possible system dependence or universal applicability of the techniques.
104 chrisfen 2586
105 chrisfen 2575 \section{Methods}
106    
107 chrisfen 2605 In each of the simulated systems, 500 distinct configurations were generated, and the electrostatic summation methods were compared via sequential application on each of these fixed configurations. The methods compared include SPME, the aforementioned Shifted Potential and Shifted Force methods - both with damping parameters ($\alpha$) of 0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, light, moderate, and heavy 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, while all all other method calculations were performed using the OOPSE molecular mechanics package.\cite{Ponder87,Meineke05}
108 chrisfen 2586
109 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}).
110 chrisfen 2586
111     \begin{figure}
112     \centering
113     \includegraphics[width=3.25in]{./slice.pdf}
114     \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.}
115 chrisfen 2601 \label{fig:argonSlice}
116 chrisfen 2586 \end{figure}
117    
118 chrisfen 2604 All of these comparisons were performed with three different cutoff radii (9, 12, and 15 \AA) to investigate the cutoff radius dependence of the various techniques. 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.\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.
119 chrisfen 2586
120 chrisfen 2575 \section{Results and Discussion}
121    
122 chrisfen 2599 \subsection{$\Delta E$ Comparison}
123 chrisfen 2601 In order to evaluate the performance of the adapted Wolf Shifted Potential and Shifted Force electrostatic summation methods for Monte Carlo simulations, the energy differences between configurations need to be compared to the results using SPME. Considering the SPME results to be the correct or desired behavior, ideal performance of a tested method is 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 gives an idea of the quality of the fit (Fig. \ref{fig:linearFit}).
124 chrisfen 2590
125     \begin{figure}
126     \centering
127     \includegraphics[width=3.25in]{./linearFit.pdf}
128 chrisfen 2595 \caption{Example least squares regression of the $\Delta E$ between configurations for the SF method against SPME in the pure water system. }
129 chrisfen 2601 \label{fig:linearFit}
130 chrisfen 2590 \end{figure}
131    
132 chrisfen 2601 With 500 independent configurations, 124,750 $\Delta E$ data points are used in a regression of a single system. Results and discussion for the individual analysis of each of the system types appear in the supporting information. To probe the applicability of each method in the general case, all the different system types were included in a single regression. The results for this regression are shown in figure \ref{fig:delE}.
133 chrisfen 2590
134 chrisfen 2594 \begin{figure}
135     \centering
136     \includegraphics[width=3.25in]{./delEplot.pdf}
137 chrisfen 2595 \caption{The results from the statistical analysis of the $\Delta$E results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii. Results close to a value of 1 (dashed line) indicate $\Delta E$ values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME. Reaction Field results do not include NaCl crystal or melt configurations.}
138 chrisfen 2601 \label{fig:delE}
139 chrisfen 2594 \end{figure}
140    
141 chrisfen 2601 In figure \ref{fig:delE}, 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. 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. Correcting the resulting charged cutoff sphere is one of the purposes of the shifted potential proposed by Wolf \textit{et al.}, 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.
142 chrisfen 2594
143 chrisfen 2599 \subsection{Force Magnitude Comparison}
144    
145 chrisfen 2601 While studying the energy differences provides insight into how comparable these methods are energetically, if we want to use these methods in Molecular Dynamics simulations, we also need to consider their effect on forces and torques. Both the magnitude and the direction of the force and torque vectors of each of the bodies in the system can be compared to those observed while using SPME. Analysis of the magnitude of these vectors can be performed in the manner described previously for comparing $\Delta E$ values, only 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 excess of 500,000 data samples for each system type. Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the force and torque vector magnitude results for the accumulated analysis over all the system types.
146 chrisfen 2594
147     \begin{figure}
148     \centering
149     \includegraphics[width=3.25in]{./frcMagplot.pdf}
150     \caption{The results from the statistical analysis of the force vector magnitude results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii. Results close to a value of 1 (dashed line) indicate force vector magnitude values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME.}
151 chrisfen 2601 \label{fig:frcMag}
152 chrisfen 2594 \end{figure}
153    
154 chrisfen 2605 The results in figure \ref{fig:frcMag} for the most part parallel those seen in the previous look at the $\Delta E$ results. 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.
155 chrisfen 2594
156 chrisfen 2599 \subsection{Torque Magnitude Comparison}
157    
158 chrisfen 2594 \begin{figure}
159     \centering
160     \includegraphics[width=3.25in]{./trqMagplot.pdf}
161     \caption{The results from the statistical analysis of the torque vector magnitude results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii. Results close to a value of 1 (dashed line) indicate torque vector magnitude values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME. Torques are only accumulated on the rigid water molecules, so these results exclude NaCl the systems.}
162 chrisfen 2601 \label{fig:trqMag}
163 chrisfen 2594 \end{figure}
164    
165 chrisfen 2601 The torque vector magnitude results in figure \ref{fig:trqMag} are similar to those seen for the forces, but more clearly show the improved behavior with increasing cutoff radius. Moderate damping is beneficial to the Shifted Potential and 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.
166 chrisfen 2594
167 chrisfen 2599 \subsection{Force and Torque Direction Comparison}
168    
169 chrisfen 2595 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 dot product of these unit vectors provides a theta value that is accumulated in a distribution function, weighted by the area on the unit sphere. Narrow distributions of theta values indicates similar to identical results between the tested method and SPME. To measure the narrowness of the resulting distributions, non-linear Gaussian fits were performed.
170 chrisfen 2594
171     \begin{figure}
172     \centering
173     \includegraphics[width=3.25in]{./gaussFit.pdf}
174     \caption{Example fitting of the angular distribution of the force vectors over all of the studied systems. The solid and dotted lines show Gaussian and Voigt fits of the distribution data respectively. Even though the Voigt profile make for a more accurate fit, the Gaussian was used due to more versatile statistical results.}
175 chrisfen 2601 \label{fig:gaussian}
176 chrisfen 2594 \end{figure}
177    
178 chrisfen 2601 Figure \ref{fig:gaussian} shows an example distribution and the non-linear fit applied. The solid line is a Gaussian profile, while the dotted line is a Voigt profile, a convolution of a Gaussian and a Lorentzian profile. Since this distribution is a measure of angular error between two different electrostatic summation methods, there is particular reason for it to adhere to a particular shape. Because of this and the Gaussian profile's more statistically meaningful properties, Gaussian fitting was used to compare all the methods considered in this study. The results (Fig. \ref{fig:frcTrqAng}) are compared through the variance ($\sigma^2$) of these non-linear fits.
179 chrisfen 2594
180     \begin{figure}
181     \centering
182     \includegraphics[width=3.25in]{./frcTrqAngplot.pdf}
183     \caption{The results from the statistical analysis of the force and torque vector angular distributions for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii. Plotted values are the variance ($\sigma^2$) of the Gaussian non-linear fits. Results close to a value of 0 (dashed line) indicate force or torque vector directions from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME. Torques are only accumulated on the rigid water molecules, so the torque vector angle results exclude NaCl the systems.}
184 chrisfen 2601 \label{fig:frcTrqAng}
185 chrisfen 2594 \end{figure}
186    
187 chrisfen 2601 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 $\sigma^2$, 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.
188 chrisfen 2594
189 chrisfen 2595 \begin{table}[htbp]
190     \centering
191     \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$^2$.}
192 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
193 chrisfen 2595 \\
194     \toprule
195     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
196     \cmidrule(lr){3-6}
197     \cmidrule(l){7-10}
198 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
199 chrisfen 2595 \midrule
200 chrisfen 2599
201     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
202     & \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} \\
203     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
204     & \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} \\
205     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
206     & \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} \\
207 chrisfen 2594
208 chrisfen 2595 \midrule
209    
210 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
211     & \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} \\
212     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
213     & \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} \\
214     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
215     & \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} \\
216 chrisfen 2595
217     \bottomrule
218     \end{tabular}
219 chrisfen 2601 \label{tab:groupAngle}
220 chrisfen 2595 \end{table}
221    
222 chrisfen 2605 Although not discussed previously, group based cutoffs can be applied to both the Shifted Potential and 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.
223 chrisfen 2595
224 chrisfen 2601 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 is arguably unnecessary when using the Shifted-Force method.
225 chrisfen 2595
226 chrisfen 2601 \subsection{Crystal Power Spectrum}
227    
228     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.
229    
230     Using the NaCl crystal as the model system, trajectories were obtained using SPME; Shifted-Force with $\alpha$ values of 0, 0.1 and 0.2 \AA$^{-1}$; and Shifted-Potential with an $\alpha$ value of 0.2 \AA$^{-1}$. To enhance the atomic motion, these simulations were run at 1000 K, near the experimental $T_m$ for NaCl. The velocity autocorrelation function (Eq. \ref{eq:vCorr})was computed on each of the trajectories.
231     \begin{equation}
232     C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
233     \label{eq:vCorr}
234     \end{equation}
235 chrisfen 2605 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 discrete Fourier transform of the autocorrelation function
236 chrisfen 2601 \begin{equation}
237 chrisfen 2605 I(\omega) = \sum^{N-1}_{\omega=0}C_v(t)e^{-i\omega t/N},
238 chrisfen 2601 \label{eq:powerSpec}
239     \end{equation}
240     where $N$ is the number of time samples in $C_v(t)$ and the frequency, $\omega=0,\ 1,\ ...,\ N-1$. The resulting spectra (Fig. \ref{fig:normalModes}) show the normal mode frequencies for the crystal under the simulated conditions.
241    
242     \begin{figure}
243     \centering
244     \includegraphics[width = 3.25in]{./nModeFTPlotDot.pdf}
245     \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. The inset shows the frequency region below 100 cm$^{-1}$ to highlight where the spectra begin to differentiate.}
246     \label{fig:normalModes}
247     \end{figure}
248    
249     Figure \ref{fig:normalModes} 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. The normal modes at frequencies below 100 cm$^{-1}$ are shifted up 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 normal mode 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 in the system. This weakening of the electrostatic interaction with distance explains why the low level normal modes are at lower frequencies for the moderately damped methods than for undamped or weakly damped methods. Consider damping on a simple real-space electrostatic potential in the form
250     \begin{equation}
251     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),
252     \end{equation}
253     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 normal modes are dependent on the damping used in the direct electrostatic sum. As the damping increases, the normal modes drop to lower frequencies. Incidentally, use of an $\alpha$ of 0.25 \AA$^{-1}$ on a simple electrostatic summation results in low frequency normal mode dynamics equivalent to a simulation using SPME. When the coefficient lowers to 0.15 \AA$^{-1}$ and below, the normal modes 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 normal mode resides near 325 cm$^{-1}$. In light of these results, the undamped Shifted-Force method producing low-lying normal modes within 10 cm$^{-1}$ of SPME is quite respectable; however, it appears as though moderate damping is required for accurate reproduction of crystal dynamics.
254     \begin{figure}
255     \centering
256     \includegraphics[width = 3.25in]{./alphaCompare.pdf}
257 chrisfen 2605 \caption{Normal modes for an NaCl crystal 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 normal modes are red-shifted towards 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.}
258 chrisfen 2601 \label{fig:dampInc}
259     \end{figure}
260    
261 chrisfen 2575 \section{Conclusions}
262    
263 chrisfen 2605 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, 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 un- to 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.
264 chrisfen 2604
265 chrisfen 2605 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 these pairwise techniques require no modifications, making them natural tools to tackle these problems. Additionally, this transferability gives it benefits over other pairwise methods, like reaction field, because estimations of physical properties, like the dielectric constant, are unnecessary.
266    
267     These results don't deprecate the use of the Ewald summation; in fact, it is the standard to which these simple pairwise sums are judged. However, these results do speak to the necessity of the Ewald summation in all molecular simulations. That a simple pairwise technique can be substituted to gain nearly all the physical effects provided by the full lattice sum makes us question whether the minimal perturbations bestowed through added complexity and increased cost are worth it.
268    
269 chrisfen 2575 \section{Acknowledgments}
270    
271 chrisfen 2594 \newpage
272    
273 chrisfen 2575 \bibliographystyle{achemso}
274     \bibliography{electrostaticMethods}
275    
276    
277     \end{document}