ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2744
Committed: Tue May 2 13:11:41 2006 UTC (18 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 61083 byte(s)
Log Message:
final changes for publication

File Contents

# User Rev Content
1 chrisfen 2575 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 2617 %\documentclass[aps,prb,preprint]{revtex4}
3 chrisfen 2741 \documentclass[11pt]{article}
4     \usepackage{endfloat}
5 chrisfen 2636 \usepackage{amsmath,bm}
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 chrisfen 2742 %\renewcommand{\baselinestretch}{1.2}
24     \renewcommand{\baselinestretch}{2}
25 chrisfen 2575 \renewcommand\citemid{\ } % no comma in optional reference note
26 chrisfen 2742 \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{2}} %doublespace captions
27     \let\Caption\caption
28     \renewcommand\caption[1]{%
29     \Caption[#1]{}%
30     }
31 gezelter 2744 \setlength{\abovecaptionskip}{1.2in}
32     \setlength{\belowcaptionskip}{1.2in}
33 chrisfen 2575
34     \begin{document}
35    
36 chrisfen 2667 \title{Is the Ewald summation still necessary? \\
37 gezelter 2744 Pairwise alternatives to the accepted standard \\
38     for long-range electrostatics}
39 chrisfen 2575
40 gezelter 2617 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
41     gezelter@nd.edu} \\
42 chrisfen 2575 Department of Chemistry and Biochemistry\\
43     University of Notre Dame\\
44     Notre Dame, Indiana 46556}
45    
46     \date{\today}
47    
48     \maketitle
49 chrisfen 2742 %\doublespacing
50 gezelter 2617
51 chrisfen 2575 \begin{abstract}
52 gezelter 2656 We investigate pairwise electrostatic interaction methods and show
53     that there are viable and computationally efficient $(\mathscr{O}(N))$
54     alternatives to the Ewald summation for typical modern molecular
55     simulations. These methods are extended from the damped and
56 chrisfen 2667 cutoff-neutralized Coulombic sum originally proposed by
57     [D. Wolf, P. Keblinski, S.~R. Phillpot, and J. Eggebrecht, {\it J. Chem. Phys.} {\bf 110}, 8255 (1999)] One of these, the damped shifted force method, shows
58 gezelter 2656 a remarkable ability to reproduce the energetic and dynamic
59     characteristics exhibited by simulations employing lattice summation
60     techniques. Comparisons were performed with this and other pairwise
61 chrisfen 2667 methods against the smooth particle mesh Ewald ({\sc spme}) summation
62     to see how well they reproduce the energetics and dynamics of a
63 gezelter 2744 variety of molecular simulations.
64 chrisfen 2575 \end{abstract}
65    
66 gezelter 2617 \newpage
67    
68 chrisfen 2575 %\narrowtext
69    
70 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71 chrisfen 2575 % BODY OF TEXT
72 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73 chrisfen 2575
74     \section{Introduction}
75    
76 gezelter 2617 In molecular simulations, proper accumulation of the electrostatic
77 gezelter 2643 interactions is essential and is one of the most
78     computationally-demanding tasks. The common molecular mechanics force
79     fields represent atomic sites with full or partial charges protected
80     by Lennard-Jones (short range) interactions. This means that nearly
81     every pair interaction involves a calculation of charge-charge forces.
82     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
83     interactions quickly become the most expensive part of molecular
84     simulations. Historically, the electrostatic pair interaction would
85     not have decayed appreciably within the typical box lengths that could
86     be feasibly simulated. In the larger systems that are more typical of
87     modern simulations, large cutoffs should be used to incorporate
88     electrostatics correctly.
89 chrisfen 2604
90 gezelter 2643 There have been many efforts to address the proper and practical
91     handling of electrostatic interactions, and these have resulted in a
92     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
93     typically classified as implicit methods (i.e., continuum dielectrics,
94     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
95     (i.e., Ewald summations, interaction shifting or
96 chrisfen 2640 truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
97 chrisfen 2639 reaction field type methods, fast multipole
98     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
99 gezelter 2643 often preferred because they physically incorporate solvent molecules
100     in the system of interest, but these methods are sometimes difficult
101     to utilize because of their high computational cost.\cite{Roux99} In
102     addition to the computational cost, there have been some questions
103     regarding possible artifacts caused by the inherent periodicity of the
104     explicit Ewald summation.\cite{Tobias01}
105 chrisfen 2639
106 chrisfen 2667 In this paper, we focus on a new set of pairwise methods devised by
107 gezelter 2643 Wolf {\it et al.},\cite{Wolf99} which we further extend. These
108     methods along with a few other mixed methods (i.e. reaction field) are
109     compared with the smooth particle mesh Ewald
110     sum,\cite{Onsager36,Essmann99} which is our reference method for
111     handling long-range electrostatic interactions. The new methods for
112     handling electrostatics have the potential to scale linearly with
113     increasing system size since they involve only a simple modification
114     to the direct pairwise sum. They also lack the added periodicity of
115     the Ewald sum, so they can be used for systems which are non-periodic
116     or which have one- or two-dimensional periodicity. Below, these
117 chrisfen 2740 methods are evaluated using a variety of model systems to
118     establish their usability in molecular simulations.
119 chrisfen 2639
120 chrisfen 2608 \subsection{The Ewald Sum}
121 chrisfen 2667 The complete accumulation of the electrostatic interactions in a system with
122 chrisfen 2639 periodic boundary conditions (PBC) requires the consideration of the
123 gezelter 2643 effect of all charges within a (cubic) simulation box as well as those
124     in the periodic replicas,
125 chrisfen 2636 \begin{equation}
126     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime \left[ \sum_{i=1}^N\sum_{j=1}^N \phi\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) \right],
127     \label{eq:PBCSum}
128     \end{equation}
129 chrisfen 2639 where the sum over $\mathbf{n}$ is a sum over all periodic box
130     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
131     prime indicates $i = j$ are neglected for $\mathbf{n} =
132     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
133     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
134     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
135 gezelter 2643 $j$, and $\phi$ is the solution to Poisson's equation
136     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
137     charge-charge interactions). In the case of monopole electrostatics,
138     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
139     non-neutral systems.
140 chrisfen 2604
141 gezelter 2643 The electrostatic summation problem was originally studied by Ewald
142 chrisfen 2636 for the case of an infinite crystal.\cite{Ewald21}. The approach he
143     took was to convert this conditionally convergent sum into two
144     absolutely convergent summations: a short-ranged real-space summation
145     and a long-ranged reciprocal-space summation,
146     \begin{equation}
147     \begin{split}
148 chrisfen 2637 V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}{|\mathbf{r}_{ij}+\mathbf{n}|} \\ &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
149 chrisfen 2636 \end{split}
150     \label{eq:EwaldSum}
151     \end{equation}
152 chrisfen 2649 where $\alpha$ is the damping or convergence parameter with units of
153     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
154     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
155     constant of the surrounding medium. The final two terms of
156 chrisfen 2636 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
157     for interacting with a surrounding dielectric.\cite{Allen87} This
158     dipolar term was neglected in early applications in molecular
159     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
160     Leeuw {\it et al.} to address situations where the unit cell has a
161 gezelter 2643 dipole moment which is magnified through replication of the periodic
162     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
163     system is said to be using conducting (or ``tin-foil'') boundary
164 chrisfen 2637 conditions, $\epsilon_{\rm S} = \infty$. Figure
165 chrisfen 2636 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
166 gezelter 2653 time. Initially, due to the small system sizes that could be
167     simulated feasibly, the entire simulation box was replicated to
168     convergence. In more modern simulations, the systems have grown large
169     enough that a real-space cutoff could potentially give convergent
170     behavior. Indeed, it has been observed that with the choice of a
171     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
172     rapidly convergent and small relative to the real-space
173     portion.\cite{Karasawa89,Kolafa92}
174 gezelter 2643
175 chrisfen 2610 \begin{figure}
176     \centering
177 gezelter 2656 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
178 gezelter 2669 \caption{The change in the need for the Ewald sum with
179     increasing computational power. A:~Initially, only small systems
180     could be studied, and the Ewald sum replicated the simulation box to
181     convergence. B:~Now, radial cutoff methods should be able to reach
182     convergence for the larger systems of charges that are common today.}
183 chrisfen 2610 \label{fig:ewaldTime}
184     \end{figure}
185    
186 gezelter 2643 The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
187 chrisfen 2649 convergence parameter $(\alpha)$ plays an important role in balancing
188 gezelter 2643 the computational cost between the direct and reciprocal-space
189     portions of the summation. The choice of this value allows one to
190     select whether the real-space or reciprocal space portion of the
191     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
192     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
193     $\alpha$ and thoughtful algorithm development, this cost can be
194     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
195     taken to reduce the cost of the Ewald summation even further is to set
196     $\alpha$ such that the real-space interactions decay rapidly, allowing
197     for a short spherical cutoff. Then the reciprocal space summation is
198     optimized. These optimizations usually involve utilization of the
199     fast Fourier transform (FFT),\cite{Hockney81} leading to the
200 chrisfen 2637 particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
201     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
202     methods, the cost of the reciprocal-space portion of the Ewald
203 gezelter 2643 summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
204     \log N)$.
205 chrisfen 2636
206 gezelter 2643 These developments and optimizations have made the use of the Ewald
207     summation routine in simulations with periodic boundary
208     conditions. However, in certain systems, such as vapor-liquid
209     interfaces and membranes, the intrinsic three-dimensional periodicity
210     can prove problematic. The Ewald sum has been reformulated to handle
211 gezelter 2744 2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
212     methods are computationally expensive.\cite{Spohr97,Yeh99} More
213 chrisfen 2740 recently, there have been several successful efforts toward reducing
214 gezelter 2744 the computational cost of 2-D lattice
215     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
216     bringing them more in line with the cost of the full 3-D summation.
217 chrisfen 2637
218 gezelter 2744
219 chrisfen 2637 Several studies have recognized that the inherent periodicity in the
220 gezelter 2643 Ewald sum can also have an effect on three-dimensional
221     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
222     Solvated proteins are essentially kept at high concentration due to
223     the periodicity of the electrostatic summation method. In these
224 chrisfen 2637 systems, the more compact folded states of a protein can be
225     artificially stabilized by the periodic replicas introduced by the
226 gezelter 2643 Ewald summation.\cite{Weber00} Thus, care must be taken when
227     considering the use of the Ewald summation where the assumed
228     periodicity would introduce spurious effects in the system dynamics.
229 chrisfen 2637
230 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
231 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
232 gezelter 2624 for the accurate accumulation of electrostatic interactions in an
233 gezelter 2643 efficient pairwise fashion. This procedure lacks the inherent
234     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
235     observed that the electrostatic interaction is effectively
236     short-ranged in condensed phase systems and that neutralization of the
237     charge contained within the cutoff radius is crucial for potential
238     stability. They devised a pairwise summation method that ensures
239     charge neutrality and gives results similar to those obtained with the
240 chrisfen 2667 Ewald summation. The resulting shifted Coulomb potential includes
241     image-charges subtracted out through placement on the cutoff sphere
242     and a distance-dependent damping function (identical to that seen in
243     the real-space portion of the Ewald sum) to aid convergence
244 chrisfen 2601 \begin{equation}
245 chrisfen 2640 V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_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\}.
246 chrisfen 2601 \label{eq:WolfPot}
247     \end{equation}
248 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
249     potential. However, neutralizing the charge contained within each
250     cutoff sphere requires the placement of a self-image charge on the
251     surface of the cutoff sphere. This additional self-term in the total
252 gezelter 2624 potential enabled Wolf {\it et al.} to obtain excellent estimates of
253 gezelter 2617 Madelung energies for many crystals.
254    
255     In order to use their charge-neutralized potential in molecular
256     dynamics simulations, Wolf \textit{et al.} suggested taking the
257     derivative of this potential prior to evaluation of the limit. This
258     procedure gives an expression for the forces,
259 chrisfen 2601 \begin{equation}
260 chrisfen 2636 F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{\left[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{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\},
261 chrisfen 2601 \label{eq:WolfForces}
262     \end{equation}
263 gezelter 2617 that incorporates both image charges and damping of the electrostatic
264     interaction.
265    
266     More recently, Zahn \textit{et al.} investigated these potential and
267     force expressions for use in simulations involving water.\cite{Zahn02}
268 gezelter 2624 In their work, they pointed out that the forces and derivative of
269     the potential are not commensurate. Attempts to use both
270 gezelter 2643 eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
271 gezelter 2624 to poor energy conservation. They correctly observed that taking the
272     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
273     derivatives gives forces for a different potential energy function
274 gezelter 2643 than the one shown in eq. (\ref{eq:WolfPot}).
275 gezelter 2617
276 gezelter 2643 Zahn \textit{et al.} introduced a modified form of this summation
277     method as a way to use the technique in Molecular Dynamics
278     simulations. They proposed a new damped Coulomb potential,
279 chrisfen 2601 \begin{equation}
280 gezelter 2643 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\},
281 chrisfen 2601 \label{eq:ZahnPot}
282     \end{equation}
283 gezelter 2643 and showed that this potential does fairly well at capturing the
284 gezelter 2617 structural and dynamic properties of water compared the same
285     properties obtained using the Ewald sum.
286 chrisfen 2601
287 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
288    
289 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
290     al.} are constructed using two different (and separable) computational
291 gezelter 2624 tricks: \begin{enumerate}
292 gezelter 2617 \item shifting through the use of image charges, and
293     \item damping the electrostatic interaction.
294 gezelter 2624 \end{enumerate} Wolf \textit{et al.} treated the
295 gezelter 2617 development of their summation method as a progressive application of
296     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
297     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
298     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
299     both techniques. It is possible, however, to separate these
300     tricks and study their effects independently.
301    
302     Starting with the original observation that the effective range of the
303     electrostatic interaction in condensed phases is considerably less
304     than $r^{-1}$, either the cutoff sphere neutralization or the
305     distance-dependent damping technique could be used as a foundation for
306     a new pairwise summation method. Wolf \textit{et al.} made the
307     observation that charge neutralization within the cutoff sphere plays
308     a significant role in energy convergence; therefore we will begin our
309     analysis with the various shifted forms that maintain this charge
310     neutralization. We can evaluate the methods of Wolf
311     \textit{et al.} and Zahn \textit{et al.} by considering the standard
312     shifted potential,
313 chrisfen 2601 \begin{equation}
314 gezelter 2643 V_\textrm{SP}(r) = \begin{cases}
315 gezelter 2617 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
316     R_\textrm{c}
317     \end{cases},
318     \label{eq:shiftingPotForm}
319     \end{equation}
320     and shifted force,
321     \begin{equation}
322 gezelter 2643 V_\textrm{SF}(r) = \begin{cases}
323 gezelter 2624 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
324 gezelter 2617 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
325 chrisfen 2601 \end{cases},
326 chrisfen 2612 \label{eq:shiftingForm}
327 chrisfen 2601 \end{equation}
328 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
329     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
330     that both the potential and the forces goes to zero at the cutoff
331     radius, while the Shifted Potential ({\sc sp}) form only ensures the
332     potential is smooth at the cutoff radius
333     ($R_\textrm{c}$).\cite{Allen87}
334    
335 gezelter 2624 The forces associated with the shifted potential are simply the forces
336     of the unshifted potential itself (when inside the cutoff sphere),
337 chrisfen 2601 \begin{equation}
338 gezelter 2643 F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
339 chrisfen 2612 \end{equation}
340 gezelter 2624 and are zero outside. Inside the cutoff sphere, the forces associated
341     with the shifted force form can be written,
342 chrisfen 2612 \begin{equation}
343 gezelter 2643 F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
344 gezelter 2624 v(r)}{dr} \right)_{r=R_\textrm{c}}.
345     \end{equation}
346    
347 gezelter 2643 If the potential, $v(r)$, is taken to be the normal Coulomb potential,
348 gezelter 2624 \begin{equation}
349     v(r) = \frac{q_i q_j}{r},
350     \label{eq:Coulomb}
351     \end{equation}
352     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
353     al.}'s undamped prescription:
354     \begin{equation}
355 gezelter 2643 V_\textrm{SP}(r) =
356 gezelter 2624 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
357     r\leqslant R_\textrm{c},
358 chrisfen 2636 \label{eq:SPPot}
359 gezelter 2624 \end{equation}
360     with associated forces,
361     \begin{equation}
362 gezelter 2643 F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
363 chrisfen 2636 \label{eq:SPForces}
364 chrisfen 2612 \end{equation}
365 gezelter 2624 These forces are identical to the forces of the standard Coulomb
366     interaction, and cutting these off at $R_c$ was addressed by Wolf
367     \textit{et al.} as undesirable. They pointed out that the effect of
368     the image charges is neglected in the forces when this form is
369     used,\cite{Wolf99} thereby eliminating any benefit from the method in
370     molecular dynamics. Additionally, there is a discontinuity in the
371     forces at the cutoff radius which results in energy drift during MD
372     simulations.
373 chrisfen 2612
374 gezelter 2624 The shifted force ({\sc sf}) form using the normal Coulomb potential
375     will give,
376 chrisfen 2612 \begin{equation}
377 gezelter 2643 V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
378 chrisfen 2612 \label{eq:SFPot}
379     \end{equation}
380 gezelter 2624 with associated forces,
381 chrisfen 2612 \begin{equation}
382 gezelter 2643 F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
383 chrisfen 2612 \label{eq:SFForces}
384     \end{equation}
385 gezelter 2624 This formulation has the benefits that there are no discontinuities at
386 gezelter 2643 the cutoff radius, while the neutralizing image charges are present in
387     both the energy and force expressions. It would be simple to add the
388     self-neutralizing term back when computing the total energy of the
389 gezelter 2624 system, thereby maintaining the agreement with the Madelung energies.
390     A side effect of this treatment is the alteration in the shape of the
391     potential that comes from the derivative term. Thus, a degree of
392     clarity about agreement with the empirical potential is lost in order
393     to gain functionality in dynamics simulations.
394 chrisfen 2612
395 chrisfen 2620 Wolf \textit{et al.} originally discussed the energetics of the
396 gezelter 2643 shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
397     insufficient for accurate determination of the energy with reasonable
398     cutoff distances. The calculated Madelung energies fluctuated around
399     the expected value as the cutoff radius was increased, but the
400     oscillations converged toward the correct value.\cite{Wolf99} A
401 gezelter 2624 damping function was incorporated to accelerate the convergence; and
402 gezelter 2643 though alternative forms for the damping function could be
403 gezelter 2624 used,\cite{Jones56,Heyes81} the complimentary error function was
404     chosen to mirror the effective screening used in the Ewald summation.
405     Incorporating this error function damping into the simple Coulomb
406     potential,
407 chrisfen 2612 \begin{equation}
408 gezelter 2624 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
409 chrisfen 2601 \label{eq:dampCoulomb}
410     \end{equation}
411 gezelter 2643 the shifted potential (eq. (\ref{eq:SPPot})) becomes
412 chrisfen 2601 \begin{equation}
413 gezelter 2643 V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
414 chrisfen 2612 \label{eq:DSPPot}
415 chrisfen 2629 \end{equation}
416 gezelter 2624 with associated forces,
417 chrisfen 2612 \begin{equation}
418 gezelter 2643 F_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
419 chrisfen 2612 \label{eq:DSPForces}
420     \end{equation}
421 gezelter 2643 Again, this damped shifted potential suffers from a
422     force-discontinuity at the cutoff radius, and the image charges play
423     no role in the forces. To remedy these concerns, one may derive a
424     {\sc sf} variant by including the derivative term in
425     eq. (\ref{eq:shiftingForm}),
426 chrisfen 2612 \begin{equation}
427 chrisfen 2620 \begin{split}
428 gezelter 2643 V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\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-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}.
429 chrisfen 2612 \label{eq:DSFPot}
430 chrisfen 2620 \end{split}
431 chrisfen 2612 \end{equation}
432 chrisfen 2636 The derivative of the above potential will lead to the following forces,
433 chrisfen 2612 \begin{equation}
434 chrisfen 2620 \begin{split}
435 gezelter 2643 F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\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] \quad r\leqslant R_\textrm{c}.
436 chrisfen 2612 \label{eq:DSFForces}
437 chrisfen 2620 \end{split}
438 chrisfen 2612 \end{equation}
439 gezelter 2643 If the damping parameter $(\alpha)$ is set to zero, the undamped case,
440     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
441     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
442 chrisfen 2601
443 chrisfen 2636 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
444     derived by Zahn \textit{et al.}; however, there are two important
445     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
446     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
447     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
448     in the Zahn potential, resulting in a potential discontinuity as
449     particles cross $R_\textrm{c}$. Second, the sign of the derivative
450     portion is different. The missing $v_\textrm{c}$ term would not
451     affect molecular dynamics simulations (although the computed energy
452     would be expected to have sudden jumps as particle distances crossed
453 gezelter 2643 $R_c$). The sign problem is a potential source of errors, however.
454     In fact, it introduces a discontinuity in the forces at the cutoff,
455     because the force function is shifted in the wrong direction and
456     doesn't cross zero at $R_\textrm{c}$.
457 chrisfen 2602
458 gezelter 2624 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
459 gezelter 2643 electrostatic summation method in which the potential and forces are
460     continuous at the cutoff radius and which incorporates the damping
461     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
462     this paper, we will evaluate exactly how good these methods ({\sc sp},
463     {\sc sf}, damping) are at reproducing the correct electrostatic
464     summation performed by the Ewald sum.
465 gezelter 2624
466     \subsection{Other alternatives}
467 gezelter 2643 In addition to the methods described above, we considered some other
468     techniques that are commonly used in molecular simulations. The
469 chrisfen 2629 simplest of these is group-based cutoffs. Though of little use for
470 gezelter 2643 charged molecules, collecting atoms into neutral groups takes
471 chrisfen 2629 advantage of the observation that the electrostatic interactions decay
472     faster than those for monopolar pairs.\cite{Steinbach94} When
473 gezelter 2643 considering these molecules as neutral groups, the relative
474     orientations of the molecules control the strength of the interactions
475     at the cutoff radius. Consequently, as these molecular particles move
476     through $R_\textrm{c}$, the energy will drift upward due to the
477     anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
478     maintain good energy conservation, both the potential and derivative
479     need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
480     This is accomplished using a standard switching function. If a smooth
481     second derivative is desired, a fifth (or higher) order polynomial can
482     be used.\cite{Andrea83}
483 gezelter 2624
484 chrisfen 2629 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
485 gezelter 2643 and to incorporate the effects of the surroundings, a method like
486     Reaction Field ({\sc rf}) can be used. The original theory for {\sc
487     rf} was originally developed by Onsager,\cite{Onsager36} and it was
488     applied in simulations for the study of water by Barker and
489     Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
490     an extension of the group-based cutoff method where the net dipole
491     within the cutoff sphere polarizes an external dielectric, which
492     reacts back on the central dipole. The same switching function
493     considerations for group-based cutoffs need to made for {\sc rf}, with
494     the additional pre-specification of a dielectric constant.
495 gezelter 2624
496 chrisfen 2608 \section{Methods}
497    
498 chrisfen 2620 In classical molecular mechanics simulations, there are two primary
499     techniques utilized to obtain information about the system of
500     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
501     techniques utilize pairwise summations of interactions between
502     particle sites, but they use these summations in different ways.
503 chrisfen 2608
504 gezelter 2645 In MC, the potential energy difference between configurations dictates
505     the progression of MC sampling. Going back to the origins of this
506     method, the acceptance criterion for the canonical ensemble laid out
507     by Metropolis \textit{et al.} states that a subsequent configuration
508     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
509     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
510     Maintaining the correct $\Delta E$ when using an alternate method for
511     handling the long-range electrostatics will ensure proper sampling
512     from the ensemble.
513 chrisfen 2608
514 gezelter 2624 In MD, the derivative of the potential governs how the system will
515 chrisfen 2620 progress in time. Consequently, the force and torque vectors on each
516 gezelter 2624 body in the system dictate how the system evolves. If the magnitude
517     and direction of these vectors are similar when using alternate
518     electrostatic summation techniques, the dynamics in the short term
519     will be indistinguishable. Because error in MD calculations is
520     cumulative, one should expect greater deviation at longer times,
521     although methods which have large differences in the force and torque
522     vectors will diverge from each other more rapidly.
523 chrisfen 2608
524 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
525 gezelter 2645
526 gezelter 2624 The pairwise summation techniques (outlined in section
527     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
528     studying the energy differences between conformations. We took the
529 gezelter 2656 {\sc spme}-computed energy difference between two conformations to be the
530 gezelter 2624 correct behavior. An ideal performance by an alternative method would
531 gezelter 2645 reproduce these energy differences exactly (even if the absolute
532     energies calculated by the methods are different). Since none of the
533     methods provide exact energy differences, we used linear least squares
534     regressions of energy gap data to evaluate how closely the methods
535     mimicked the Ewald energy gaps. Unitary results for both the
536     correlation (slope) and correlation coefficient for these regressions
537 gezelter 2656 indicate perfect agreement between the alternative method and {\sc spme}.
538 gezelter 2645 Sample correlation plots for two alternate methods are shown in
539     Fig. \ref{fig:linearFit}.
540 chrisfen 2608
541 chrisfen 2609 \begin{figure}
542     \centering
543 chrisfen 2741 \includegraphics[width = \linewidth]{./dualLinear.pdf}
544 gezelter 2645 \caption{Example least squares regressions of the configuration energy
545     differences for SPC/E water systems. The upper plot shows a data set
546     with a poor correlation coefficient ($R^2$), while the lower plot
547     shows a data set with a good correlation coefficient.}
548     \label{fig:linearFit}
549 chrisfen 2609 \end{figure}
550    
551 chrisfen 2740 Each of the seven system types (detailed in section \ref{sec:RepSims})
552     were represented using 500 independent configurations. Thus, each of
553     the alternative (non-Ewald) electrostatic summation methods was
554     evaluated using an accumulated 873,250 configurational energy
555     differences.
556 chrisfen 2609
557 gezelter 2624 Results and discussion for the individual analysis of each of the
558 chrisfen 2742 system types appear in the supporting information,\cite{EPAPSdeposit}
559     while the cumulative results over all the investigated systems appears
560     below in section \ref{sec:EnergyResults}.
561 gezelter 2624
562 chrisfen 2609 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
563 gezelter 2624 We evaluated the pairwise methods (outlined in section
564     \ref{sec:ESMethods}) for use in MD simulations by
565     comparing the force and torque vectors with those obtained using the
566 gezelter 2656 reference Ewald summation ({\sc spme}). Both the magnitude and the
567 gezelter 2624 direction of these vectors on each of the bodies in the system were
568     analyzed. For the magnitude of these vectors, linear least squares
569     regression analyses were performed as described previously for
570     comparing $\Delta E$ values. Instead of a single energy difference
571     between two system configurations, we compared the magnitudes of the
572     forces (and torques) on each molecule in each configuration. For a
573     system of 1000 water molecules and 40 ions, there are 1040 force
574     vectors and 1000 torque vectors. With 500 configurations, this
575     results in 520,000 force and 500,000 torque vector comparisons.
576     Additionally, data from seven different system types was aggregated
577     before the comparison was made.
578 chrisfen 2609
579 gezelter 2624 The {\it directionality} of the force and torque vectors was
580     investigated through measurement of the angle ($\theta$) formed
581 gezelter 2656 between those computed from the particular method and those from {\sc spme},
582 chrisfen 2610 \begin{equation}
583 gezelter 2645 \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
584 chrisfen 2610 \end{equation}
585 gezelter 2656 where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
586 chrisfen 2651 vector computed using method M. Each of these $\theta$ values was
587     accumulated in a distribution function and weighted by the area on the
588 chrisfen 2652 unit sphere. Since this distribution is a measure of angular error
589     between two different electrostatic summation methods, there is no
590     {\it a priori} reason for the profile to adhere to any specific
591     shape. Thus, gaussian fits were used to measure the width of the
592 chrisfen 2667 resulting distributions. The variance ($\sigma^2$) was extracted from
593     each of these fits and was used to compare distribution widths.
594     Values of $\sigma^2$ near zero indicate vector directions
595     indistinguishable from those calculated when using the reference
596     method ({\sc spme}).
597 gezelter 2624
598     \subsection{Short-time Dynamics}
599 gezelter 2645
600     The effects of the alternative electrostatic summation methods on the
601     short-time dynamics of charged systems were evaluated by considering a
602     NaCl crystal at a temperature of 1000 K. A subset of the best
603     performing pairwise methods was used in this comparison. The NaCl
604     crystal was chosen to avoid possible complications from the treatment
605     of orientational motion in molecular systems. All systems were
606     started with the same initial positions and velocities. Simulations
607     were performed under the microcanonical ensemble, and velocity
608 chrisfen 2638 autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
609     of the trajectories,
610 chrisfen 2609 \begin{equation}
611 gezelter 2656 C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
612 chrisfen 2609 \label{eq:vCorr}
613     \end{equation}
614 chrisfen 2638 Velocity autocorrelation functions require detailed short time data,
615     thus velocity information was saved every 2 fs over 10 ps
616     trajectories. Because the NaCl crystal is composed of two different
617     atom types, the average of the two resulting velocity autocorrelation
618     functions was used for comparisons.
619    
620     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
621 gezelter 2645
622     The effects of the same subset of alternative electrostatic methods on
623     the {\it long-time} dynamics of charged systems were evaluated using
624 chrisfen 2667 the same model system (NaCl crystals at 1000~K). The power spectrum
625 gezelter 2645 ($I(\omega)$) was obtained via Fourier transform of the velocity
626     autocorrelation function, \begin{equation} I(\omega) =
627     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
628 chrisfen 2609 \label{eq:powerSpec}
629     \end{equation}
630 chrisfen 2638 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
631     NaCl crystal is composed of two different atom types, the average of
632 gezelter 2645 the two resulting power spectra was used for comparisons. Simulations
633     were performed under the microcanonical ensemble, and velocity
634 chrisfen 2740 information was saved every 5~fs over 100~ps trajectories.
635 chrisfen 2609
636     \subsection{Representative Simulations}\label{sec:RepSims}
637 chrisfen 2740 A variety of representative molecular simulations were analyzed to
638     determine the relative effectiveness of the pairwise summation
639     techniques in reproducing the energetics and dynamics exhibited by
640     {\sc spme}. We wanted to span the space of typical molecular
641     simulations (i.e. from liquids of neutral molecules to ionic
642     crystals), so the systems studied were:
643 chrisfen 2599 \begin{enumerate}
644 gezelter 2645 \item liquid water (SPC/E),\cite{Berendsen87}
645     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
646     \item NaCl crystals,
647     \item NaCl melts,
648     \item a low ionic strength solution of NaCl in water (0.11 M),
649     \item a high ionic strength solution of NaCl in water (1.1 M), and
650     \item a 6 \AA\ radius sphere of Argon in water.
651 chrisfen 2599 \end{enumerate}
652 chrisfen 2620 By utilizing the pairwise techniques (outlined in section
653     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
654 gezelter 2645 charged particles, and mixtures of the two, we hope to discern under
655     which conditions it will be possible to use one of the alternative
656     summation methodologies instead of the Ewald sum.
657 chrisfen 2586
658 gezelter 2645 For the solid and liquid water configurations, configurations were
659     taken at regular intervals from high temperature trajectories of 1000
660     SPC/E water molecules. Each configuration was equilibrated
661     independently at a lower temperature (300~K for the liquid, 200~K for
662     the crystal). The solid and liquid NaCl systems consisted of 500
663     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
664     these systems were selected and equilibrated in the same manner as the
665 chrisfen 2667 water systems. In order to introduce measurable fluctuations in the
666     configuration energy differences, the crystalline simulations were
667     equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid
668     NaCl configurations needed to represent a fully disordered array of
669     point charges, so the high temperature of 7000~K was selected for
670     equilibration. The ionic solutions were made by solvating 4 (or 40)
671     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
672     water positions were then randomly swapped, and the resulting
673     configurations were again equilibrated individually. Finally, for the
674     Argon / Water ``charge void'' systems, the identities of all the SPC/E
675     waters within 6 \AA\ of the center of the equilibrated water
676     configurations were converted to argon.
677 chrisfen 2586
678 gezelter 2645 These procedures guaranteed us a set of representative configurations
679 gezelter 2653 from chemically-relevant systems sampled from appropriate
680     ensembles. Force field parameters for the ions and Argon were taken
681 gezelter 2645 from the force field utilized by {\sc oopse}.\cite{Meineke05}
682    
683     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
684     We compared the following alternative summation methods with results
685 gezelter 2656 from the reference method ({\sc spme}):
686 gezelter 2645 \begin{itemize}
687     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
688     and 0.3 \AA$^{-1}$,
689     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
690     and 0.3 \AA$^{-1}$,
691     \item reaction field with an infinite dielectric constant, and
692     \item an unmodified cutoff.
693     \end{itemize}
694     Group-based cutoffs with a fifth-order polynomial switching function
695     were utilized for the reaction field simulations. Additionally, we
696 gezelter 2656 investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
697     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
698     implementation of {\sc spme},\cite{Ponder87} while all other calculations
699 gezelter 2653 were performed using the {\sc oopse} molecular mechanics
700 gezelter 2645 package.\cite{Meineke05} All other portions of the energy calculation
701     (i.e. Lennard-Jones interactions) were handled in exactly the same
702     manner across all systems and configurations.
703 chrisfen 2586
704 chrisfen 2667 The alternative methods were also evaluated with three different
705 chrisfen 2649 cutoff radii (9, 12, and 15 \AA). As noted previously, the
706     convergence parameter ($\alpha$) plays a role in the balance of the
707     real-space and reciprocal-space portions of the Ewald calculation.
708     Typical molecular mechanics packages set this to a value dependent on
709     the cutoff radius and a tolerance (typically less than $1 \times
710     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
711 gezelter 2653 increasing accuracy at the expense of computational time spent on the
712     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
713 gezelter 2656 The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
714     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
715 gezelter 2653 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
716     respectively.
717 chrisfen 2609
718 chrisfen 2575 \section{Results and Discussion}
719    
720 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
721 chrisfen 2620 In order to evaluate the performance of the pairwise electrostatic
722     summation methods for Monte Carlo simulations, the energy differences
723     between configurations were compared to the values obtained when using
724 gezelter 2656 {\sc spme}. The results for the subsequent regression analysis are shown in
725 chrisfen 2620 figure \ref{fig:delE}.
726 chrisfen 2590
727     \begin{figure}
728     \centering
729 chrisfen 2741 \includegraphics[width=5.5in]{./delEplot.pdf}
730 gezelter 2645 \caption{Statistical analysis of the quality of configurational energy
731     differences for a given electrostatic method compared with the
732     reference Ewald sum. Results with a value equal to 1 (dashed line)
733     indicate $\Delta E$ values indistinguishable from those obtained using
734 gezelter 2656 {\sc spme}. Different values of the cutoff radius are indicated with
735 gezelter 2645 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
736     inverted triangles).}
737 chrisfen 2601 \label{fig:delE}
738 chrisfen 2594 \end{figure}
739    
740 gezelter 2645 The most striking feature of this plot is how well the Shifted Force
741     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
742     differences. For the undamped {\sc sf} method, and the
743     moderately-damped {\sc sp} methods, the results are nearly
744     indistinguishable from the Ewald results. The other common methods do
745     significantly less well.
746 chrisfen 2594
747 gezelter 2645 The unmodified cutoff method is essentially unusable. This is not
748     surprising since hard cutoffs give large energy fluctuations as atoms
749     or molecules move in and out of the cutoff
750     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
751     some degree by using group based cutoffs with a switching
752     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
753     significant improvement using the group-switched cutoff because the
754     salt and salt solution systems contain non-neutral groups. Interested
755     readers can consult the accompanying supporting information for a
756 chrisfen 2742 comparison where all groups are neutral.\cite{EPAPSdeposit}
757 gezelter 2645
758 gezelter 2653 For the {\sc sp} method, inclusion of electrostatic damping improves
759     the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
760 gezelter 2656 shows an excellent correlation and quality of fit with the {\sc spme}
761 gezelter 2653 results, particularly with a cutoff radius greater than 12
762 gezelter 2645 \AA . Use of a larger damping parameter is more helpful for the
763     shortest cutoff shown, but it has a detrimental effect on simulations
764     with larger cutoffs.
765 chrisfen 2609
766 gezelter 2653 In the {\sc sf} sets, increasing damping results in progressively {\it
767     worse} correlation with Ewald. Overall, the undamped case is the best
768 gezelter 2645 performing set, as the correlation and quality of fits are
769     consistently superior regardless of the cutoff distance. The undamped
770     case is also less computationally demanding (because no evaluation of
771     the complementary error function is required).
772    
773     The reaction field results illustrates some of that method's
774     limitations, primarily that it was developed for use in homogenous
775     systems; although it does provide results that are an improvement over
776     those from an unmodified cutoff.
777    
778 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
779 chrisfen 2599
780 chrisfen 2620 Evaluation of pairwise methods for use in Molecular Dynamics
781     simulations requires consideration of effects on the forces and
782 gezelter 2653 torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
783     regression results for the force and torque vector magnitudes,
784     respectively. The data in these figures was generated from an
785     accumulation of the statistics from all of the system types.
786 chrisfen 2594
787     \begin{figure}
788     \centering
789 chrisfen 2741 \includegraphics[width=5.5in]{./frcMagplot.pdf}
790 chrisfen 2651 \caption{Statistical analysis of the quality of the force vector
791     magnitudes for a given electrostatic method compared with the
792     reference Ewald sum. Results with a value equal to 1 (dashed line)
793     indicate force magnitude values indistinguishable from those obtained
794 gezelter 2656 using {\sc spme}. Different values of the cutoff radius are indicated with
795 chrisfen 2651 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
796     inverted triangles).}
797 chrisfen 2601 \label{fig:frcMag}
798 chrisfen 2594 \end{figure}
799    
800 gezelter 2653 Again, it is striking how well the Shifted Potential and Shifted Force
801 gezelter 2656 methods are doing at reproducing the {\sc spme} forces. The undamped and
802 gezelter 2653 weakly-damped {\sc sf} method gives the best agreement with Ewald.
803     This is perhaps expected because this method explicitly incorporates a
804     smooth transition in the forces at the cutoff radius as well as the
805     neutralizing image charges.
806    
807 chrisfen 2620 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
808     in the previous $\Delta E$ section. The unmodified cutoff results are
809     poor, but using group based cutoffs and a switching function provides
810 gezelter 2653 an improvement much more significant than what was seen with $\Delta
811     E$.
812    
813     With moderate damping and a large enough cutoff radius, the {\sc sp}
814     method is generating usable forces. Further increases in damping,
815 chrisfen 2620 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
816 gezelter 2653 detrimental to simulations with larger cutoff radii.
817    
818     The reaction field results are surprisingly good, considering the poor
819 chrisfen 2620 quality of the fits for the $\Delta E$ results. There is still a
820 gezelter 2653 considerable degree of scatter in the data, but the forces correlate
821     well with the Ewald forces in general. We note that the reaction
822     field calculations do not include the pure NaCl systems, so these
823 chrisfen 2620 results are partly biased towards conditions in which the method
824     performs more favorably.
825 chrisfen 2594
826     \begin{figure}
827     \centering
828 chrisfen 2741 \includegraphics[width=5.5in]{./trqMagplot.pdf}
829 chrisfen 2651 \caption{Statistical analysis of the quality of the torque vector
830     magnitudes for a given electrostatic method compared with the
831     reference Ewald sum. Results with a value equal to 1 (dashed line)
832     indicate torque magnitude values indistinguishable from those obtained
833 gezelter 2656 using {\sc spme}. Different values of the cutoff radius are indicated with
834 chrisfen 2651 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
835     inverted triangles).}
836 chrisfen 2601 \label{fig:trqMag}
837 chrisfen 2594 \end{figure}
838    
839 gezelter 2653 Molecular torques were only available from the systems which contained
840     rigid molecules (i.e. the systems containing water). The data in
841     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
842 chrisfen 2594
843 gezelter 2653 Torques appear to be much more sensitive to charges at a longer
844     distance. The striking feature in comparing the new electrostatic
845 gezelter 2656 methods with {\sc spme} is how much the agreement improves with increasing
846 gezelter 2653 cutoff radius. Again, the weakly damped and undamped {\sc sf} method
847 gezelter 2656 appears to be reproducing the {\sc spme} torques most accurately.
848 gezelter 2653
849     Water molecules are dipolar, and the reaction field method reproduces
850     the effect of the surrounding polarized medium on each of the
851     molecular bodies. Therefore it is not surprising that reaction field
852     performs best of all of the methods on molecular torques.
853    
854 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
855 chrisfen 2599
856 gezelter 2653 It is clearly important that a new electrostatic method can reproduce
857     the magnitudes of the force and torque vectors obtained via the Ewald
858     sum. However, the {\it directionality} of these vectors will also be
859     vital in calculating dynamical quantities accurately. Force and
860     torque directionalities were investigated by measuring the angles
861     formed between these vectors and the same vectors calculated using
862 gezelter 2656 {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
863 gezelter 2653 variance ($\sigma^2$) of the Gaussian fits of the angle error
864     distributions of the combined set over all system types.
865 chrisfen 2594
866     \begin{figure}
867     \centering
868 chrisfen 2741 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
869 gezelter 2653 \caption{Statistical analysis of the width of the angular distribution
870     that the force and torque vectors from a given electrostatic method
871     make with their counterparts obtained using the reference Ewald sum.
872     Results with a variance ($\sigma^2$) equal to zero (dashed line)
873     indicate force and torque directions indistinguishable from those
874 gezelter 2656 obtained using {\sc spme}. Different values of the cutoff radius are
875 gezelter 2653 indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
876     and 15\AA\ = inverted triangles).}
877 chrisfen 2601 \label{fig:frcTrqAng}
878 chrisfen 2594 \end{figure}
879    
880 chrisfen 2620 Both the force and torque $\sigma^2$ results from the analysis of the
881     total accumulated system data are tabulated in figure
882 gezelter 2653 \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
883 gezelter 2656 sp}) method would be essentially unusable for molecular dynamics
884     unless the damping function is added. The Shifted Force ({\sc sf})
885     method, however, is generating force and torque vectors which are
886     within a few degrees of the Ewald results even with weak (or no)
887     damping.
888 chrisfen 2594
889 gezelter 2653 All of the sets (aside from the over-damped case) show the improvement
890     afforded by choosing a larger cutoff radius. Increasing the cutoff
891     from 9 to 12 \AA\ typically results in a halving of the width of the
892 gezelter 2656 distribution, with a similar improvement when going from 12 to 15
893 gezelter 2653 \AA .
894    
895     The undamped {\sc sf}, group-based cutoff, and reaction field methods
896     all do equivalently well at capturing the direction of both the force
897 gezelter 2656 and torque vectors. Using the electrostatic damping improves the
898     angular behavior significantly for the {\sc sp} and moderately for the
899     {\sc sf} methods. Overdamping is detrimental to both methods. Again
900     it is important to recognize that the force vectors cover all
901     particles in all seven systems, while torque vectors are only
902     available for neutral molecular groups. Damping is more beneficial to
903 gezelter 2653 charged bodies, and this observation is investigated further in the
904 chrisfen 2742 accompanying supporting information.\cite{EPAPSdeposit}
905 gezelter 2653
906     Although not discussed previously, group based cutoffs can be applied
907 gezelter 2656 to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
908     will reintroduce small discontinuities at the cutoff radius, but the
909     effects of these can be minimized by utilizing a switching function.
910     Though there are no significant benefits or drawbacks observed in
911     $\Delta E$ and the force and torque magnitudes when doing this, there
912     is a measurable improvement in the directionality of the forces and
913     torques. Table \ref{tab:groupAngle} shows the angular variances
914     obtained using group based cutoffs along with the results seen in
915     figure \ref{fig:frcTrqAng}. The {\sc sp} (with an $\alpha$ of 0.2
916     \AA$^{-1}$ or smaller) shows much narrower angular distributions when
917     using group-based cutoffs. The {\sc sf} method likewise shows
918     improvement in the undamped and lightly damped cases.
919 gezelter 2653
920 chrisfen 2595 \begin{table}[htbp]
921 gezelter 2656 \centering
922     \caption{Statistical analysis of the angular
923     distributions that the force (upper) and torque (lower) vectors
924     from a given electrostatic method make with their counterparts
925     obtained using the reference Ewald sum. Calculations were
926     performed both with (Y) and without (N) group based cutoffs and a
927     switching function. The $\alpha$ values have units of \AA$^{-1}$
928     and the variance values have units of degrees$^2$.}
929    
930 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
931 chrisfen 2595 \\
932     \toprule
933     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
934     \cmidrule(lr){3-6}
935     \cmidrule(l){7-10}
936 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
937 chrisfen 2595 \midrule
938 chrisfen 2599
939     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
940     & \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} \\
941     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
942     & \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} \\
943     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
944     & \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} \\
945 chrisfen 2594
946 chrisfen 2595 \midrule
947    
948 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
949     & \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} \\
950     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
951     & \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} \\
952     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
953     & \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} \\
954 chrisfen 2595
955     \bottomrule
956     \end{tabular}
957 chrisfen 2601 \label{tab:groupAngle}
958 chrisfen 2595 \end{table}
959    
960 gezelter 2656 One additional trend in table \ref{tab:groupAngle} is that the
961     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
962     increases, something that is more obvious with group-based cutoffs.
963     The complimentary error function inserted into the potential weakens
964     the electrostatic interaction as the value of $\alpha$ is increased.
965     However, at larger values of $\alpha$, it is possible to overdamp the
966     electrostatic interaction and to remove it completely. Kast
967 gezelter 2653 \textit{et al.} developed a method for choosing appropriate $\alpha$
968     values for these types of electrostatic summation methods by fitting
969     to $g(r)$ data, and their methods indicate optimal values of 0.34,
970     0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
971     respectively.\cite{Kast03} These appear to be reasonable choices to
972     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
973     these findings, choices this high would introduce error in the
974 gezelter 2656 molecular torques, particularly for the shorter cutoffs. Based on our
975     observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
976     but damping may be unnecessary when using the {\sc sf} method.
977 chrisfen 2595
978 chrisfen 2638 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
979 chrisfen 2601
980 gezelter 2653 Zahn {\it et al.} investigated the structure and dynamics of water
981     using eqs. (\ref{eq:ZahnPot}) and
982     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
983     that a method similar (but not identical with) the damped {\sc sf}
984     method resulted in properties very similar to those obtained when
985     using the Ewald summation. The properties they studied (pair
986     distribution functions, diffusion constants, and velocity and
987     orientational correlation functions) may not be particularly sensitive
988     to the long-range and collective behavior that governs the
989 gezelter 2656 low-frequency behavior in crystalline systems. Additionally, the
990     ionic crystals are the worst case scenario for the pairwise methods
991     because they lack the reciprocal space contribution contained in the
992     Ewald summation.
993 chrisfen 2601
994 gezelter 2653 We are using two separate measures to probe the effects of these
995     alternative electrostatic methods on the dynamics in crystalline
996     materials. For short- and intermediate-time dynamics, we are
997     computing the velocity autocorrelation function, and for long-time
998     and large length-scale collective motions, we are looking at the
999     low-frequency portion of the power spectrum.
1000    
1001 chrisfen 2601 \begin{figure}
1002     \centering
1003 chrisfen 2741 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
1004 gezelter 2656 \caption{Velocity autocorrelation functions of NaCl crystals at
1005     1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1006 gezelter 2653 sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
1007     the first minimum. The times to first collision are nearly identical,
1008     but differences can be seen in the peaks and troughs, where the
1009     undamped and weakly damped methods are stiffer than the moderately
1010 gezelter 2656 damped and {\sc spme} methods.}
1011 chrisfen 2638 \label{fig:vCorrPlot}
1012     \end{figure}
1013    
1014 gezelter 2656 The short-time decay of the velocity autocorrelation function through
1015     the first collision are nearly identical in figure
1016     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1017     how the methods differ. The undamped {\sc sf} method has deeper
1018     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1019     any of the other methods. As the damping parameter ($\alpha$) is
1020     increased, these peaks are smoothed out, and the {\sc sf} method
1021     approaches the {\sc spme} results. With $\alpha$ values of 0.2 \AA$^{-1}$,
1022     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1023     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1024     and {\sc sp} potentials become nearly identical with increased
1025     damping. However, this appears to indicate that once damping is
1026     utilized, the details of the form of the potential (and forces)
1027     constructed out of the damped electrostatic interaction are less
1028     important.
1029 chrisfen 2638
1030     \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1031    
1032 gezelter 2656 To evaluate how the differences between the methods affect the
1033     collective long-time motion, we computed power spectra from long-time
1034     traces of the velocity autocorrelation function. The power spectra for
1035     the best-performing alternative methods are shown in
1036     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1037     a cubic switching function between 40 and 50 ps was used to reduce the
1038     ringing resulting from data truncation. This procedure had no
1039     noticeable effect on peak location or magnitude.
1040 chrisfen 2638
1041     \begin{figure}
1042     \centering
1043 chrisfen 2741 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1044 chrisfen 2651 \caption{Power spectra obtained from the velocity auto-correlation
1045 gezelter 2656 functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf}
1046     ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset
1047     shows the frequency region below 100 cm$^{-1}$ to highlight where the
1048     spectra differ.}
1049 chrisfen 2610 \label{fig:methodPS}
1050 chrisfen 2601 \end{figure}
1051    
1052 gezelter 2656 While the high frequency regions of the power spectra for the
1053     alternative methods are quantitatively identical with Ewald spectrum,
1054     the low frequency region shows how the summation methods differ.
1055     Considering the low-frequency inset (expanded in the upper frame of
1056     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1057     correlated motions are blue-shifted when using undamped or weakly
1058     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1059     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1060     correlated motion to the Ewald method (which has a convergence
1061     parameter of 0.3119 \AA$^{-1}$). This weakening of the electrostatic
1062     interaction with increased damping explains why the long-ranged
1063     correlated motions are at lower frequencies for the moderately damped
1064     methods than for undamped or weakly damped methods.
1065    
1066     To isolate the role of the damping constant, we have computed the
1067     spectra for a single method ({\sc sf}) with a range of damping
1068     constants and compared this with the {\sc spme} spectrum.
1069     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1070     electrostatic damping red-shifts the lowest frequency phonon modes.
1071     However, even without any electrostatic damping, the {\sc sf} method
1072     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1073     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1074     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1075     Most} of the collective behavior in the crystal is accurately captured
1076     using the {\sc sf} method. Quantitative agreement with Ewald can be
1077     obtained using moderate damping in addition to the shifting at the
1078     cutoff distance.
1079    
1080 chrisfen 2601 \begin{figure}
1081     \centering
1082 chrisfen 2741 \includegraphics[width = \linewidth]{./increasedDamping.pdf}
1083 gezelter 2656 \caption{Effect of damping on the two lowest-frequency phonon modes in
1084 chrisfen 2667 the NaCl crystal at 1000~K. The undamped shifted force ({\sc sf})
1085 gezelter 2656 method is off by less than 10 cm$^{-1}$, and increasing the
1086     electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1087     with the power spectrum obtained using the Ewald sum. Overdamping can
1088     result in underestimates of frequencies of the long-wavelength
1089     motions.}
1090 chrisfen 2601 \label{fig:dampInc}
1091     \end{figure}
1092    
1093 chrisfen 2575 \section{Conclusions}
1094    
1095 chrisfen 2620 This investigation of pairwise electrostatic summation techniques
1096 gezelter 2656 shows that there are viable and computationally efficient alternatives
1097     to the Ewald summation. These methods are derived from the damped and
1098     cutoff-neutralized Coulombic sum originally proposed by Wolf
1099     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1100     method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1101     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1102     energetic and dynamic characteristics exhibited by simulations
1103     employing lattice summation techniques. The cumulative energy
1104     difference results showed the undamped {\sc sf} and moderately damped
1105     {\sc sp} methods produced results nearly identical to {\sc spme}. Similarly
1106     for the dynamic features, the undamped or moderately damped {\sc sf}
1107     and moderately damped {\sc sp} methods produce force and torque vector
1108     magnitude and directions very similar to the expected values. These
1109     results translate into long-time dynamic behavior equivalent to that
1110     produced in simulations using {\sc spme}.
1111 chrisfen 2604
1112 gezelter 2656 As in all purely-pairwise cutoff methods, these methods are expected
1113     to scale approximately {\it linearly} with system size, and they are
1114     easily parallelizable. This should result in substantial reductions
1115     in the computational cost of performing large simulations.
1116    
1117 chrisfen 2620 Aside from the computational cost benefit, these techniques have
1118     applicability in situations where the use of the Ewald sum can prove
1119 gezelter 2656 problematic. Of greatest interest is their potential use in
1120     interfacial systems, where the unmodified lattice sum techniques
1121     artificially accentuate the periodicity of the system in an
1122     undesirable manner. There have been alterations to the standard Ewald
1123     techniques, via corrections and reformulations, to compensate for
1124     these systems; but the pairwise techniques discussed here require no
1125     modifications, making them natural tools to tackle these problems.
1126     Additionally, this transferability gives them benefits over other
1127     pairwise methods, like reaction field, because estimations of physical
1128     properties (e.g. the dielectric constant) are unnecessary.
1129 chrisfen 2605
1130 gezelter 2656 If a researcher is using Monte Carlo simulations of large chemical
1131     systems containing point charges, most structural features will be
1132     accurately captured using the undamped {\sc sf} method or the {\sc sp}
1133     method with an electrostatic damping of 0.2 \AA$^{-1}$. These methods
1134     would also be appropriate for molecular dynamics simulations where the
1135     data of interest is either structural or short-time dynamical
1136     quantities. For long-time dynamics and collective motions, the safest
1137     pairwise method we have evaluated is the {\sc sf} method with an
1138     electrostatic damping between 0.2 and 0.25
1139     \AA$^{-1}$.
1140 chrisfen 2605
1141 gezelter 2656 We are not suggesting that there is any flaw with the Ewald sum; in
1142     fact, it is the standard by which these simple pairwise sums have been
1143     judged. However, these results do suggest that in the typical
1144     simulations performed today, the Ewald summation may no longer be
1145     required to obtain the level of accuracy most researchers have come to
1146     expect.
1147    
1148 chrisfen 2575 \section{Acknowledgments}
1149 gezelter 2656 Support for this project was provided by the National Science
1150     Foundation under grant CHE-0134881. The authors would like to thank
1151     Steve Corcelli and Ed Maginn for helpful discussions and comments.
1152    
1153 chrisfen 2594 \newpage
1154    
1155 gezelter 2617 \bibliographystyle{jcp2}
1156 chrisfen 2575 \bibliography{electrostaticMethods}
1157    
1158    
1159     \end{document}