ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2659
Committed: Wed Mar 22 21:20:40 2006 UTC (18 years, 5 months ago) by chrisfen
Content type: application/x-tex
File size: 61504 byte(s)
Log Message:
added some files and updated a figure

File Contents

# User Rev Content
1 chrisfen 2575 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 2617 %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4 chrisfen 2575 \usepackage{endfloat}
5 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     \renewcommand{\baselinestretch}{1.2}
24     \renewcommand\citemid{\ } % no comma in optional reference note
25    
26     \begin{document}
27    
28 gezelter 2656 \title{Is the Ewald Summation necessary? \\
29     Pairwise alternatives to the accepted standard for \\
30     long-range electrostatics}
31 chrisfen 2575
32 gezelter 2617 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
33     gezelter@nd.edu} \\
34 chrisfen 2575 Department of Chemistry and Biochemistry\\
35     University of Notre Dame\\
36     Notre Dame, Indiana 46556}
37    
38     \date{\today}
39    
40     \maketitle
41 gezelter 2617 \doublespacing
42    
43 chrisfen 2605 \nobibliography{}
44 chrisfen 2575 \begin{abstract}
45 gezelter 2656 We investigate pairwise electrostatic interaction methods and show
46     that there are viable and computationally efficient $(\mathscr{O}(N))$
47     alternatives to the Ewald summation for typical modern molecular
48     simulations. These methods are extended from the damped and
49     cutoff-neutralized Coulombic sum originally proposed by Wolf
50     \textit{et al.} One of these, the damped shifted force method, shows
51     a remarkable ability to reproduce the energetic and dynamic
52     characteristics exhibited by simulations employing lattice summation
53     techniques. Comparisons were performed with this and other pairwise
54     methods against the smooth particle mesh Ewald ({\sc spme}) summation to see
55     how well they reproduce the energetics and dynamics of a variety of
56     simulation types.
57 chrisfen 2575 \end{abstract}
58    
59 gezelter 2617 \newpage
60    
61 chrisfen 2575 %\narrowtext
62    
63 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64 chrisfen 2575 % BODY OF TEXT
65 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66 chrisfen 2575
67     \section{Introduction}
68    
69 gezelter 2617 In molecular simulations, proper accumulation of the electrostatic
70 gezelter 2643 interactions is essential and is one of the most
71     computationally-demanding tasks. The common molecular mechanics force
72     fields represent atomic sites with full or partial charges protected
73     by Lennard-Jones (short range) interactions. This means that nearly
74     every pair interaction involves a calculation of charge-charge forces.
75     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
76     interactions quickly become the most expensive part of molecular
77     simulations. Historically, the electrostatic pair interaction would
78     not have decayed appreciably within the typical box lengths that could
79     be feasibly simulated. In the larger systems that are more typical of
80     modern simulations, large cutoffs should be used to incorporate
81     electrostatics correctly.
82 chrisfen 2604
83 gezelter 2643 There have been many efforts to address the proper and practical
84     handling of electrostatic interactions, and these have resulted in a
85     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
86     typically classified as implicit methods (i.e., continuum dielectrics,
87     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
88     (i.e., Ewald summations, interaction shifting or
89 chrisfen 2640 truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
90 chrisfen 2639 reaction field type methods, fast multipole
91     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
92 gezelter 2643 often preferred because they physically incorporate solvent molecules
93     in the system of interest, but these methods are sometimes difficult
94     to utilize because of their high computational cost.\cite{Roux99} In
95     addition to the computational cost, there have been some questions
96     regarding possible artifacts caused by the inherent periodicity of the
97     explicit Ewald summation.\cite{Tobias01}
98 chrisfen 2639
99 gezelter 2643 In this paper, we focus on a new set of shifted methods devised by
100     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
101     methods along with a few other mixed methods (i.e. reaction field) are
102     compared with the smooth particle mesh Ewald
103     sum,\cite{Onsager36,Essmann99} which is our reference method for
104     handling long-range electrostatic interactions. The new methods for
105     handling electrostatics have the potential to scale linearly with
106     increasing system size since they involve only a simple modification
107     to the direct pairwise sum. They also lack the added periodicity of
108     the Ewald sum, so they can be used for systems which are non-periodic
109     or which have one- or two-dimensional periodicity. Below, these
110     methods are evaluated using a variety of model systems to establish
111 chrisfen 2640 their usability in molecular simulations.
112 chrisfen 2639
113 chrisfen 2608 \subsection{The Ewald Sum}
114 chrisfen 2639 The complete accumulation electrostatic interactions in a system with
115     periodic boundary conditions (PBC) requires the consideration of the
116 gezelter 2643 effect of all charges within a (cubic) simulation box as well as those
117     in the periodic replicas,
118 chrisfen 2636 \begin{equation}
119     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],
120     \label{eq:PBCSum}
121     \end{equation}
122 chrisfen 2639 where the sum over $\mathbf{n}$ is a sum over all periodic box
123     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
124     prime indicates $i = j$ are neglected for $\mathbf{n} =
125     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
126     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
127     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
128 gezelter 2643 $j$, and $\phi$ is the solution to Poisson's equation
129     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
130     charge-charge interactions). In the case of monopole electrostatics,
131     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
132     non-neutral systems.
133 chrisfen 2604
134 gezelter 2643 The electrostatic summation problem was originally studied by Ewald
135 chrisfen 2636 for the case of an infinite crystal.\cite{Ewald21}. The approach he
136     took was to convert this conditionally convergent sum into two
137     absolutely convergent summations: a short-ranged real-space summation
138     and a long-ranged reciprocal-space summation,
139     \begin{equation}
140     \begin{split}
141 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,
142 chrisfen 2636 \end{split}
143     \label{eq:EwaldSum}
144     \end{equation}
145 chrisfen 2649 where $\alpha$ is the damping or convergence parameter with units of
146     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
147     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
148     constant of the surrounding medium. The final two terms of
149 chrisfen 2636 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
150     for interacting with a surrounding dielectric.\cite{Allen87} This
151     dipolar term was neglected in early applications in molecular
152     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
153     Leeuw {\it et al.} to address situations where the unit cell has a
154 gezelter 2643 dipole moment which is magnified through replication of the periodic
155     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
156     system is said to be using conducting (or ``tin-foil'') boundary
157 chrisfen 2637 conditions, $\epsilon_{\rm S} = \infty$. Figure
158 chrisfen 2636 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
159 gezelter 2653 time. Initially, due to the small system sizes that could be
160     simulated feasibly, the entire simulation box was replicated to
161     convergence. In more modern simulations, the systems have grown large
162     enough that a real-space cutoff could potentially give convergent
163     behavior. Indeed, it has been observed that with the choice of a
164     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
165     rapidly convergent and small relative to the real-space
166     portion.\cite{Karasawa89,Kolafa92}
167 gezelter 2643
168 chrisfen 2610 \begin{figure}
169     \centering
170 gezelter 2656 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
171 gezelter 2653 \caption{The change in the application of the Ewald sum with
172     increasing computational power. Initially, only small systems could
173     be studied, and the Ewald sum replicated the simulation box to
174     convergence. Now, much larger systems of charges are investigated
175     with fixed-distance cutoffs.}
176 chrisfen 2610 \label{fig:ewaldTime}
177     \end{figure}
178    
179 gezelter 2643 The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
180 chrisfen 2649 convergence parameter $(\alpha)$ plays an important role in balancing
181 gezelter 2643 the computational cost between the direct and reciprocal-space
182     portions of the summation. The choice of this value allows one to
183     select whether the real-space or reciprocal space portion of the
184     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
185     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
186     $\alpha$ and thoughtful algorithm development, this cost can be
187     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
188     taken to reduce the cost of the Ewald summation even further is to set
189     $\alpha$ such that the real-space interactions decay rapidly, allowing
190     for a short spherical cutoff. Then the reciprocal space summation is
191     optimized. These optimizations usually involve utilization of the
192     fast Fourier transform (FFT),\cite{Hockney81} leading to the
193 chrisfen 2637 particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
194     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
195     methods, the cost of the reciprocal-space portion of the Ewald
196 gezelter 2643 summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
197     \log N)$.
198 chrisfen 2636
199 gezelter 2643 These developments and optimizations have made the use of the Ewald
200     summation routine in simulations with periodic boundary
201     conditions. However, in certain systems, such as vapor-liquid
202     interfaces and membranes, the intrinsic three-dimensional periodicity
203     can prove problematic. The Ewald sum has been reformulated to handle
204     2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
205     new methods are computationally expensive.\cite{Spohr97,Yeh99}
206     Inclusion of a correction term in the Ewald summation is a possible
207     direction for handling 2D systems while still enabling the use of the
208     modern optimizations.\cite{Yeh99}
209 chrisfen 2637
210     Several studies have recognized that the inherent periodicity in the
211 gezelter 2643 Ewald sum can also have an effect on three-dimensional
212     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
213     Solvated proteins are essentially kept at high concentration due to
214     the periodicity of the electrostatic summation method. In these
215 chrisfen 2637 systems, the more compact folded states of a protein can be
216     artificially stabilized by the periodic replicas introduced by the
217 gezelter 2643 Ewald summation.\cite{Weber00} Thus, care must be taken when
218     considering the use of the Ewald summation where the assumed
219     periodicity would introduce spurious effects in the system dynamics.
220 chrisfen 2637
221 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
222 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
223 gezelter 2624 for the accurate accumulation of electrostatic interactions in an
224 gezelter 2643 efficient pairwise fashion. This procedure lacks the inherent
225     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
226     observed that the electrostatic interaction is effectively
227     short-ranged in condensed phase systems and that neutralization of the
228     charge contained within the cutoff radius is crucial for potential
229     stability. They devised a pairwise summation method that ensures
230     charge neutrality and gives results similar to those obtained with the
231     Ewald summation. The resulting shifted Coulomb potential
232     (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
233     placement on the cutoff sphere and a distance-dependent damping
234     function (identical to that seen in the real-space portion of the
235     Ewald sum) to aid convergence
236 chrisfen 2601 \begin{equation}
237 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\}.
238 chrisfen 2601 \label{eq:WolfPot}
239     \end{equation}
240 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
241     potential. However, neutralizing the charge contained within each
242     cutoff sphere requires the placement of a self-image charge on the
243     surface of the cutoff sphere. This additional self-term in the total
244 gezelter 2624 potential enabled Wolf {\it et al.} to obtain excellent estimates of
245 gezelter 2617 Madelung energies for many crystals.
246    
247     In order to use their charge-neutralized potential in molecular
248     dynamics simulations, Wolf \textit{et al.} suggested taking the
249     derivative of this potential prior to evaluation of the limit. This
250     procedure gives an expression for the forces,
251 chrisfen 2601 \begin{equation}
252 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\},
253 chrisfen 2601 \label{eq:WolfForces}
254     \end{equation}
255 gezelter 2617 that incorporates both image charges and damping of the electrostatic
256     interaction.
257    
258     More recently, Zahn \textit{et al.} investigated these potential and
259     force expressions for use in simulations involving water.\cite{Zahn02}
260 gezelter 2624 In their work, they pointed out that the forces and derivative of
261     the potential are not commensurate. Attempts to use both
262 gezelter 2643 eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
263 gezelter 2624 to poor energy conservation. They correctly observed that taking the
264     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
265     derivatives gives forces for a different potential energy function
266 gezelter 2643 than the one shown in eq. (\ref{eq:WolfPot}).
267 gezelter 2617
268 gezelter 2643 Zahn \textit{et al.} introduced a modified form of this summation
269     method as a way to use the technique in Molecular Dynamics
270     simulations. They proposed a new damped Coulomb potential,
271 chrisfen 2601 \begin{equation}
272 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\},
273 chrisfen 2601 \label{eq:ZahnPot}
274     \end{equation}
275 gezelter 2643 and showed that this potential does fairly well at capturing the
276 gezelter 2617 structural and dynamic properties of water compared the same
277     properties obtained using the Ewald sum.
278 chrisfen 2601
279 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
280    
281 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
282     al.} are constructed using two different (and separable) computational
283 gezelter 2624 tricks: \begin{enumerate}
284 gezelter 2617 \item shifting through the use of image charges, and
285     \item damping the electrostatic interaction.
286 gezelter 2624 \end{enumerate} Wolf \textit{et al.} treated the
287 gezelter 2617 development of their summation method as a progressive application of
288     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
289     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
290     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
291     both techniques. It is possible, however, to separate these
292     tricks and study their effects independently.
293    
294     Starting with the original observation that the effective range of the
295     electrostatic interaction in condensed phases is considerably less
296     than $r^{-1}$, either the cutoff sphere neutralization or the
297     distance-dependent damping technique could be used as a foundation for
298     a new pairwise summation method. Wolf \textit{et al.} made the
299     observation that charge neutralization within the cutoff sphere plays
300     a significant role in energy convergence; therefore we will begin our
301     analysis with the various shifted forms that maintain this charge
302     neutralization. We can evaluate the methods of Wolf
303     \textit{et al.} and Zahn \textit{et al.} by considering the standard
304     shifted potential,
305 chrisfen 2601 \begin{equation}
306 gezelter 2643 V_\textrm{SP}(r) = \begin{cases}
307 gezelter 2617 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
308     R_\textrm{c}
309     \end{cases},
310     \label{eq:shiftingPotForm}
311     \end{equation}
312     and shifted force,
313     \begin{equation}
314 gezelter 2643 V_\textrm{SF}(r) = \begin{cases}
315 gezelter 2624 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
316 gezelter 2617 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
317 chrisfen 2601 \end{cases},
318 chrisfen 2612 \label{eq:shiftingForm}
319 chrisfen 2601 \end{equation}
320 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
321     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
322     that both the potential and the forces goes to zero at the cutoff
323     radius, while the Shifted Potential ({\sc sp}) form only ensures the
324     potential is smooth at the cutoff radius
325     ($R_\textrm{c}$).\cite{Allen87}
326    
327 gezelter 2624 The forces associated with the shifted potential are simply the forces
328     of the unshifted potential itself (when inside the cutoff sphere),
329 chrisfen 2601 \begin{equation}
330 gezelter 2643 F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
331 chrisfen 2612 \end{equation}
332 gezelter 2624 and are zero outside. Inside the cutoff sphere, the forces associated
333     with the shifted force form can be written,
334 chrisfen 2612 \begin{equation}
335 gezelter 2643 F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
336 gezelter 2624 v(r)}{dr} \right)_{r=R_\textrm{c}}.
337     \end{equation}
338    
339 gezelter 2643 If the potential, $v(r)$, is taken to be the normal Coulomb potential,
340 gezelter 2624 \begin{equation}
341     v(r) = \frac{q_i q_j}{r},
342     \label{eq:Coulomb}
343     \end{equation}
344     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
345     al.}'s undamped prescription:
346     \begin{equation}
347 gezelter 2643 V_\textrm{SP}(r) =
348 gezelter 2624 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
349     r\leqslant R_\textrm{c},
350 chrisfen 2636 \label{eq:SPPot}
351 gezelter 2624 \end{equation}
352     with associated forces,
353     \begin{equation}
354 gezelter 2643 F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
355 chrisfen 2636 \label{eq:SPForces}
356 chrisfen 2612 \end{equation}
357 gezelter 2624 These forces are identical to the forces of the standard Coulomb
358     interaction, and cutting these off at $R_c$ was addressed by Wolf
359     \textit{et al.} as undesirable. They pointed out that the effect of
360     the image charges is neglected in the forces when this form is
361     used,\cite{Wolf99} thereby eliminating any benefit from the method in
362     molecular dynamics. Additionally, there is a discontinuity in the
363     forces at the cutoff radius which results in energy drift during MD
364     simulations.
365 chrisfen 2612
366 gezelter 2624 The shifted force ({\sc sf}) form using the normal Coulomb potential
367     will give,
368 chrisfen 2612 \begin{equation}
369 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}.
370 chrisfen 2612 \label{eq:SFPot}
371     \end{equation}
372 gezelter 2624 with associated forces,
373 chrisfen 2612 \begin{equation}
374 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}.
375 chrisfen 2612 \label{eq:SFForces}
376     \end{equation}
377 gezelter 2624 This formulation has the benefits that there are no discontinuities at
378 gezelter 2643 the cutoff radius, while the neutralizing image charges are present in
379     both the energy and force expressions. It would be simple to add the
380     self-neutralizing term back when computing the total energy of the
381 gezelter 2624 system, thereby maintaining the agreement with the Madelung energies.
382     A side effect of this treatment is the alteration in the shape of the
383     potential that comes from the derivative term. Thus, a degree of
384     clarity about agreement with the empirical potential is lost in order
385     to gain functionality in dynamics simulations.
386 chrisfen 2612
387 chrisfen 2620 Wolf \textit{et al.} originally discussed the energetics of the
388 gezelter 2643 shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
389     insufficient for accurate determination of the energy with reasonable
390     cutoff distances. The calculated Madelung energies fluctuated around
391     the expected value as the cutoff radius was increased, but the
392     oscillations converged toward the correct value.\cite{Wolf99} A
393 gezelter 2624 damping function was incorporated to accelerate the convergence; and
394 gezelter 2643 though alternative forms for the damping function could be
395 gezelter 2624 used,\cite{Jones56,Heyes81} the complimentary error function was
396     chosen to mirror the effective screening used in the Ewald summation.
397     Incorporating this error function damping into the simple Coulomb
398     potential,
399 chrisfen 2612 \begin{equation}
400 gezelter 2624 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
401 chrisfen 2601 \label{eq:dampCoulomb}
402     \end{equation}
403 gezelter 2643 the shifted potential (eq. (\ref{eq:SPPot})) becomes
404 chrisfen 2601 \begin{equation}
405 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},
406 chrisfen 2612 \label{eq:DSPPot}
407 chrisfen 2629 \end{equation}
408 gezelter 2624 with associated forces,
409 chrisfen 2612 \begin{equation}
410 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}.
411 chrisfen 2612 \label{eq:DSPForces}
412     \end{equation}
413 gezelter 2643 Again, this damped shifted potential suffers from a
414     force-discontinuity at the cutoff radius, and the image charges play
415     no role in the forces. To remedy these concerns, one may derive a
416     {\sc sf} variant by including the derivative term in
417     eq. (\ref{eq:shiftingForm}),
418 chrisfen 2612 \begin{equation}
419 chrisfen 2620 \begin{split}
420 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}.
421 chrisfen 2612 \label{eq:DSFPot}
422 chrisfen 2620 \end{split}
423 chrisfen 2612 \end{equation}
424 chrisfen 2636 The derivative of the above potential will lead to the following forces,
425 chrisfen 2612 \begin{equation}
426 chrisfen 2620 \begin{split}
427 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}.
428 chrisfen 2612 \label{eq:DSFForces}
429 chrisfen 2620 \end{split}
430 chrisfen 2612 \end{equation}
431 gezelter 2643 If the damping parameter $(\alpha)$ is set to zero, the undamped case,
432     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
433     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
434 chrisfen 2601
435 chrisfen 2636 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
436     derived by Zahn \textit{et al.}; however, there are two important
437     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
438     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
439     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
440     in the Zahn potential, resulting in a potential discontinuity as
441     particles cross $R_\textrm{c}$. Second, the sign of the derivative
442     portion is different. The missing $v_\textrm{c}$ term would not
443     affect molecular dynamics simulations (although the computed energy
444     would be expected to have sudden jumps as particle distances crossed
445 gezelter 2643 $R_c$). The sign problem is a potential source of errors, however.
446     In fact, it introduces a discontinuity in the forces at the cutoff,
447     because the force function is shifted in the wrong direction and
448     doesn't cross zero at $R_\textrm{c}$.
449 chrisfen 2602
450 gezelter 2624 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
451 gezelter 2643 electrostatic summation method in which the potential and forces are
452     continuous at the cutoff radius and which incorporates the damping
453     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
454     this paper, we will evaluate exactly how good these methods ({\sc sp},
455     {\sc sf}, damping) are at reproducing the correct electrostatic
456     summation performed by the Ewald sum.
457 gezelter 2624
458     \subsection{Other alternatives}
459 gezelter 2643 In addition to the methods described above, we considered some other
460     techniques that are commonly used in molecular simulations. The
461 chrisfen 2629 simplest of these is group-based cutoffs. Though of little use for
462 gezelter 2643 charged molecules, collecting atoms into neutral groups takes
463 chrisfen 2629 advantage of the observation that the electrostatic interactions decay
464     faster than those for monopolar pairs.\cite{Steinbach94} When
465 gezelter 2643 considering these molecules as neutral groups, the relative
466     orientations of the molecules control the strength of the interactions
467     at the cutoff radius. Consequently, as these molecular particles move
468     through $R_\textrm{c}$, the energy will drift upward due to the
469     anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
470     maintain good energy conservation, both the potential and derivative
471     need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
472     This is accomplished using a standard switching function. If a smooth
473     second derivative is desired, a fifth (or higher) order polynomial can
474     be used.\cite{Andrea83}
475 gezelter 2624
476 chrisfen 2629 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
477 gezelter 2643 and to incorporate the effects of the surroundings, a method like
478     Reaction Field ({\sc rf}) can be used. The original theory for {\sc
479     rf} was originally developed by Onsager,\cite{Onsager36} and it was
480     applied in simulations for the study of water by Barker and
481     Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
482     an extension of the group-based cutoff method where the net dipole
483     within the cutoff sphere polarizes an external dielectric, which
484     reacts back on the central dipole. The same switching function
485     considerations for group-based cutoffs need to made for {\sc rf}, with
486     the additional pre-specification of a dielectric constant.
487 gezelter 2624
488 chrisfen 2608 \section{Methods}
489    
490 chrisfen 2620 In classical molecular mechanics simulations, there are two primary
491     techniques utilized to obtain information about the system of
492     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
493     techniques utilize pairwise summations of interactions between
494     particle sites, but they use these summations in different ways.
495 chrisfen 2608
496 gezelter 2645 In MC, the potential energy difference between configurations dictates
497     the progression of MC sampling. Going back to the origins of this
498     method, the acceptance criterion for the canonical ensemble laid out
499     by Metropolis \textit{et al.} states that a subsequent configuration
500     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
501     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
502     Maintaining the correct $\Delta E$ when using an alternate method for
503     handling the long-range electrostatics will ensure proper sampling
504     from the ensemble.
505 chrisfen 2608
506 gezelter 2624 In MD, the derivative of the potential governs how the system will
507 chrisfen 2620 progress in time. Consequently, the force and torque vectors on each
508 gezelter 2624 body in the system dictate how the system evolves. If the magnitude
509     and direction of these vectors are similar when using alternate
510     electrostatic summation techniques, the dynamics in the short term
511     will be indistinguishable. Because error in MD calculations is
512     cumulative, one should expect greater deviation at longer times,
513     although methods which have large differences in the force and torque
514     vectors will diverge from each other more rapidly.
515 chrisfen 2608
516 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
517 gezelter 2645
518 gezelter 2624 The pairwise summation techniques (outlined in section
519     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
520     studying the energy differences between conformations. We took the
521 gezelter 2656 {\sc spme}-computed energy difference between two conformations to be the
522 gezelter 2624 correct behavior. An ideal performance by an alternative method would
523 gezelter 2645 reproduce these energy differences exactly (even if the absolute
524     energies calculated by the methods are different). Since none of the
525     methods provide exact energy differences, we used linear least squares
526     regressions of energy gap data to evaluate how closely the methods
527     mimicked the Ewald energy gaps. Unitary results for both the
528     correlation (slope) and correlation coefficient for these regressions
529 gezelter 2656 indicate perfect agreement between the alternative method and {\sc spme}.
530 gezelter 2645 Sample correlation plots for two alternate methods are shown in
531     Fig. \ref{fig:linearFit}.
532 chrisfen 2608
533 chrisfen 2609 \begin{figure}
534     \centering
535 chrisfen 2619 \includegraphics[width = \linewidth]{./dualLinear.pdf}
536 gezelter 2645 \caption{Example least squares regressions of the configuration energy
537     differences for SPC/E water systems. The upper plot shows a data set
538     with a poor correlation coefficient ($R^2$), while the lower plot
539     shows a data set with a good correlation coefficient.}
540     \label{fig:linearFit}
541 chrisfen 2609 \end{figure}
542    
543 gezelter 2624 Each system type (detailed in section \ref{sec:RepSims}) was
544     represented using 500 independent configurations. Additionally, we
545 gezelter 2645 used seven different system types, so each of the alternative
546 gezelter 2624 (non-Ewald) electrostatic summation methods was evaluated using
547     873,250 configurational energy differences.
548 chrisfen 2609
549 gezelter 2624 Results and discussion for the individual analysis of each of the
550     system types appear in the supporting information, while the
551     cumulative results over all the investigated systems appears below in
552     section \ref{sec:EnergyResults}.
553    
554 chrisfen 2609 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
555 gezelter 2624 We evaluated the pairwise methods (outlined in section
556     \ref{sec:ESMethods}) for use in MD simulations by
557     comparing the force and torque vectors with those obtained using the
558 gezelter 2656 reference Ewald summation ({\sc spme}). Both the magnitude and the
559 gezelter 2624 direction of these vectors on each of the bodies in the system were
560     analyzed. For the magnitude of these vectors, linear least squares
561     regression analyses were performed as described previously for
562     comparing $\Delta E$ values. Instead of a single energy difference
563     between two system configurations, we compared the magnitudes of the
564     forces (and torques) on each molecule in each configuration. For a
565     system of 1000 water molecules and 40 ions, there are 1040 force
566     vectors and 1000 torque vectors. With 500 configurations, this
567     results in 520,000 force and 500,000 torque vector comparisons.
568     Additionally, data from seven different system types was aggregated
569     before the comparison was made.
570 chrisfen 2609
571 gezelter 2624 The {\it directionality} of the force and torque vectors was
572     investigated through measurement of the angle ($\theta$) formed
573 gezelter 2656 between those computed from the particular method and those from {\sc spme},
574 chrisfen 2610 \begin{equation}
575 gezelter 2645 \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
576 chrisfen 2610 \end{equation}
577 gezelter 2656 where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
578 chrisfen 2651 vector computed using method M. Each of these $\theta$ values was
579     accumulated in a distribution function and weighted by the area on the
580 chrisfen 2652 unit sphere. Since this distribution is a measure of angular error
581     between two different electrostatic summation methods, there is no
582     {\it a priori} reason for the profile to adhere to any specific
583     shape. Thus, gaussian fits were used to measure the width of the
584     resulting distributions.
585 chrisfen 2651 %
586     %\begin{figure}
587     %\centering
588     %\includegraphics[width = \linewidth]{./gaussFit.pdf}
589     %\caption{Sample fit of the angular distribution of the force vectors
590     %accumulated using all of the studied systems. Gaussian fits were used
591     %to obtain values for the variance in force and torque vectors.}
592     %\label{fig:gaussian}
593     %\end{figure}
594     %
595     %Figure \ref{fig:gaussian} shows an example distribution with applied
596     %non-linear fits. The solid line is a Gaussian profile, while the
597     %dotted line is a Voigt profile, a convolution of a Gaussian and a
598     %Lorentzian.
599     %Since this distribution is a measure of angular error between two
600     %different electrostatic summation methods, there is no {\it a priori}
601     %reason for the profile to adhere to any specific shape.
602     %Gaussian fits was used to compare all the tested methods.
603     The variance ($\sigma^2$) was extracted from each of these fits and
604     was used to compare distribution widths. Values of $\sigma^2$ near
605     zero indicate vector directions indistinguishable from those
606 gezelter 2656 calculated when using the reference method ({\sc spme}).
607 gezelter 2624
608     \subsection{Short-time Dynamics}
609 gezelter 2645
610     The effects of the alternative electrostatic summation methods on the
611     short-time dynamics of charged systems were evaluated by considering a
612     NaCl crystal at a temperature of 1000 K. A subset of the best
613     performing pairwise methods was used in this comparison. The NaCl
614     crystal was chosen to avoid possible complications from the treatment
615     of orientational motion in molecular systems. All systems were
616     started with the same initial positions and velocities. Simulations
617     were performed under the microcanonical ensemble, and velocity
618 chrisfen 2638 autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
619     of the trajectories,
620 chrisfen 2609 \begin{equation}
621 gezelter 2656 C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
622 chrisfen 2609 \label{eq:vCorr}
623     \end{equation}
624 chrisfen 2638 Velocity autocorrelation functions require detailed short time data,
625     thus velocity information was saved every 2 fs over 10 ps
626     trajectories. Because the NaCl crystal is composed of two different
627     atom types, the average of the two resulting velocity autocorrelation
628     functions was used for comparisons.
629    
630     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
631 gezelter 2645
632     The effects of the same subset of alternative electrostatic methods on
633     the {\it long-time} dynamics of charged systems were evaluated using
634     the same model system (NaCl crystals at 1000K). The power spectrum
635     ($I(\omega)$) was obtained via Fourier transform of the velocity
636     autocorrelation function, \begin{equation} I(\omega) =
637     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
638 chrisfen 2609 \label{eq:powerSpec}
639     \end{equation}
640 chrisfen 2638 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
641     NaCl crystal is composed of two different atom types, the average of
642 gezelter 2645 the two resulting power spectra was used for comparisons. Simulations
643     were performed under the microcanonical ensemble, and velocity
644     information was saved every 5 fs over 100 ps trajectories.
645 chrisfen 2609
646     \subsection{Representative Simulations}\label{sec:RepSims}
647 gezelter 2645 A variety of representative simulations were analyzed to determine the
648     relative effectiveness of the pairwise summation techniques in
649 gezelter 2656 reproducing the energetics and dynamics exhibited by {\sc spme}. We wanted
650 gezelter 2645 to span the space of modern simulations (i.e. from liquids of neutral
651     molecules to ionic crystals), so the systems studied were:
652 chrisfen 2599 \begin{enumerate}
653 gezelter 2645 \item liquid water (SPC/E),\cite{Berendsen87}
654     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
655     \item NaCl crystals,
656     \item NaCl melts,
657     \item a low ionic strength solution of NaCl in water (0.11 M),
658     \item a high ionic strength solution of NaCl in water (1.1 M), and
659     \item a 6 \AA\ radius sphere of Argon in water.
660 chrisfen 2599 \end{enumerate}
661 chrisfen 2620 By utilizing the pairwise techniques (outlined in section
662     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
663 gezelter 2645 charged particles, and mixtures of the two, we hope to discern under
664     which conditions it will be possible to use one of the alternative
665     summation methodologies instead of the Ewald sum.
666 chrisfen 2586
667 gezelter 2645 For the solid and liquid water configurations, configurations were
668     taken at regular intervals from high temperature trajectories of 1000
669     SPC/E water molecules. Each configuration was equilibrated
670     independently at a lower temperature (300~K for the liquid, 200~K for
671     the crystal). The solid and liquid NaCl systems consisted of 500
672     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
673     these systems were selected and equilibrated in the same manner as the
674     water systems. The equilibrated temperatures were 1000~K for the NaCl
675     crystal and 7000~K for the liquid. The ionic solutions were made by
676     solvating 4 (or 40) ions in a periodic box containing 1000 SPC/E water
677     molecules. Ion and water positions were then randomly swapped, and
678     the resulting configurations were again equilibrated individually.
679     Finally, for the Argon / Water ``charge void'' systems, the identities
680     of all the SPC/E waters within 6 \AA\ of the center of the
681 chrisfen 2651 equilibrated water configurations were converted to argon.
682     %(Fig. \ref{fig:argonSlice}).
683 chrisfen 2586
684 gezelter 2645 These procedures guaranteed us a set of representative configurations
685 gezelter 2653 from chemically-relevant systems sampled from appropriate
686     ensembles. Force field parameters for the ions and Argon were taken
687 gezelter 2645 from the force field utilized by {\sc oopse}.\cite{Meineke05}
688    
689 chrisfen 2651 %\begin{figure}
690     %\centering
691     %\includegraphics[width = \linewidth]{./slice.pdf}
692     %\caption{A slice from the center of a water box used in a charge void
693     %simulation. The darkened region represents the boundary sphere within
694     %which the water molecules were converted to argon atoms.}
695     %\label{fig:argonSlice}
696     %\end{figure}
697 chrisfen 2586
698 gezelter 2645 \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
699     We compared the following alternative summation methods with results
700 gezelter 2656 from the reference method ({\sc spme}):
701 gezelter 2645 \begin{itemize}
702     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
703     and 0.3 \AA$^{-1}$,
704     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
705     and 0.3 \AA$^{-1}$,
706     \item reaction field with an infinite dielectric constant, and
707     \item an unmodified cutoff.
708     \end{itemize}
709     Group-based cutoffs with a fifth-order polynomial switching function
710     were utilized for the reaction field simulations. Additionally, we
711 gezelter 2656 investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
712     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
713     implementation of {\sc spme},\cite{Ponder87} while all other calculations
714 gezelter 2653 were performed using the {\sc oopse} molecular mechanics
715 gezelter 2645 package.\cite{Meineke05} All other portions of the energy calculation
716     (i.e. Lennard-Jones interactions) were handled in exactly the same
717     manner across all systems and configurations.
718 chrisfen 2586
719 gezelter 2645 The althernative methods were also evaluated with three different
720 chrisfen 2649 cutoff radii (9, 12, and 15 \AA). As noted previously, the
721     convergence parameter ($\alpha$) plays a role in the balance of the
722     real-space and reciprocal-space portions of the Ewald calculation.
723     Typical molecular mechanics packages set this to a value dependent on
724     the cutoff radius and a tolerance (typically less than $1 \times
725     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
726 gezelter 2653 increasing accuracy at the expense of computational time spent on the
727     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
728 gezelter 2656 The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
729     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
730 gezelter 2653 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
731     respectively.
732 chrisfen 2609
733 chrisfen 2575 \section{Results and Discussion}
734    
735 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
736 chrisfen 2620 In order to evaluate the performance of the pairwise electrostatic
737     summation methods for Monte Carlo simulations, the energy differences
738     between configurations were compared to the values obtained when using
739 gezelter 2656 {\sc spme}. The results for the subsequent regression analysis are shown in
740 chrisfen 2620 figure \ref{fig:delE}.
741 chrisfen 2590
742     \begin{figure}
743     \centering
744 gezelter 2617 \includegraphics[width=5.5in]{./delEplot.pdf}
745 gezelter 2645 \caption{Statistical analysis of the quality of configurational energy
746     differences for a given electrostatic method compared with the
747     reference Ewald sum. Results with a value equal to 1 (dashed line)
748     indicate $\Delta E$ values indistinguishable from those obtained using
749 gezelter 2656 {\sc spme}. Different values of the cutoff radius are indicated with
750 gezelter 2645 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
751     inverted triangles).}
752 chrisfen 2601 \label{fig:delE}
753 chrisfen 2594 \end{figure}
754    
755 gezelter 2645 The most striking feature of this plot is how well the Shifted Force
756     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
757     differences. For the undamped {\sc sf} method, and the
758     moderately-damped {\sc sp} methods, the results are nearly
759     indistinguishable from the Ewald results. The other common methods do
760     significantly less well.
761 chrisfen 2594
762 gezelter 2645 The unmodified cutoff method is essentially unusable. This is not
763     surprising since hard cutoffs give large energy fluctuations as atoms
764     or molecules move in and out of the cutoff
765     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
766     some degree by using group based cutoffs with a switching
767     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
768     significant improvement using the group-switched cutoff because the
769     salt and salt solution systems contain non-neutral groups. Interested
770     readers can consult the accompanying supporting information for a
771     comparison where all groups are neutral.
772    
773 gezelter 2653 For the {\sc sp} method, inclusion of electrostatic damping improves
774     the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
775 gezelter 2656 shows an excellent correlation and quality of fit with the {\sc spme}
776 gezelter 2653 results, particularly with a cutoff radius greater than 12
777 gezelter 2645 \AA . Use of a larger damping parameter is more helpful for the
778     shortest cutoff shown, but it has a detrimental effect on simulations
779     with larger cutoffs.
780 chrisfen 2609
781 gezelter 2653 In the {\sc sf} sets, increasing damping results in progressively {\it
782     worse} correlation with Ewald. Overall, the undamped case is the best
783 gezelter 2645 performing set, as the correlation and quality of fits are
784     consistently superior regardless of the cutoff distance. The undamped
785     case is also less computationally demanding (because no evaluation of
786     the complementary error function is required).
787    
788     The reaction field results illustrates some of that method's
789     limitations, primarily that it was developed for use in homogenous
790     systems; although it does provide results that are an improvement over
791     those from an unmodified cutoff.
792    
793 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
794 chrisfen 2599
795 chrisfen 2620 Evaluation of pairwise methods for use in Molecular Dynamics
796     simulations requires consideration of effects on the forces and
797 gezelter 2653 torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
798     regression results for the force and torque vector magnitudes,
799     respectively. The data in these figures was generated from an
800     accumulation of the statistics from all of the system types.
801 chrisfen 2594
802     \begin{figure}
803     \centering
804 gezelter 2617 \includegraphics[width=5.5in]{./frcMagplot.pdf}
805 chrisfen 2651 \caption{Statistical analysis of the quality of the force vector
806     magnitudes for a given electrostatic method compared with the
807     reference Ewald sum. Results with a value equal to 1 (dashed line)
808     indicate force magnitude values indistinguishable from those obtained
809 gezelter 2656 using {\sc spme}. Different values of the cutoff radius are indicated with
810 chrisfen 2651 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
811     inverted triangles).}
812 chrisfen 2601 \label{fig:frcMag}
813 chrisfen 2594 \end{figure}
814    
815 gezelter 2653 Again, it is striking how well the Shifted Potential and Shifted Force
816 gezelter 2656 methods are doing at reproducing the {\sc spme} forces. The undamped and
817 gezelter 2653 weakly-damped {\sc sf} method gives the best agreement with Ewald.
818     This is perhaps expected because this method explicitly incorporates a
819     smooth transition in the forces at the cutoff radius as well as the
820     neutralizing image charges.
821    
822 chrisfen 2620 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
823     in the previous $\Delta E$ section. The unmodified cutoff results are
824     poor, but using group based cutoffs and a switching function provides
825 gezelter 2653 an improvement much more significant than what was seen with $\Delta
826     E$.
827    
828     With moderate damping and a large enough cutoff radius, the {\sc sp}
829     method is generating usable forces. Further increases in damping,
830 chrisfen 2620 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
831 gezelter 2653 detrimental to simulations with larger cutoff radii.
832    
833     The reaction field results are surprisingly good, considering the poor
834 chrisfen 2620 quality of the fits for the $\Delta E$ results. There is still a
835 gezelter 2653 considerable degree of scatter in the data, but the forces correlate
836     well with the Ewald forces in general. We note that the reaction
837     field calculations do not include the pure NaCl systems, so these
838 chrisfen 2620 results are partly biased towards conditions in which the method
839     performs more favorably.
840 chrisfen 2594
841     \begin{figure}
842     \centering
843 gezelter 2617 \includegraphics[width=5.5in]{./trqMagplot.pdf}
844 chrisfen 2651 \caption{Statistical analysis of the quality of the torque vector
845     magnitudes for a given electrostatic method compared with the
846     reference Ewald sum. Results with a value equal to 1 (dashed line)
847     indicate torque magnitude values indistinguishable from those obtained
848 gezelter 2656 using {\sc spme}. Different values of the cutoff radius are indicated with
849 chrisfen 2651 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
850     inverted triangles).}
851 chrisfen 2601 \label{fig:trqMag}
852 chrisfen 2594 \end{figure}
853    
854 gezelter 2653 Molecular torques were only available from the systems which contained
855     rigid molecules (i.e. the systems containing water). The data in
856     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
857 chrisfen 2594
858 gezelter 2653 Torques appear to be much more sensitive to charges at a longer
859     distance. The striking feature in comparing the new electrostatic
860 gezelter 2656 methods with {\sc spme} is how much the agreement improves with increasing
861 gezelter 2653 cutoff radius. Again, the weakly damped and undamped {\sc sf} method
862 gezelter 2656 appears to be reproducing the {\sc spme} torques most accurately.
863 gezelter 2653
864     Water molecules are dipolar, and the reaction field method reproduces
865     the effect of the surrounding polarized medium on each of the
866     molecular bodies. Therefore it is not surprising that reaction field
867     performs best of all of the methods on molecular torques.
868    
869 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
870 chrisfen 2599
871 gezelter 2653 It is clearly important that a new electrostatic method can reproduce
872     the magnitudes of the force and torque vectors obtained via the Ewald
873     sum. However, the {\it directionality} of these vectors will also be
874     vital in calculating dynamical quantities accurately. Force and
875     torque directionalities were investigated by measuring the angles
876     formed between these vectors and the same vectors calculated using
877 gezelter 2656 {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
878 gezelter 2653 variance ($\sigma^2$) of the Gaussian fits of the angle error
879     distributions of the combined set over all system types.
880 chrisfen 2594
881     \begin{figure}
882     \centering
883 gezelter 2617 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
884 gezelter 2653 \caption{Statistical analysis of the width of the angular distribution
885     that the force and torque vectors from a given electrostatic method
886     make with their counterparts obtained using the reference Ewald sum.
887     Results with a variance ($\sigma^2$) equal to zero (dashed line)
888     indicate force and torque directions indistinguishable from those
889 gezelter 2656 obtained using {\sc spme}. Different values of the cutoff radius are
890 gezelter 2653 indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
891     and 15\AA\ = inverted triangles).}
892 chrisfen 2601 \label{fig:frcTrqAng}
893 chrisfen 2594 \end{figure}
894    
895 chrisfen 2620 Both the force and torque $\sigma^2$ results from the analysis of the
896     total accumulated system data are tabulated in figure
897 gezelter 2653 \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
898 gezelter 2656 sp}) method would be essentially unusable for molecular dynamics
899     unless the damping function is added. The Shifted Force ({\sc sf})
900     method, however, is generating force and torque vectors which are
901     within a few degrees of the Ewald results even with weak (or no)
902     damping.
903 chrisfen 2594
904 gezelter 2653 All of the sets (aside from the over-damped case) show the improvement
905     afforded by choosing a larger cutoff radius. Increasing the cutoff
906     from 9 to 12 \AA\ typically results in a halving of the width of the
907 gezelter 2656 distribution, with a similar improvement when going from 12 to 15
908 gezelter 2653 \AA .
909    
910     The undamped {\sc sf}, group-based cutoff, and reaction field methods
911     all do equivalently well at capturing the direction of both the force
912 gezelter 2656 and torque vectors. Using the electrostatic damping improves the
913     angular behavior significantly for the {\sc sp} and moderately for the
914     {\sc sf} methods. Overdamping is detrimental to both methods. Again
915     it is important to recognize that the force vectors cover all
916     particles in all seven systems, while torque vectors are only
917     available for neutral molecular groups. Damping is more beneficial to
918 gezelter 2653 charged bodies, and this observation is investigated further in the
919     accompanying supporting information.
920    
921     Although not discussed previously, group based cutoffs can be applied
922 gezelter 2656 to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
923     will reintroduce small discontinuities at the cutoff radius, but the
924     effects of these can be minimized by utilizing a switching function.
925     Though there are no significant benefits or drawbacks observed in
926     $\Delta E$ and the force and torque magnitudes when doing this, there
927     is a measurable improvement in the directionality of the forces and
928     torques. Table \ref{tab:groupAngle} shows the angular variances
929     obtained using group based cutoffs along with the results seen in
930     figure \ref{fig:frcTrqAng}. The {\sc sp} (with an $\alpha$ of 0.2
931     \AA$^{-1}$ or smaller) shows much narrower angular distributions when
932     using group-based cutoffs. The {\sc sf} method likewise shows
933     improvement in the undamped and lightly damped cases.
934 gezelter 2653
935 chrisfen 2595 \begin{table}[htbp]
936 gezelter 2656 \centering
937     \caption{Statistical analysis of the angular
938     distributions that the force (upper) and torque (lower) vectors
939     from a given electrostatic method make with their counterparts
940     obtained using the reference Ewald sum. Calculations were
941     performed both with (Y) and without (N) group based cutoffs and a
942     switching function. The $\alpha$ values have units of \AA$^{-1}$
943     and the variance values have units of degrees$^2$.}
944    
945 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
946 chrisfen 2595 \\
947     \toprule
948     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
949     \cmidrule(lr){3-6}
950     \cmidrule(l){7-10}
951 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
952 chrisfen 2595 \midrule
953 chrisfen 2599
954     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
955     & \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} \\
956     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
957     & \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} \\
958     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
959     & \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} \\
960 chrisfen 2594
961 chrisfen 2595 \midrule
962    
963 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
964     & \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} \\
965     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
966     & \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} \\
967     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
968     & \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} \\
969 chrisfen 2595
970     \bottomrule
971     \end{tabular}
972 chrisfen 2601 \label{tab:groupAngle}
973 chrisfen 2595 \end{table}
974    
975 gezelter 2656 One additional trend in table \ref{tab:groupAngle} is that the
976     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
977     increases, something that is more obvious with group-based cutoffs.
978     The complimentary error function inserted into the potential weakens
979     the electrostatic interaction as the value of $\alpha$ is increased.
980     However, at larger values of $\alpha$, it is possible to overdamp the
981     electrostatic interaction and to remove it completely. Kast
982 gezelter 2653 \textit{et al.} developed a method for choosing appropriate $\alpha$
983     values for these types of electrostatic summation methods by fitting
984     to $g(r)$ data, and their methods indicate optimal values of 0.34,
985     0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
986     respectively.\cite{Kast03} These appear to be reasonable choices to
987     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
988     these findings, choices this high would introduce error in the
989 gezelter 2656 molecular torques, particularly for the shorter cutoffs. Based on our
990     observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
991     but damping may be unnecessary when using the {\sc sf} method.
992 chrisfen 2595
993 chrisfen 2638 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
994 chrisfen 2601
995 gezelter 2653 Zahn {\it et al.} investigated the structure and dynamics of water
996     using eqs. (\ref{eq:ZahnPot}) and
997     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
998     that a method similar (but not identical with) the damped {\sc sf}
999     method resulted in properties very similar to those obtained when
1000     using the Ewald summation. The properties they studied (pair
1001     distribution functions, diffusion constants, and velocity and
1002     orientational correlation functions) may not be particularly sensitive
1003     to the long-range and collective behavior that governs the
1004 gezelter 2656 low-frequency behavior in crystalline systems. Additionally, the
1005     ionic crystals are the worst case scenario for the pairwise methods
1006     because they lack the reciprocal space contribution contained in the
1007     Ewald summation.
1008 chrisfen 2601
1009 gezelter 2653 We are using two separate measures to probe the effects of these
1010     alternative electrostatic methods on the dynamics in crystalline
1011     materials. For short- and intermediate-time dynamics, we are
1012     computing the velocity autocorrelation function, and for long-time
1013     and large length-scale collective motions, we are looking at the
1014     low-frequency portion of the power spectrum.
1015    
1016 chrisfen 2601 \begin{figure}
1017     \centering
1018 chrisfen 2638 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
1019 gezelter 2656 \caption{Velocity autocorrelation functions of NaCl crystals at
1020     1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1021 gezelter 2653 sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
1022     the first minimum. The times to first collision are nearly identical,
1023     but differences can be seen in the peaks and troughs, where the
1024     undamped and weakly damped methods are stiffer than the moderately
1025 gezelter 2656 damped and {\sc spme} methods.}
1026 chrisfen 2638 \label{fig:vCorrPlot}
1027     \end{figure}
1028    
1029 gezelter 2656 The short-time decay of the velocity autocorrelation function through
1030     the first collision are nearly identical in figure
1031     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1032     how the methods differ. The undamped {\sc sf} method has deeper
1033     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1034     any of the other methods. As the damping parameter ($\alpha$) is
1035     increased, these peaks are smoothed out, and the {\sc sf} method
1036     approaches the {\sc spme} results. With $\alpha$ values of 0.2 \AA$^{-1}$,
1037     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1038     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1039     and {\sc sp} potentials become nearly identical with increased
1040     damping. However, this appears to indicate that once damping is
1041     utilized, the details of the form of the potential (and forces)
1042     constructed out of the damped electrostatic interaction are less
1043     important.
1044 chrisfen 2638
1045     \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1046    
1047 gezelter 2656 To evaluate how the differences between the methods affect the
1048     collective long-time motion, we computed power spectra from long-time
1049     traces of the velocity autocorrelation function. The power spectra for
1050     the best-performing alternative methods are shown in
1051     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1052     a cubic switching function between 40 and 50 ps was used to reduce the
1053     ringing resulting from data truncation. This procedure had no
1054     noticeable effect on peak location or magnitude.
1055 chrisfen 2638
1056     \begin{figure}
1057     \centering
1058 gezelter 2617 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1059 chrisfen 2651 \caption{Power spectra obtained from the velocity auto-correlation
1060 gezelter 2656 functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf}
1061     ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset
1062     shows the frequency region below 100 cm$^{-1}$ to highlight where the
1063     spectra differ.}
1064 chrisfen 2610 \label{fig:methodPS}
1065 chrisfen 2601 \end{figure}
1066    
1067 gezelter 2656 While the high frequency regions of the power spectra for the
1068     alternative methods are quantitatively identical with Ewald spectrum,
1069     the low frequency region shows how the summation methods differ.
1070     Considering the low-frequency inset (expanded in the upper frame of
1071     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1072     correlated motions are blue-shifted when using undamped or weakly
1073     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1074     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1075     correlated motion to the Ewald method (which has a convergence
1076     parameter of 0.3119 \AA$^{-1}$). This weakening of the electrostatic
1077     interaction with increased damping explains why the long-ranged
1078     correlated motions are at lower frequencies for the moderately damped
1079     methods than for undamped or weakly damped methods.
1080    
1081     To isolate the role of the damping constant, we have computed the
1082     spectra for a single method ({\sc sf}) with a range of damping
1083     constants and compared this with the {\sc spme} spectrum.
1084     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1085     electrostatic damping red-shifts the lowest frequency phonon modes.
1086     However, even without any electrostatic damping, the {\sc sf} method
1087     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1088     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1089     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1090     Most} of the collective behavior in the crystal is accurately captured
1091     using the {\sc sf} method. Quantitative agreement with Ewald can be
1092     obtained using moderate damping in addition to the shifting at the
1093     cutoff distance.
1094    
1095 chrisfen 2601 \begin{figure}
1096     \centering
1097 chrisfen 2659 \includegraphics[width = \linewidth]{./increasedDamping.pdf}
1098 gezelter 2656 \caption{Effect of damping on the two lowest-frequency phonon modes in
1099     the NaCl crystal at 1000K. The undamped shifted force ({\sc sf})
1100     method is off by less than 10 cm$^{-1}$, and increasing the
1101     electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1102     with the power spectrum obtained using the Ewald sum. Overdamping can
1103     result in underestimates of frequencies of the long-wavelength
1104     motions.}
1105 chrisfen 2601 \label{fig:dampInc}
1106     \end{figure}
1107    
1108 chrisfen 2575 \section{Conclusions}
1109    
1110 chrisfen 2620 This investigation of pairwise electrostatic summation techniques
1111 gezelter 2656 shows that there are viable and computationally efficient alternatives
1112     to the Ewald summation. These methods are derived from the damped and
1113     cutoff-neutralized Coulombic sum originally proposed by Wolf
1114     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1115     method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1116     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1117     energetic and dynamic characteristics exhibited by simulations
1118     employing lattice summation techniques. The cumulative energy
1119     difference results showed the undamped {\sc sf} and moderately damped
1120     {\sc sp} methods produced results nearly identical to {\sc spme}. Similarly
1121     for the dynamic features, the undamped or moderately damped {\sc sf}
1122     and moderately damped {\sc sp} methods produce force and torque vector
1123     magnitude and directions very similar to the expected values. These
1124     results translate into long-time dynamic behavior equivalent to that
1125     produced in simulations using {\sc spme}.
1126 chrisfen 2604
1127 gezelter 2656 As in all purely-pairwise cutoff methods, these methods are expected
1128     to scale approximately {\it linearly} with system size, and they are
1129     easily parallelizable. This should result in substantial reductions
1130     in the computational cost of performing large simulations.
1131    
1132 chrisfen 2620 Aside from the computational cost benefit, these techniques have
1133     applicability in situations where the use of the Ewald sum can prove
1134 gezelter 2656 problematic. Of greatest interest is their potential use in
1135     interfacial systems, where the unmodified lattice sum techniques
1136     artificially accentuate the periodicity of the system in an
1137     undesirable manner. There have been alterations to the standard Ewald
1138     techniques, via corrections and reformulations, to compensate for
1139     these systems; but the pairwise techniques discussed here require no
1140     modifications, making them natural tools to tackle these problems.
1141     Additionally, this transferability gives them benefits over other
1142     pairwise methods, like reaction field, because estimations of physical
1143     properties (e.g. the dielectric constant) are unnecessary.
1144 chrisfen 2605
1145 gezelter 2656 If a researcher is using Monte Carlo simulations of large chemical
1146     systems containing point charges, most structural features will be
1147     accurately captured using the undamped {\sc sf} method or the {\sc sp}
1148     method with an electrostatic damping of 0.2 \AA$^{-1}$. These methods
1149     would also be appropriate for molecular dynamics simulations where the
1150     data of interest is either structural or short-time dynamical
1151     quantities. For long-time dynamics and collective motions, the safest
1152     pairwise method we have evaluated is the {\sc sf} method with an
1153     electrostatic damping between 0.2 and 0.25
1154     \AA$^{-1}$.
1155 chrisfen 2605
1156 gezelter 2656 We are not suggesting that there is any flaw with the Ewald sum; in
1157     fact, it is the standard by which these simple pairwise sums have been
1158     judged. However, these results do suggest that in the typical
1159     simulations performed today, the Ewald summation may no longer be
1160     required to obtain the level of accuracy most researchers have come to
1161     expect.
1162    
1163 chrisfen 2575 \section{Acknowledgments}
1164 gezelter 2656 Support for this project was provided by the National Science
1165     Foundation under grant CHE-0134881. The authors would like to thank
1166     Steve Corcelli and Ed Maginn for helpful discussions and comments.
1167    
1168 chrisfen 2594 \newpage
1169    
1170 gezelter 2617 \bibliographystyle{jcp2}
1171 chrisfen 2575 \bibliography{electrostaticMethods}
1172    
1173    
1174     \end{document}