ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2624
Committed: Wed Mar 15 17:09:09 2006 UTC (18 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 47755 byte(s)
Log Message:
paper

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 %\documentclass[aps,prb,preprint]{revtex4}
3 \documentclass[11pt]{article}
4 \usepackage{endfloat}
5 \usepackage{amsmath}
6 \usepackage{amssymb}
7 \usepackage{epsf}
8 \usepackage{times}
9 \usepackage{mathptmx}
10 \usepackage{setspace}
11 \usepackage{tabularx}
12 \usepackage{graphicx}
13 \usepackage{booktabs}
14 \usepackage{bibentry}
15 \usepackage{mathrsfs}
16 \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 \title{Is the Ewald Summation necessary? Pairwise alternatives to the accepted standard for long-range electrostatics}
29
30 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
31 gezelter@nd.edu} \\
32 Department of Chemistry and Biochemistry\\
33 University of Notre Dame\\
34 Notre Dame, Indiana 46556}
35
36 \date{\today}
37
38 \maketitle
39 \doublespacing
40
41 \nobibliography{}
42 \begin{abstract}
43 A new method for accumulating electrostatic interactions was derived
44 from the previous efforts described in \bibentry{Wolf99} and
45 \bibentry{Zahn02} as a possible replacement for lattice sum methods in
46 molecular simulations. Comparisons were performed with this and other
47 pairwise electrostatic summation techniques against the smooth
48 particle mesh Ewald (SPME) summation to see how well they reproduce
49 the energetics and dynamics of a variety of simulation types. The
50 newly derived Shifted-Force technique shows a remarkable ability to
51 reproduce the behavior exhibited in simulations using SPME with an
52 $\mathscr{O}(N)$ computational cost, equivalent to merely the
53 real-space portion of the lattice summation.
54
55 \end{abstract}
56
57 \newpage
58
59 %\narrowtext
60
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 % BODY OF TEXT
63 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64
65 \section{Introduction}
66
67 In molecular simulations, proper accumulation of the electrostatic
68 interactions is considered one of the most essential and
69 computationally demanding tasks. The common molecular mechanics force
70 fields are founded on representation of the atomic sites centered on
71 full or partial charges shielded by Lennard-Jones type interactions.
72 This means that nearly every pair interaction involves an
73 charge-charge calculation. Coupled with $r^{-1}$ decay, the monopole
74 interactions quickly become a burden for molecular systems of all
75 sizes. For example, in small systems, the electrostatic pair
76 interaction may not have decayed appreciably within the box length
77 leading to an effect excluded from the pair interactions within a unit
78 box. In large systems, excessively large cutoffs need to be used to
79 accurately incorporate their effect, and since the computational cost
80 increases proportionally with the cutoff sphere, it quickly becomes an
81 impractical task to perform these calculations.
82
83 \subsection{The Ewald Sum}
84 blah blah blah Ewald Sum Important blah blah blah
85
86 \begin{figure}
87 \centering
88 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
89 \caption{How the application of the Ewald summation has changed with
90 the increase in computer power. Initially, only small numbers of
91 particles could be studied, and the Ewald sum acted to replicate the
92 unit cell charge distribution out to convergence. Now, much larger
93 systems of charges are investigated with fixed distance cutoffs. The
94 calculated structure factor is used to sum out to great distance, and
95 a surrounding dielectric term is included.}
96 \label{fig:ewaldTime}
97 \end{figure}
98
99 \subsection{The Wolf and Zahn Methods}
100 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
101 for the accurate accumulation of electrostatic interactions in an
102 efficient pairwise fashion.\cite{Wolf99} Wolf \textit{et al.} observed
103 that the electrostatic interaction is effectively short-ranged in
104 condensed phase systems and that neutralization of the charge
105 contained within the cutoff radius is crucial for potential
106 stability. They devised a pairwise summation method that ensures
107 charge neutrality and gives results similar to those obtained with
108 the Ewald summation. The resulting shifted Coulomb potential
109 (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
110 placement on the cutoff sphere and a distance-dependent damping
111 function (identical to that seen in the real-space portion of the
112 Ewald sum) to aid convergence
113 \begin{equation}
114 V_{\textrm{Wolf}}(r_{ij})= \frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
115 \label{eq:WolfPot}
116 \end{equation}
117 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
118 potential. However, neutralizing the charge contained within each
119 cutoff sphere requires the placement of a self-image charge on the
120 surface of the cutoff sphere. This additional self-term in the total
121 potential enabled Wolf {\it et al.} to obtain excellent estimates of
122 Madelung energies for many crystals.
123
124 In order to use their charge-neutralized potential in molecular
125 dynamics simulations, Wolf \textit{et al.} suggested taking the
126 derivative of this potential prior to evaluation of the limit. This
127 procedure gives an expression for the forces,
128 \begin{equation}
129 F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{-\left[\frac{\textrm{erfc}(\alpha r_{ij})}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r_{ij}^2)}}{r_{ij}}\right]+\left[\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right]\right\},
130 \label{eq:WolfForces}
131 \end{equation}
132 that incorporates both image charges and damping of the electrostatic
133 interaction.
134
135 More recently, Zahn \textit{et al.} investigated these potential and
136 force expressions for use in simulations involving water.\cite{Zahn02}
137 In their work, they pointed out that the forces and derivative of
138 the potential are not commensurate. Attempts to use both
139 Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
140 to poor energy conservation. They correctly observed that taking the
141 limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
142 derivatives gives forces for a different potential energy function
143 than the one shown in Eq. (\ref{eq:WolfPot}).
144
145 Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
146 method'' as a way to use this technique in Molecular Dynamics
147 simulations. Taking the integral of the forces shown in equation
148 \ref{eq:WolfForces}, they proposed a new damped Coulomb
149 potential,
150 \begin{equation}
151 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\}.
152 \label{eq:ZahnPot}
153 \end{equation}
154 They showed that this potential does fairly well at capturing the
155 structural and dynamic properties of water compared the same
156 properties obtained using the Ewald sum.
157
158 \subsection{Simple Forms for Pairwise Electrostatics}
159
160 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
161 al.} are constructed using two different (and separable) computational
162 tricks: \begin{enumerate}
163 \item shifting through the use of image charges, and
164 \item damping the electrostatic interaction.
165 \end{enumerate} Wolf \textit{et al.} treated the
166 development of their summation method as a progressive application of
167 these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
168 their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
169 post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
170 both techniques. It is possible, however, to separate these
171 tricks and study their effects independently.
172
173 Starting with the original observation that the effective range of the
174 electrostatic interaction in condensed phases is considerably less
175 than $r^{-1}$, either the cutoff sphere neutralization or the
176 distance-dependent damping technique could be used as a foundation for
177 a new pairwise summation method. Wolf \textit{et al.} made the
178 observation that charge neutralization within the cutoff sphere plays
179 a significant role in energy convergence; therefore we will begin our
180 analysis with the various shifted forms that maintain this charge
181 neutralization. We can evaluate the methods of Wolf
182 \textit{et al.} and Zahn \textit{et al.} by considering the standard
183 shifted potential,
184 \begin{equation}
185 v_\textrm{SP}(r) = \begin{cases}
186 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
187 R_\textrm{c}
188 \end{cases},
189 \label{eq:shiftingPotForm}
190 \end{equation}
191 and shifted force,
192 \begin{equation}
193 v_\textrm{SF}(r) = \begin{cases}
194 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
195 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
196 \end{cases},
197 \label{eq:shiftingForm}
198 \end{equation}
199 functions where $v(r)$ is the unshifted form of the potential, and
200 $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
201 that both the potential and the forces goes to zero at the cutoff
202 radius, while the Shifted Potential ({\sc sp}) form only ensures the
203 potential is smooth at the cutoff radius
204 ($R_\textrm{c}$).\cite{Allen87}
205
206 The forces associated with the shifted potential are simply the forces
207 of the unshifted potential itself (when inside the cutoff sphere),
208 \begin{equation}
209 F_{\textrm{SP}} = \left( \frac{d v(r)}{dr} \right),
210 \end{equation}
211 and are zero outside. Inside the cutoff sphere, the forces associated
212 with the shifted force form can be written,
213 \begin{equation}
214 F_{\textrm{SF}} = \left( \frac{d v(r)}{dr} \right) - \left(\frac{d
215 v(r)}{dr} \right)_{r=R_\textrm{c}}.
216 \end{equation}
217
218 If the potential ($v(r)$) is taken to be the normal Coulomb potential,
219 \begin{equation}
220 v(r) = \frac{q_i q_j}{r},
221 \label{eq:Coulomb}
222 \end{equation}
223 then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
224 al.}'s undamped prescription:
225 \begin{equation}
226 V_\textrm{SP}(r) =
227 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
228 r\leqslant R_\textrm{c},
229 \label{eq:WolfSP}
230 \end{equation}
231 with associated forces,
232 \begin{equation}
233 F_\textrm{SP}(r) = q_iq_j\left(-\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
234 \label{eq:FWolfSP}
235 \end{equation}
236 These forces are identical to the forces of the standard Coulomb
237 interaction, and cutting these off at $R_c$ was addressed by Wolf
238 \textit{et al.} as undesirable. They pointed out that the effect of
239 the image charges is neglected in the forces when this form is
240 used,\cite{Wolf99} thereby eliminating any benefit from the method in
241 molecular dynamics. Additionally, there is a discontinuity in the
242 forces at the cutoff radius which results in energy drift during MD
243 simulations.
244
245 The shifted force ({\sc sf}) form using the normal Coulomb potential
246 will give,
247 \begin{equation}
248 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}.
249 \label{eq:SFPot}
250 \end{equation}
251 with associated forces,
252 \begin{equation}
253 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}.
254 \label{eq:SFForces}
255 \end{equation}
256 This formulation has the benefits that there are no discontinuities at
257 the cutoff distance, while the neutralizing image charges are present
258 in both the energy and force expressions. It would be simple to add
259 the self-neutralizing term back when computing the total energy of the
260 system, thereby maintaining the agreement with the Madelung energies.
261 A side effect of this treatment is the alteration in the shape of the
262 potential that comes from the derivative term. Thus, a degree of
263 clarity about agreement with the empirical potential is lost in order
264 to gain functionality in dynamics simulations.
265
266 Wolf \textit{et al.} originally discussed the energetics of the
267 shifted Coulomb potential (Eq. \ref{eq:WolfSP}), and they found that
268 it was still insufficient for accurate determination of the energy
269 with reasonable cutoff distances. The calculated Madelung energies
270 fluctuate around the expected value with increasing cutoff radius, but
271 the oscillations converge toward the correct value.\cite{Wolf99} A
272 damping function was incorporated to accelerate the convergence; and
273 though alternative functional forms could be
274 used,\cite{Jones56,Heyes81} the complimentary error function was
275 chosen to mirror the effective screening used in the Ewald summation.
276 Incorporating this error function damping into the simple Coulomb
277 potential,
278 \begin{equation}
279 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
280 \label{eq:dampCoulomb}
281 \end{equation}
282 the shifted potential (Eq. \ref{eq:WolfSP}) can be recovered
283 \textit{via} equation \ref{eq:shiftingForm},
284 \begin{equation}
285 v_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}(\alpha r)}{r}-\frac{\textrm{erfc}(\alpha R_\textrm{c})}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c}.
286 \label{eq:DSPPot}
287 \end{equation},
288 with associated forces,
289 \begin{equation}
290 f_{\textrm{DSP}}(r) = q_iq_j\left(-\frac{\textrm{erfc}(\alpha r)}{r^2}-\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r^2)}}{r}\right) \quad r\leqslant R_\textrm{c}.
291 \label{eq:DSPForces}
292 \end{equation}
293 Again, this damped shifted potential suffers from a discontinuity and
294 a lack of the image charges in the forces. To remedy these concerns,
295 one may derive a Shifted-Force variant by including the derivative
296 term in equation \ref{eq:shiftingForm},
297 \begin{equation}
298 \begin{split}
299 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}.
300 \label{eq:DSFPot}
301 \end{split}
302 \end{equation}
303 The derivative of the above potential gives the following forces,
304 \begin{equation}
305 \begin{split}
306 f_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&-\left(\frac{\textrm{erfc}(\alpha r)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{(-\alpha^2r^2)}}{r}\right) \\ &\left.+\left(\frac{\textrm{erfc}(\alpha R_{\textrm{c}})}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
307 \label{eq:DSFForces}
308 \end{split}
309 \end{equation}
310
311 This new Shifted-Force potential is similar to equation
312 \ref{eq:ZahnPot} derived by Zahn \textit{et al.}; however, there are
313 two important differences.\cite{Zahn02} First, the $v_\textrm{c}$ term
314 from eq. (\ref{eq:shiftingForm}) is equal to
315 eq. (\ref{eq:dampCoulomb}) with $r$ replaced by $R_\textrm{c}$. This
316 term is {\it not} present in the Zahn potential, resulting in a
317 potential discontinuity as particles cross $R_\textrm{c}$. Second,
318 the sign of the derivative portion is different. The missing
319 $v_\textrm{c}$ term would not affect molecular dynamics simulations
320 (although the computed energy would be expected to have sudden jumps
321 as particle distances crossed $R_c$). The sign problem would be a
322 potential source of errors, however. In fact, it introduces a
323 discontinuity in the forces at the cutoff, because the force function
324 is shifted in the wrong direction and doesn't cross zero at
325 $R_\textrm{c}$.
326
327 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
328 electrostatic summation method that is continuous in both the
329 potential and forces and which incorporates the damping function
330 proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
331 paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
332 sf}, damping) are at reproducing the correct electrostatic summation
333 performed by the Ewald sum.
334
335 \subsection{Other alternatives}
336
337 Reaction Field blah
338
339 Group-based cutoff blah
340
341
342 \section{Methods}
343
344 In classical molecular mechanics simulations, there are two primary
345 techniques utilized to obtain information about the system of
346 interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
347 techniques utilize pairwise summations of interactions between
348 particle sites, but they use these summations in different ways.
349
350 In MC, the potential energy difference between two subsequent
351 configurations dictates the progression of MC sampling. Going back to
352 the origins of this method, the acceptance criterion for the canonical
353 ensemble laid out by Metropolis \textit{et al.} states that a
354 subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
355 \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
356 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
357 alternate method for handling the long-range electrostatics will
358 ensure proper sampling from the ensemble.
359
360 In MD, the derivative of the potential governs how the system will
361 progress in time. Consequently, the force and torque vectors on each
362 body in the system dictate how the system evolves. If the magnitude
363 and direction of these vectors are similar when using alternate
364 electrostatic summation techniques, the dynamics in the short term
365 will be indistinguishable. Because error in MD calculations is
366 cumulative, one should expect greater deviation at longer times,
367 although methods which have large differences in the force and torque
368 vectors will diverge from each other more rapidly.
369
370 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
371 The pairwise summation techniques (outlined in section
372 \ref{sec:ESMethods}) were evaluated for use in MC simulations by
373 studying the energy differences between conformations. We took the
374 SPME-computed energy difference between two conformations to be the
375 correct behavior. An ideal performance by an alternative method would
376 reproduce these energy differences exactly. Since none of the methods
377 provide exact energy differences, we used linear least squares
378 regressions of the $\Delta E$ values between configurations using SPME
379 against $\Delta E$ values using tested methods provides a quantitative
380 comparison of this agreement. Unitary results for both the
381 correlation and correlation coefficient for these regressions indicate
382 equivalent energetic results between the method under consideration
383 and electrostatics handled using SPME. Sample correlation plots for
384 two alternate methods are shown in Fig. \ref{fig:linearFit}.
385
386 \begin{figure}
387 \centering
388 \includegraphics[width = \linewidth]{./dualLinear.pdf}
389 \caption{Example least squares regressions of the configuration energy differences for SPC/E water systems. The upper plot shows a data set with a poor correlation coefficient ($R^2$), while the lower plot shows a data set with a good correlation coefficient.}
390 \label{fig:linearFit}
391 \end{figure}
392
393 Each system type (detailed in section \ref{sec:RepSims}) was
394 represented using 500 independent configurations. Additionally, we
395 used seven different system types, so each of the alternate
396 (non-Ewald) electrostatic summation methods was evaluated using
397 873,250 configurational energy differences.
398
399 Results and discussion for the individual analysis of each of the
400 system types appear in the supporting information, while the
401 cumulative results over all the investigated systems appears below in
402 section \ref{sec:EnergyResults}.
403
404 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
405 We evaluated the pairwise methods (outlined in section
406 \ref{sec:ESMethods}) for use in MD simulations by
407 comparing the force and torque vectors with those obtained using the
408 reference Ewald summation (SPME). Both the magnitude and the
409 direction of these vectors on each of the bodies in the system were
410 analyzed. For the magnitude of these vectors, linear least squares
411 regression analyses were performed as described previously for
412 comparing $\Delta E$ values. Instead of a single energy difference
413 between two system configurations, we compared the magnitudes of the
414 forces (and torques) on each molecule in each configuration. For a
415 system of 1000 water molecules and 40 ions, there are 1040 force
416 vectors and 1000 torque vectors. With 500 configurations, this
417 results in 520,000 force and 500,000 torque vector comparisons.
418 Additionally, data from seven different system types was aggregated
419 before the comparison was made.
420
421 The {\it directionality} of the force and torque vectors was
422 investigated through measurement of the angle ($\theta$) formed
423 between those computed from the particular method and those from SPME,
424 \begin{equation}
425 \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
426 \end{equation}
427 where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
428 force vector computed using method $M$.
429
430 Each of these $\theta$ values was accumulated in a distribution
431 function, weighted by the area on the unit sphere. Non-linear
432 Gaussian fits were used to measure the width of the resulting
433 distributions.
434
435 \begin{figure}
436 \centering
437 \includegraphics[width = \linewidth]{./gaussFit.pdf}
438 \caption{Sample fit of the angular distribution of the force vectors over all of the studied systems. Gaussian fits were used to obtain values for the variance in force and torque vectors used in the following figure.}
439 \label{fig:gaussian}
440 \end{figure}
441
442 Figure \ref{fig:gaussian} shows an example distribution with applied
443 non-linear fits. The solid line is a Gaussian profile, while the
444 dotted line is a Voigt profile, a convolution of a Gaussian and a
445 Lorentzian. Since this distribution is a measure of angular error
446 between two different electrostatic summation methods, there is no
447 {\it a priori} reason for the profile to adhere to any specific shape.
448 Gaussian fits was used to compare all the tested methods. The
449 variance ($\sigma^2$) was extracted from each of these fits and was
450 used to compare distribution widths. Values of $\sigma^2$ near zero
451 indicate vector directions indistinguishable from those calculated
452 when using the reference method (SPME).
453
454 \subsection{Short-time Dynamics}
455
456 \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
457 Evaluation of the long-time dynamics of charged systems was performed
458 by considering the NaCl crystal system while using a subset of the
459 best performing pairwise methods. The NaCl crystal was chosen to
460 avoid possible complications involving the propagation techniques of
461 orientational motion in molecular systems. To enhance the atomic
462 motion, these crystals were equilibrated at 1000 K, near the
463 experimental $T_m$ for NaCl. Simulations were performed under the
464 microcanonical ensemble, and velocity autocorrelation functions
465 (Eq. \ref{eq:vCorr}) were computed for each of the trajectories,
466 \begin{equation}
467 C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
468 \label{eq:vCorr}
469 \end{equation}
470 Velocity autocorrelation functions require detailed short time data
471 and long trajectories for good statistics, thus velocity information
472 was saved every 5 fs over 100 ps trajectories. The power spectrum
473 ($I(\omega)$) is obtained via Fourier transform of the autocorrelation
474 function
475 \begin{equation}
476 I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
477 \label{eq:powerSpec}
478 \end{equation}
479 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$.
480
481 \subsection{Representative Simulations}\label{sec:RepSims}
482 A variety of common and representative simulations were analyzed to
483 determine the relative effectiveness of the pairwise summation
484 techniques in reproducing the energetics and dynamics exhibited by
485 SPME. The studied systems were as follows:
486 \begin{enumerate}
487 \item Liquid Water
488 \item Crystalline Water (Ice I$_\textrm{c}$)
489 \item NaCl Crystal
490 \item NaCl Melt
491 \item Low Ionic Strength Solution of NaCl in Water
492 \item High Ionic Strength Solution of NaCl in Water
493 \item 6 \AA\ Radius Sphere of Argon in Water
494 \end{enumerate}
495 By utilizing the pairwise techniques (outlined in section
496 \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
497 charged particles, and mixtures of the two, we can comment on possible
498 system dependence and/or universal applicability of the techniques.
499
500 Generation of the system configurations was dependent on the system
501 type. For the solid and liquid water configurations, configuration
502 snapshots were taken at regular intervals from higher temperature 1000
503 SPC/E water molecule trajectories and each equilibrated individually.
504 The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
505 ions and were selected and equilibrated in the same fashion as the
506 water systems. For the low and high ionic strength NaCl solutions, 4
507 and 40 ions were first solvated in a 1000 water molecule boxes
508 respectively. Ion and water positions were then randomly swapped, and
509 the resulting configurations were again equilibrated individually.
510 Finally, for the Argon/Water "charge void" systems, the identities of
511 all the SPC/E waters within 6 \AA\ of the center of the equilibrated
512 water configurations were converted to argon
513 (Fig. \ref{fig:argonSlice}).
514
515 \begin{figure}
516 \centering
517 \includegraphics[width = \linewidth]{./slice.pdf}
518 \caption{A slice from the center of a water box used in a charge void simulation. The darkened region represents the boundary sphere within which the water molecules were converted to argon atoms.}
519 \label{fig:argonSlice}
520 \end{figure}
521
522 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
523 Electrostatic summation method comparisons were performed using SPME,
524 the Shifted-Potential and Shifted-Force methods - both with damping
525 parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
526 moderate, and strong damping respectively), reaction field with an
527 infinite dielectric constant, and an unmodified cutoff. Group-based
528 cutoffs with a fifth-order polynomial switching function were
529 necessary for the reaction field simulations and were utilized in the
530 SP, SF, and pure cutoff methods for comparison to the standard lack of
531 group-based cutoffs with a hard truncation. The SPME calculations
532 were performed using the TINKER implementation of SPME,\cite{Ponder87}
533 while all other method calculations were performed using the OOPSE
534 molecular mechanics package.\cite{Meineke05}
535
536 These methods were additionally evaluated with three different cutoff
537 radii (9, 12, and 15 \AA) to investigate possible cutoff radius
538 dependence. It should be noted that the damping parameter chosen in
539 SPME, or so called ``Ewald Coefficient", has a significant effect on
540 the energies and forces calculated. Typical molecular mechanics
541 packages default this to a value dependent on the cutoff radius and a
542 tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller
543 tolerances are typically associated with increased accuracy in the
544 real-space portion of the summation.\cite{Essmann95} The default
545 TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
546 calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
547 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
548
549 \section{Results and Discussion}
550
551 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
552 In order to evaluate the performance of the pairwise electrostatic
553 summation methods for Monte Carlo simulations, the energy differences
554 between configurations were compared to the values obtained when using
555 SPME. The results for the subsequent regression analysis are shown in
556 figure \ref{fig:delE}.
557
558 \begin{figure}
559 \centering
560 \includegraphics[width=5.5in]{./delEplot.pdf}
561 \caption{Statistical analysis of the quality of configurational energy differences for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate $\Delta E$ values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
562 \label{fig:delE}
563 \end{figure}
564
565 In this figure, it is apparent that it is unreasonable to expect
566 realistic results using an unmodified cutoff. This is not all that
567 surprising since this results in large energy fluctuations as atoms
568 move in and out of the cutoff radius. These fluctuations can be
569 alleviated to some degree by using group based cutoffs with a
570 switching function.\cite{Steinbach94} The Group Switch Cutoff row
571 doesn't show a significant improvement in this plot because the salt
572 and salt solution systems contain non-neutral groups, see the
573 accompanying supporting information for a comparison where all groups
574 are neutral.
575
576 Correcting the resulting charged cutoff sphere is one of the purposes
577 of the damped Coulomb summation proposed by Wolf \textit{et
578 al.},\cite{Wolf99} and this correction indeed improves the results as
579 seen in the Shifted-Potental rows. While the undamped case of this
580 method is a significant improvement over the pure cutoff, it still
581 doesn't correlate that well with SPME. Inclusion of potential damping
582 improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
583 an excellent correlation and quality of fit with the SPME results,
584 particularly with a cutoff radius greater than 12 \AA . Use of a
585 larger damping parameter is more helpful for the shortest cutoff
586 shown, but it has a detrimental effect on simulations with larger
587 cutoffs. In the Shifted-Force sets, increasing damping results in
588 progressively poorer correlation. Overall, the undamped case is the
589 best performing set, as the correlation and quality of fits are
590 consistently superior regardless of the cutoff distance. This result
591 is beneficial in that the undamped case is less computationally
592 prohibitive do to the lack of complimentary error function calculation
593 when performing the electrostatic pair interaction. The reaction
594 field results illustrates some of that method's limitations, primarily
595 that it was developed for use in homogenous systems; although it does
596 provide results that are an improvement over those from an unmodified
597 cutoff.
598
599 \subsection{Magnitudes of the Force and Torque Vectors}
600
601 Evaluation of pairwise methods for use in Molecular Dynamics
602 simulations requires consideration of effects on the forces and
603 torques. Investigation of the force and torque vector magnitudes
604 provides a measure of the strength of these values relative to SPME.
605 Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
606 force and torque vector magnitude regression results for the
607 accumulated analysis over all the system types.
608
609 \begin{figure}
610 \centering
611 \includegraphics[width=5.5in]{./frcMagplot.pdf}
612 \caption{Statistical analysis of the quality of the force vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate force magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
613 \label{fig:frcMag}
614 \end{figure}
615
616 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
617 in the previous $\Delta E$ section. The unmodified cutoff results are
618 poor, but using group based cutoffs and a switching function provides
619 a improvement much more significant than what was seen with $\Delta
620 E$. Looking at the Shifted-Potential sets, the slope and $R^2$
621 improve with the use of damping to an optimal result of 0.2 \AA
622 $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping,
623 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
624 detrimental to simulations with larger cutoff radii. The undamped
625 Shifted-Force method gives forces in line with those obtained using
626 SPME, and use of a damping function results in minor improvement. The
627 reaction field results are surprisingly good, considering the poor
628 quality of the fits for the $\Delta E$ results. There is still a
629 considerable degree of scatter in the data, but it correlates well in
630 general. To be fair, we again note that the reaction field
631 calculations do not encompass NaCl crystal and melt systems, so these
632 results are partly biased towards conditions in which the method
633 performs more favorably.
634
635 \begin{figure}
636 \centering
637 \includegraphics[width=5.5in]{./trqMagplot.pdf}
638 \caption{Statistical analysis of the quality of the torque vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate torque magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
639 \label{fig:trqMag}
640 \end{figure}
641
642 To evaluate the torque vector magnitudes, the data set from which
643 values are drawn is limited to rigid molecules in the systems
644 (i.e. water molecules). In spite of this smaller sampling pool, the
645 torque vector magnitude results in figure \ref{fig:trqMag} are still
646 similar to those seen for the forces; however, they more clearly show
647 the improved behavior that comes with increasing the cutoff radius.
648 Moderate damping is beneficial to the Shifted-Potential and helpful
649 yet possibly unnecessary with the Shifted-Force method, and they also
650 show that over-damping adversely effects all cutoff radii rather than
651 showing an improvement for systems with short cutoffs. The reaction
652 field method performs well when calculating the torques, better than
653 the Shifted Force method over this limited data set.
654
655 \subsection{Directionality of the Force and Torque Vectors}
656
657 Having force and torque vectors with magnitudes that are well
658 correlated to SPME is good, but if they are not pointing in the proper
659 direction the results will be incorrect. These vector directions were
660 investigated through measurement of the angle formed between them and
661 those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared
662 through the variance ($\sigma^2$) of the Gaussian fits of the angle
663 error distributions of the combined set over all system types.
664
665 \begin{figure}
666 \centering
667 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
668 \caption{Statistical analysis of the quality of the Gaussian fit of the force and torque vector angular distributions for a given electrostatic method compared with the reference Ewald sum. Results with a variance ($\sigma^2$) equal to zero (dashed line) indicate force and torque directions indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
669 \label{fig:frcTrqAng}
670 \end{figure}
671
672 Both the force and torque $\sigma^2$ results from the analysis of the
673 total accumulated system data are tabulated in figure
674 \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case
675 show the improvement afforded by choosing a longer simulation cutoff.
676 Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
677 of the distribution widths, with a similar improvement going from 12
678 to 15 \AA . The undamped Shifted-Force, Group Based Cutoff, and
679 Reaction Field methods all do equivalently well at capturing the
680 direction of both the force and torque vectors. Using damping
681 improves the angular behavior significantly for the Shifted-Potential
682 and moderately for the Shifted-Force methods. Increasing the damping
683 too far is destructive for both methods, particularly to the torque
684 vectors. Again it is important to recognize that the force vectors
685 cover all particles in the systems, while torque vectors are only
686 available for neutral molecular groups. Damping appears to have a
687 more beneficial effect on non-neutral bodies, and this observation is
688 investigated further in the accompanying supporting information.
689
690 \begin{table}[htbp]
691 \centering
692 \caption{Variance ($\sigma^2$) of the force (top set) and torque (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods. Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function. The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}
693 \begin{tabular}{@{} ccrrrrrrrr @{}}
694 \\
695 \toprule
696 & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
697 \cmidrule(lr){3-6}
698 \cmidrule(l){7-10}
699 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
700 \midrule
701
702 9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
703 & \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} \\
704 12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
705 & \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} \\
706 15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
707 & \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} \\
708
709 \midrule
710
711 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
712 & \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} \\
713 12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
714 & \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} \\
715 15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
716 & \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} \\
717
718 \bottomrule
719 \end{tabular}
720 \label{tab:groupAngle}
721 \end{table}
722
723 Although not discussed previously, group based cutoffs can be applied
724 to both the Shifted-Potential and Shifted-Force methods. Use off a
725 switching function corrects for the discontinuities that arise when
726 atoms of a group exit the cutoff before the group's center of mass.
727 Though there are no significant benefit or drawbacks observed in
728 $\Delta E$ and vector magnitude results when doing this, there is a
729 measurable improvement in the vector angle results. Table
730 \ref{tab:groupAngle} shows the angular variance values obtained using
731 group based cutoffs and a switching function alongside the standard
732 results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
733 The Shifted-Potential shows much narrower angular distributions for
734 both the force and torque vectors when using an $\alpha$ of 0.2
735 \AA$^{-1}$ or less, while Shifted-Force shows improvements in the
736 undamped and lightly damped cases. Thus, by calculating the
737 electrostatic interactions in terms of molecular pairs rather than
738 atomic pairs, the direction of the force and torque vectors are
739 determined more accurately.
740
741 One additional trend to recognize in table \ref{tab:groupAngle} is
742 that the $\sigma^2$ values for both Shifted-Potential and
743 Shifted-Force converge as $\alpha$ increases, something that is easier
744 to see when using group based cutoffs. Looking back on figures
745 \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
746 behavior clearly at large $\alpha$ and cutoff values. The reason for
747 this is that the complimentary error function inserted into the
748 potential weakens the electrostatic interaction as $\alpha$ increases.
749 Thus, at larger values of $\alpha$, both the summation method types
750 progress toward non-interacting functions, so care is required in
751 choosing large damping functions lest one generate an undesirable loss
752 in the pair interaction. Kast \textit{et al.} developed a method for
753 choosing appropriate $\alpha$ values for these types of electrostatic
754 summation methods by fitting to $g(r)$ data, and their methods
755 indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
756 values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
757 to be reasonable choices to obtain proper MC behavior
758 (Fig. \ref{fig:delE}); however, based on these findings, choices this
759 high would introduce error in the molecular torques, particularly for
760 the shorter cutoffs. Based on the above findings, empirical damping
761 up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
762 unnecessary when using the Shifted-Force method.
763
764 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
765
766 In the previous studies using a Shifted-Force variant of the damped
767 Wolf coulomb potential, the structure and dynamics of water were
768 investigated rather extensively.\cite{Zahn02,Kast03} Their results
769 indicated that the damped Shifted-Force method results in properties
770 very similar to those obtained when using the Ewald summation.
771 Considering the statistical results shown above, the good performance
772 of this method is not that surprising. Rather than consider the same
773 systems and simply recapitulate their results, we decided to look at
774 the solid state dynamical behavior obtained using the best performing
775 summation methods from the above results.
776
777 \begin{figure}
778 \centering
779 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
780 \caption{Power spectra obtained from the velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, Shifted-Force ($\alpha$ = 0, 0.1, \& 0.2), and Shifted-Potential ($\alpha$ = 0.2). Apodization of the correlation functions via a cubic switching function between 40 and 50 ps was used to clear up the spectral noise resulting from data truncation, and had no noticeable effect on peak location or magnitude. The inset shows the frequency region below 100 cm$^{-1}$ to highlight where the spectra begin to differ.}
781 \label{fig:methodPS}
782 \end{figure}
783
784 Figure \ref{fig:methodPS} shows the power spectra for the NaCl
785 crystals (from averaged Na and Cl ion velocity autocorrelation
786 functions) using the stated electrostatic summation methods. While
787 high frequency peaks of all the spectra overlap, showing the same
788 general features, the low frequency region shows how the summation
789 methods differ. Considering the low-frequency inset (expanded in the
790 upper frame of figure \ref{fig:dampInc}), at frequencies below 100
791 cm$^{-1}$, the correlated motions are blue-shifted when using undamped
792 or weakly damped Shifted-Force. When using moderate damping ($\alpha
793 = 0.2$ \AA$^{-1}$) both the Shifted-Force and Shifted-Potential
794 methods give near identical correlated motion behavior as the Ewald
795 method (which has a damping value of 0.3119). The damping acts as a
796 distance dependent Gaussian screening of the point charges for the
797 pairwise summation methods. This weakening of the electrostatic
798 interaction with distance explains why the long-ranged correlated
799 motions are at lower frequencies for the moderately damped methods
800 than for undamped or weakly damped methods. To see this effect more
801 clearly, we show how damping strength affects a simple real-space
802 electrostatic potential,
803 \begin{equation}
804 V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
805 \end{equation}
806 where $S(r)$ is a switching function that smoothly zeroes the
807 potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
808 the low frequency motions are dependent on the damping used in the
809 direct electrostatic sum. As the damping increases, the peaks drop to
810 lower frequencies. Incidentally, use of an $\alpha$ of 0.25
811 \AA$^{-1}$ on a simple electrostatic summation results in low
812 frequency correlated dynamics equivalent to a simulation using SPME.
813 When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
814 shift to higher frequency in exponential fashion. Though not shown,
815 the spectrum for the simple undamped electrostatic potential is
816 blue-shifted such that the lowest frequency peak resides near 325
817 cm$^{-1}$. In light of these results, the undamped Shifted-Force
818 method producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is
819 quite respectable; however, it appears as though moderate damping is
820 required for accurate reproduction of crystal dynamics.
821 \begin{figure}
822 \centering
823 \includegraphics[width = \linewidth]{./comboSquare.pdf}
824 \caption{Upper: Zoomed inset from figure \ref{fig:methodPS}. As the damping value for the Shifted-Force potential increases, the low-frequency peaks red-shift. Lower: Low-frequency correlated motions for NaCl crystals at 1000 K when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$. As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME. The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
825 \label{fig:dampInc}
826 \end{figure}
827
828 \section{Conclusions}
829
830 This investigation of pairwise electrostatic summation techniques
831 shows that there are viable and more computationally efficient
832 electrostatic summation techniques than the Ewald summation, chiefly
833 methods derived from the damped Coulombic sum originally proposed by
834 Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
835 Shifted-Force method, reformulated above as equation \ref{eq:SFPot},
836 shows a remarkable ability to reproduce the energetic and dynamic
837 characteristics exhibited by simulations employing lattice summation
838 techniques. The cumulative energy difference results showed the
839 undamped Shifted-Force and moderately damped Shifted-Potential methods
840 produced results nearly identical to SPME. Similarly for the dynamic
841 features, the undamped or moderately damped Shifted-Force and
842 moderately damped Shifted-Potential methods produce force and torque
843 vector magnitude and directions very similar to the expected values.
844 These results translate into long-time dynamic behavior equivalent to
845 that produced in simulations using SPME.
846
847 Aside from the computational cost benefit, these techniques have
848 applicability in situations where the use of the Ewald sum can prove
849 problematic. Primary among them is their use in interfacial systems,
850 where the unmodified lattice sum techniques artificially accentuate
851 the periodicity of the system in an undesirable manner. There have
852 been alterations to the standard Ewald techniques, via corrections and
853 reformulations, to compensate for these systems; but the pairwise
854 techniques discussed here require no modifications, making them
855 natural tools to tackle these problems. Additionally, this
856 transferability gives them benefits over other pairwise methods, like
857 reaction field, because estimations of physical properties (e.g. the
858 dielectric constant) are unnecessary.
859
860 We are not suggesting any flaw with the Ewald sum; in fact, it is the
861 standard by which these simple pairwise sums are judged. However,
862 these results do suggest that in the typical simulations performed
863 today, the Ewald summation may no longer be required to obtain the
864 level of accuracy most researcher have come to expect
865
866 \section{Acknowledgments}
867 \newpage
868
869 \bibliographystyle{jcp2}
870 \bibliography{electrostaticMethods}
871
872
873 \end{document}