ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
(Generate patch)

Comparing trunk/electrostaticMethodsPaper/electrostaticMethods.tex (file contents):
Revision 2599 by chrisfen, Tue Feb 28 14:09:55 2006 UTC vs.
Revision 2637 by chrisfen, Sun Mar 19 02:48:19 2006 UTC

# Line 1 | Line 1
1   %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 < \documentclass[12pt]{article}
2 > %\documentclass[aps,prb,preprint]{revtex4}
3 > \documentclass[11pt]{article}
4   \usepackage{endfloat}
5 < \usepackage{amsmath}
5 > \usepackage{amsmath,bm}
6   \usepackage{amssymb}
6 %\usepackage{ifsym}
7   \usepackage{epsf}
8   \usepackage{times}
9 < \usepackage{mathptm}
9 > \usepackage{mathptmx}
10   \usepackage{setspace}
11   \usepackage{tabularx}
12   \usepackage{graphicx}
13   \usepackage{booktabs}
14 < %\usepackage{berkeley}
14 > \usepackage{bibentry}
15 > \usepackage{mathrsfs}
16   \usepackage[ref]{overcite}
17   \pagestyle{plain}
18   \pagenumbering{arabic}
# Line 24 | Line 25
25  
26   \begin{document}
27  
28 < \title{Is the Ewald Summation necessary in typical molecular simulations: Alternatives to the accepted standard of cutoff policies}
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 \\
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}
# Line 34 | Line 36 | Notre Dame, Indiana 46556}
36   \date{\today}
37  
38   \maketitle
39 < %\doublespacing
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 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62   %                              BODY OF TEXT
63 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64  
65   \section{Introduction}
66  
67 < In this paper, a variety of simulation situations were analyzed to determine the relative effectiveness of the adapted Wolf spherical truncation schemes at reproducing the results obtained using a smooth particle mesh Ewald (SPME) summation technique.  In addition to the Shifted-Potential and Shifted-Force adapted Wolf methods, both reaction field and uncorrected cutoff methods were included for comparison purposes.  The general usability of these methods in both Monte Carlo and Molecular Dynamics calculations was assessed through statistical analysis over the combined results from all of the following studied systems:
68 < \begin{enumerate}
69 < \item Liquid Water
70 < \item Crystalline Water (Ice I$_\textrm{c}$)
71 < \item NaCl Crystal
72 < \item NaCl Melt
73 < \item Low Ionic Strength Solution of NaCl in Water
74 < \item High Ionic Strength Solution of NaCl in Water
75 < \item 6 \AA\  Radius Sphere of Argon in Water
76 < \end{enumerate}
77 < Additional discussion on the results from the individual systems was also performed to identify limitations of the considered methods in specific systems.
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 < \section{Methods}
83 > \subsection{The Ewald Sum}
84 > The complete accumulation electrostatic interactions in a system with periodic boundary conditions (PBC) requires the consideration of the effect of all charges within a simulation box, as well as those in the periodic replicas,
85 > \begin{equation}
86 > 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],
87 > \label{eq:PBCSum}
88 > \end{equation}
89 > where the sum over $\mathbf{n}$ is a sum over all periodic box replicas
90 > with integer coordinates $\mathbf{n} = (l,m,n)$, and the prime indicates
91 > $i = j$ are neglected for $\mathbf{n} = 0$.\cite{deLeeuw80} Within the
92 > sum, $N$ is the number of electrostatic particles, $\mathbf{r}_{ij}$ is
93 > $\mathbf{r}_j - \mathbf{r}_i$, $L$ is the cell length, $\bm{\Omega}_{i,j}$ are
94 > the Euler angles for $i$ and $j$, and $\phi$ is Poisson's equation
95 > ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for charge-charge
96 > interactions). In the case of monopole electrostatics,
97 > eq. (\ref{eq:PBCSum}) is conditionally convergent and is discontiuous
98 > for non-neutral systems.
99  
100 < In each of the simulated systems, 500 distinct configurations were generated, and the electrostatic summation methods were compared via sequential application on each of these fixed configurations.  The methods compared include SPME, the aforementioned Shifted Potential and Shifted Force methods - both with damping parameters ($\alpha$) of 0, 0.1, 0.2, and 0.3 \AA$^{-1}$, reaction field with an infinite dielectric constant, and an unmodified cutoff.  Group-based cutoffs with a fifth-order polynomial switching function were necessary for the reaction field simulations and were utilized in the SP, SF, and pure cutoff methods for comparison to the standard lack of group-based cutoffs with a hard truncation.  
100 > This electrostatic summation problem was originally studied by Ewald
101 > for the case of an infinite crystal.\cite{Ewald21}. The approach he
102 > took was to convert this conditionally convergent sum into two
103 > absolutely convergent summations: a short-ranged real-space summation
104 > and a long-ranged reciprocal-space summation,
105 > \begin{equation}
106 > \begin{split}
107 > 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,
108 > \end{split}
109 > \label{eq:EwaldSum}
110 > \end{equation}
111 > where $\alpha$ is a damping parameter, or separation constant, with
112 > units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and equal
113 > $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
114 > constant of the encompassing medium. The final two terms of
115 > eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
116 > for interacting with a surrounding dielectric.\cite{Allen87} This
117 > dipolar term was neglected in early applications in molecular
118 > simulations,\cite{Brush66,Woodcock71} until it was introduced by de
119 > Leeuw {\it et al.} to address situations where the unit cell has a
120 > dipole moment and this dipole moment gets magnified through
121 > replication of the periodic images.\cite{deLeeuw80,Smith81} If this
122 > term is taken to be zero, the system is using conducting boundary
123 > conditions, $\epsilon_{\rm S} = \infty$. Figure
124 > \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
125 > time.  Initially, due to the small size of systems, the entire
126 > simulation box was replicated to convergence.  Currently, we balance a
127 > spherical real-space cutoff with the reciprocal sum and consider the
128 > surrounding dielectric.
129 > \begin{figure}
130 > \centering
131 > \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
132 > \caption{How the application of the Ewald summation has changed with
133 > the increase in computer power.  Initially, only small numbers of
134 > particles could be studied, and the Ewald sum acted to replicate the
135 > unit cell charge distribution out to convergence.  Now, much larger
136 > systems of charges are investigated with fixed distance cutoffs.  The
137 > calculated structure factor is used to sum out to great distance, and
138 > a surrounding dielectric term is included.}
139 > \label{fig:ewaldTime}
140 > \end{figure}
141  
142 < Generation of the system configurations was dependent on the system type.  For the solid and liquid water configurations, configuration snapshots were taken at regular intervals from higher temperature 1000 SPC/E water molecule trajectories and each equilibrated individually.  The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl- ions and were selected and equilibrated in the same fashion as the water systems.  For the low and high ionic strength NaCl solutions, 4 and 40 ions were first solvated in a 1000 water molecule boxes respectively.  Ion and water positions were then randomly swapped, and the resulting configurations were again equilibrated individually.  Finally, for the Argon/Water "charge void" systems, the identities of all the SPC/E waters within 6 \AA\ of the center of the equilibrated water configurations were converted to argon (Fig. \ref{argonSlice}).
142 > The Ewald summation in the straight-forward form is an
143 > $\mathscr{O}(N^2)$ algorithm.  The separation constant $(\alpha)$
144 > plays an important role in the computational cost balance between the
145 > direct and reciprocal-space portions of the summation.  The choice of
146 > the magnitude of this value allows one to select whether the
147 > real-space or reciprocal space portion of the summation is an
148 > $\mathscr{O}(N^2)$ calcualtion (with the other being
149 > $\mathscr{O}(N)$).\cite{Sagui99} With appropriate choice of $\alpha$
150 > and thoughtful algorithm development, this cost can be brought down to
151 > $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
152 > reduce the cost of the Ewald summation further is to set $\alpha$ such
153 > that the real-space interactions decay rapidly, allowing for a short
154 > spherical cutoff, and then optimize the reciprocal space summation.
155 > These optimizations usually involve the utilization of the fast
156 > Fourier transform (FFT),\cite{Hockney81} leading to the
157 > particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
158 > methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
159 > methods, the cost of the reciprocal-space portion of the Ewald
160 > summation is from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N \log N)$.
161 >
162 > These developments and optimizations have led the use of the Ewald
163 > summation to become routine in simulations with periodic boundary
164 > conditions. However, in certain systems the intrinsic three
165 > dimensional periodicity can prove to be problematic, such as two
166 > dimensional surfaces and membranes.  The Ewald sum has been
167 > reformulated to handle 2D
168 > systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the new
169 > methods have been found to be computationally
170 > expensive.\cite{Spohr97,Yeh99} Inclusion of a correction term in the
171 > full Ewald summation is a possible direction for enabling the handling
172 > of 2D systems and the inclusion of the optimizations described
173 > previously.\cite{Yeh99}
174 >
175 > Several studies have recognized that the inherent periodicity in the
176 > Ewald sum can also have an effect on systems that have the same
177 > dimensionality.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
178 > Good examples are solvated proteins kept at high relative
179 > concentration due to the periodicity of the electrostatics.  In these
180 > systems, the more compact folded states of a protein can be
181 > artificially stabilized by the periodic replicas introduced by the
182 > Ewald summation.\cite{Weber00} Thus, care ought to be taken when
183 > considering the use of the Ewald summation where the intrinsic
184 > perodicity may negatively affect the system dynamics.
185 >
186 >
187 > \subsection{The Wolf and Zahn Methods}
188 > In a recent paper by Wolf \textit{et al.}, a procedure was outlined
189 > for the accurate accumulation of electrostatic interactions in an
190 > efficient pairwise fashion and lacks the inherent periodicity of the
191 > Ewald summation.\cite{Wolf99} Wolf \textit{et al.} observed that the
192 > electrostatic interaction is effectively short-ranged in condensed
193 > phase systems and that neutralization of the charge contained within
194 > the cutoff radius is crucial for potential stability. They devised a
195 > pairwise summation method that ensures charge neutrality and gives
196 > results similar to those obtained with the Ewald summation.  The
197 > resulting shifted Coulomb potential (Eq. \ref{eq:WolfPot}) includes
198 > image-charges subtracted out through placement on the cutoff sphere
199 > and a distance-dependent damping function (identical to that seen in
200 > the real-space portion of the Ewald sum) to aid convergence
201 > \begin{equation}
202 > 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\}.
203 > \label{eq:WolfPot}
204 > \end{equation}
205 > Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
206 > potential.  However, neutralizing the charge contained within each
207 > cutoff sphere requires the placement of a self-image charge on the
208 > surface of the cutoff sphere.  This additional self-term in the total
209 > potential enabled Wolf {\it et al.}  to obtain excellent estimates of
210 > Madelung energies for many crystals.
211 >
212 > In order to use their charge-neutralized potential in molecular
213 > dynamics simulations, Wolf \textit{et al.} suggested taking the
214 > derivative of this potential prior to evaluation of the limit.  This
215 > procedure gives an expression for the forces,
216 > \begin{equation}
217 > 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\},
218 > \label{eq:WolfForces}
219 > \end{equation}
220 > that incorporates both image charges and damping of the electrostatic
221 > interaction.
222 >
223 > More recently, Zahn \textit{et al.} investigated these potential and
224 > force expressions for use in simulations involving water.\cite{Zahn02}
225 > In their work, they pointed out that the forces and derivative of
226 > the potential are not commensurate.  Attempts to use both
227 > Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
228 > to poor energy conservation.  They correctly observed that taking the
229 > limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
230 > derivatives gives forces for a different potential energy function
231 > than the one shown in Eq. (\ref{eq:WolfPot}).
232 >
233 > Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
234 > method'' as a way to use this technique in Molecular Dynamics
235 > simulations.  Taking the integral of the forces shown in equation
236 > \ref{eq:WolfForces}, they proposed a new damped Coulomb
237 > potential,
238 > \begin{equation}
239 > 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\}.
240 > \label{eq:ZahnPot}
241 > \end{equation}
242 > They showed that this potential does fairly well at capturing the
243 > structural and dynamic properties of water compared the same
244 > properties obtained using the Ewald sum.
245 >
246 > \subsection{Simple Forms for Pairwise Electrostatics}
247 >
248 > The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
249 > al.} are constructed using two different (and separable) computational
250 > tricks: \begin{enumerate}
251 > \item shifting through the use of image charges, and
252 > \item damping the electrostatic interaction.
253 > \end{enumerate}  Wolf \textit{et al.} treated the
254 > development of their summation method as a progressive application of
255 > these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
256 > their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
257 > post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
258 > both techniques.  It is possible, however, to separate these
259 > tricks and study their effects independently.
260  
261 + Starting with the original observation that the effective range of the
262 + electrostatic interaction in condensed phases is considerably less
263 + than $r^{-1}$, either the cutoff sphere neutralization or the
264 + distance-dependent damping technique could be used as a foundation for
265 + a new pairwise summation method.  Wolf \textit{et al.} made the
266 + observation that charge neutralization within the cutoff sphere plays
267 + a significant role in energy convergence; therefore we will begin our
268 + analysis with the various shifted forms that maintain this charge
269 + neutralization.  We can evaluate the methods of Wolf
270 + \textit{et al.}  and Zahn \textit{et al.} by considering the standard
271 + shifted potential,
272 + \begin{equation}
273 + v_\textrm{SP}(r) =      \begin{cases}
274 + v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
275 + R_\textrm{c}  
276 + \end{cases},
277 + \label{eq:shiftingPotForm}
278 + \end{equation}
279 + and shifted force,
280 + \begin{equation}
281 + v_\textrm{SF}(r) =      \begin{cases}
282 + v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
283 + &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
284 +                                                \end{cases},
285 + \label{eq:shiftingForm}
286 + \end{equation}
287 + functions where $v(r)$ is the unshifted form of the potential, and
288 + $v_c$ is $v(R_\textrm{c})$.  The Shifted Force ({\sc sf}) form ensures
289 + that both the potential and the forces goes to zero at the cutoff
290 + radius, while the Shifted Potential ({\sc sp}) form only ensures the
291 + potential is smooth at the cutoff radius
292 + ($R_\textrm{c}$).\cite{Allen87}
293 +
294 + The forces associated with the shifted potential are simply the forces
295 + of the unshifted potential itself (when inside the cutoff sphere),
296 + \begin{equation}
297 + f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
298 + \end{equation}
299 + and are zero outside.  Inside the cutoff sphere, the forces associated
300 + with the shifted force form can be written,
301 + \begin{equation}
302 + f_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
303 + v(r)}{dr} \right)_{r=R_\textrm{c}}.
304 + \end{equation}
305 +
306 + If the potential ($v(r)$) is taken to be the normal Coulomb potential,
307 + \begin{equation}
308 + v(r) = \frac{q_i q_j}{r},
309 + \label{eq:Coulomb}
310 + \end{equation}
311 + then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
312 + al.}'s undamped prescription:
313 + \begin{equation}
314 + v_\textrm{SP}(r) =
315 + q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
316 + r\leqslant R_\textrm{c},
317 + \label{eq:SPPot}
318 + \end{equation}
319 + with associated forces,
320 + \begin{equation}
321 + f_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
322 + \label{eq:SPForces}
323 + \end{equation}
324 + These forces are identical to the forces of the standard Coulomb
325 + interaction, and cutting these off at $R_c$ was addressed by Wolf
326 + \textit{et al.} as undesirable.  They pointed out that the effect of
327 + the image charges is neglected in the forces when this form is
328 + used,\cite{Wolf99} thereby eliminating any benefit from the method in
329 + molecular dynamics.  Additionally, there is a discontinuity in the
330 + forces at the cutoff radius which results in energy drift during MD
331 + simulations.
332 +
333 + The shifted force ({\sc sf}) form using the normal Coulomb potential
334 + will give,
335 + \begin{equation}
336 + 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}.
337 + \label{eq:SFPot}
338 + \end{equation}
339 + with associated forces,
340 + \begin{equation}
341 + 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}.
342 + \label{eq:SFForces}
343 + \end{equation}
344 + This formulation has the benefits that there are no discontinuities at
345 + the cutoff distance, while the neutralizing image charges are present
346 + in both the energy and force expressions.  It would be simple to add
347 + the self-neutralizing term back when computing the total energy of the
348 + system, thereby maintaining the agreement with the Madelung energies.
349 + A side effect of this treatment is the alteration in the shape of the
350 + potential that comes from the derivative term.  Thus, a degree of
351 + clarity about agreement with the empirical potential is lost in order
352 + to gain functionality in dynamics simulations.
353 +
354 + Wolf \textit{et al.} originally discussed the energetics of the
355 + shifted Coulomb potential (Eq. \ref{eq:SPPot}), and they found that
356 + it was still insufficient for accurate determination of the energy
357 + with reasonable cutoff distances.  The calculated Madelung energies
358 + fluctuate around the expected value with increasing cutoff radius, but
359 + the oscillations converge toward the correct value.\cite{Wolf99} A
360 + damping function was incorporated to accelerate the convergence; and
361 + though alternative functional forms could be
362 + used,\cite{Jones56,Heyes81} the complimentary error function was
363 + chosen to mirror the effective screening used in the Ewald summation.
364 + Incorporating this error function damping into the simple Coulomb
365 + potential,
366 + \begin{equation}
367 + v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
368 + \label{eq:dampCoulomb}
369 + \end{equation}
370 + the shifted potential (Eq. (\ref{eq:SPPot})) can be reacquired using
371 + eq. (\ref{eq:shiftingForm}),
372 + \begin{equation}
373 + 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},
374 + \label{eq:DSPPot}
375 + \end{equation}
376 + with associated forces,
377 + \begin{equation}
378 + 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}.
379 + \label{eq:DSPForces}
380 + \end{equation}
381 + Again, this damped shifted potential suffers from a discontinuity and
382 + a lack of the image charges in the forces.  To remedy these concerns,
383 + one may derive a {\sc sf} variant by including  the derivative
384 + term in eq. (\ref{eq:shiftingForm}),
385 + \begin{equation}
386 + \begin{split}
387 + 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}.
388 + \label{eq:DSFPot}
389 + \end{split}
390 + \end{equation}
391 + The derivative of the above potential will lead to the following forces,
392 + \begin{equation}
393 + \begin{split}
394 + 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}.
395 + \label{eq:DSFForces}
396 + \end{split}
397 + \end{equation}
398 + If the damping parameter $(\alpha)$ is chosen to be zero, the undamped
399 + case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
400 + from eqs. (\ref{eq:DSPPot}-\ref{eq:DSFForces}).
401 +
402 + This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
403 + derived by Zahn \textit{et al.}; however, there are two important
404 + differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
405 + eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
406 + with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present
407 + in the Zahn potential, resulting in a potential discontinuity as
408 + particles cross $R_\textrm{c}$.  Second, the sign of the derivative
409 + portion is different.  The missing $v_\textrm{c}$ term would not
410 + affect molecular dynamics simulations (although the computed energy
411 + would be expected to have sudden jumps as particle distances crossed
412 + $R_c$).  The sign problem would be a potential source of errors,
413 + however.  In fact, it introduces a discontinuity in the forces at the
414 + cutoff, because the force function is shifted in the wrong direction
415 + and doesn't cross zero at $R_\textrm{c}$.
416 +
417 + Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
418 + electrostatic summation method that is continuous in both the
419 + potential and forces and which incorporates the damping function
420 + proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
421 + paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
422 + sf}, damping) are at reproducing the correct electrostatic summation
423 + performed by the Ewald sum.
424 +
425 + \subsection{Other alternatives}
426 + In addition to the methods described above, we will consider some
427 + other techniques that commonly get used in molecular simulations.  The
428 + simplest of these is group-based cutoffs.  Though of little use for
429 + non-neutral molecules, collecting atoms into neutral groups takes
430 + advantage of the observation that the electrostatic interactions decay
431 + faster than those for monopolar pairs.\cite{Steinbach94} When
432 + considering these molecules as groups, an orientational aspect is
433 + introduced to the interactions.  Consequently, as these molecular
434 + particles move through $R_\textrm{c}$, the energy will drift upward
435 + due to the anisotropy of the net molecular dipole
436 + interactions.\cite{Rahman71} To maintain good energy conservation,
437 + both the potential and derivative need to be smoothly switched to zero
438 + at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
439 + switching function,
440 + \begin{equation}
441 + S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
442 + \frac{(R_\textrm{c}+2r-3r_\textrm{sw})(R_\textrm{c}-r)^2}{(R_\textrm{c}-r_\textrm{sw})^3} &\quad r_\textrm{sw}<r\leqslant R_\textrm{c} \\
443 + 0 &\quad r>R_\textrm{c}
444 + \end{cases},
445 + \end{equation}
446 + where the above form is for a cubic function.  If a smooth second
447 + derivative is desired, a fifth (or higher) order polynomial can be
448 + used.\cite{Andrea83}
449 +
450 + Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
451 + and to incorporate their effect, a method like Reaction Field ({\sc
452 + rf}) can be used.  The original theory for {\sc rf} was originally
453 + developed by Onsager,\cite{Onsager36} and it was applied in
454 + simulations for the study of water by Barker and Watts.\cite{Barker73}
455 + In application, it is simply an extension of the group-based cutoff
456 + method where the net dipole within the cutoff sphere polarizes an
457 + external dielectric, which reacts back on the central dipole.  The
458 + same switching function considerations for group-based cutoffs need to
459 + made for {\sc rf}, with the additional pre-specification of a
460 + dielectric constant.
461 +
462 + \section{Methods}
463 +
464 + In classical molecular mechanics simulations, there are two primary
465 + techniques utilized to obtain information about the system of
466 + interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
467 + techniques utilize pairwise summations of interactions between
468 + particle sites, but they use these summations in different ways.
469 +
470 + In MC, the potential energy difference between two subsequent
471 + configurations dictates the progression of MC sampling.  Going back to
472 + the origins of this method, the acceptance criterion for the canonical
473 + ensemble laid out by Metropolis \textit{et al.} states that a
474 + subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
475 + \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
476 + 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
477 + alternate method for handling the long-range electrostatics will
478 + ensure proper sampling from the ensemble.
479 +
480 + In MD, the derivative of the potential governs how the system will
481 + progress in time.  Consequently, the force and torque vectors on each
482 + body in the system dictate how the system evolves.  If the magnitude
483 + and direction of these vectors are similar when using alternate
484 + electrostatic summation techniques, the dynamics in the short term
485 + will be indistinguishable.  Because error in MD calculations is
486 + cumulative, one should expect greater deviation at longer times,
487 + although methods which have large differences in the force and torque
488 + vectors will diverge from each other more rapidly.
489 +
490 + \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
491 + The pairwise summation techniques (outlined in section
492 + \ref{sec:ESMethods}) were evaluated for use in MC simulations by
493 + studying the energy differences between conformations.  We took the
494 + SPME-computed energy difference between two conformations to be the
495 + correct behavior. An ideal performance by an alternative method would
496 + reproduce these energy differences exactly.  Since none of the methods
497 + provide exact energy differences, we used linear least squares
498 + regressions of the $\Delta E$ values between configurations using SPME
499 + against $\Delta E$ values using tested methods provides a quantitative
500 + comparison of this agreement.  Unitary results for both the
501 + correlation and correlation coefficient for these regressions indicate
502 + equivalent energetic results between the method under consideration
503 + and electrostatics handled using SPME.  Sample correlation plots for
504 + two alternate methods are shown in Fig. \ref{fig:linearFit}.
505 +
506   \begin{figure}
507   \centering
508 < \includegraphics[width=3.25in]{./slice.pdf}
509 < \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.}
510 < \label{argonSlice}
508 > \includegraphics[width = \linewidth]{./dualLinear.pdf}
509 > \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.}
510 > \label{fig:linearFit}
511   \end{figure}
512  
513 < All of these comparisons were performed with three different cutoff radii (9, 12, and 15 \AA) to investigate the cutoff radius dependence of the various techniques.  It should be noted that the damping parameter chosen in SPME, or so called ``Ewald Coefficient", has a significant effect on the energies and forces calculated.  Typical molecular mechanics packages default this to a value dependent on the cutoff radius and a tolerance (typically less than $1 \times 10^{-5}$ kcal/mol).  We chose a tolerance of $1 \times 10^{-8}$, resulting in Ewald Coefficients of 0.4200, 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
513 > Each system type (detailed in section \ref{sec:RepSims}) was
514 > represented using 500 independent configurations.  Additionally, we
515 > used seven different system types, so each of the alternate
516 > (non-Ewald) electrostatic summation methods was evaluated using
517 > 873,250 configurational energy differences.
518  
519 < \section{Results and Discussion}
519 > Results and discussion for the individual analysis of each of the
520 > system types appear in the supporting information, while the
521 > cumulative results over all the investigated systems appears below in
522 > section \ref{sec:EnergyResults}.
523  
524 < \subsection{$\Delta E$ Comparison}
525 < In order to evaluate the performance of the adapted Wolf Shifted Potential and Shifted Force electrostatic summation methods for Monte Carlo simulations, the energy differences between configurations need to be compared to the results using SPME.  Considering the SPME results to be the correct or desired behavior, ideal performance of a tested method is taken to be agreement between the energy differences calculated.  Linear least squares regression of the $\Delta E$ values between configurations using SPME against $\Delta E$ values using tested methods provides a quantitative comparison of this agreement.  Unitary results for both the correlation and correlation coefficient for these regressions indicate equivalent energetic results between the methods.  The correlation is the slope of the plotted data while the correlation coefficient ($R^2$) is a measure of the of the data scatter around the fitted line and gives an idea of the quality of the fit (Fig. \ref{linearFit}).
524 > \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
525 > We evaluated the pairwise methods (outlined in section
526 > \ref{sec:ESMethods}) for use in MD simulations by
527 > comparing the force and torque vectors with those obtained using the
528 > reference Ewald summation (SPME).  Both the magnitude and the
529 > direction of these vectors on each of the bodies in the system were
530 > analyzed.  For the magnitude of these vectors, linear least squares
531 > regression analyses were performed as described previously for
532 > comparing $\Delta E$ values.  Instead of a single energy difference
533 > between two system configurations, we compared the magnitudes of the
534 > forces (and torques) on each molecule in each configuration.  For a
535 > system of 1000 water molecules and 40 ions, there are 1040 force
536 > vectors and 1000 torque vectors.  With 500 configurations, this
537 > results in 520,000 force and 500,000 torque vector comparisons.
538 > Additionally, data from seven different system types was aggregated
539 > before the comparison was made.
540  
541 + The {\it directionality} of the force and torque vectors was
542 + investigated through measurement of the angle ($\theta$) formed
543 + between those computed from the particular method and those from SPME,
544 + \begin{equation}
545 + \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
546 + \end{equation}
547 + where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
548 + force vector computed using method $M$.  
549 +
550 + Each of these $\theta$ values was accumulated in a distribution
551 + function, weighted by the area on the unit sphere.  Non-linear
552 + Gaussian fits were used to measure the width of the resulting
553 + distributions.
554 +
555   \begin{figure}
556   \centering
557 < \includegraphics[width=3.25in]{./linearFit.pdf}
558 < \caption{Example least squares regression of the $\Delta E$ between configurations for the SF method against SPME in the pure water system.  }
559 < \label{linearFit}
557 > \includegraphics[width = \linewidth]{./gaussFit.pdf}
558 > \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.}
559 > \label{fig:gaussian}
560   \end{figure}
561  
562 < With 500 independent configurations, 124,750 $\Delta E$ data points are used in a regression of a single system.  Results and discussion for the individual analysis of each of the system types appear in the supporting information.  To probe the applicability of each method in the general case, all the different system types were included in a single regression.  The results for this regression are shown in figure \ref{delE}.  
562 > Figure \ref{fig:gaussian} shows an example distribution with applied
563 > non-linear fits.  The solid line is a Gaussian profile, while the
564 > dotted line is a Voigt profile, a convolution of a Gaussian and a
565 > Lorentzian.  Since this distribution is a measure of angular error
566 > between two different electrostatic summation methods, there is no
567 > {\it a priori} reason for the profile to adhere to any specific shape.
568 > Gaussian fits was used to compare all the tested methods.  The
569 > variance ($\sigma^2$) was extracted from each of these fits and was
570 > used to compare distribution widths.  Values of $\sigma^2$ near zero
571 > indicate vector directions indistinguishable from those calculated
572 > when using the reference method (SPME).
573  
574 + \subsection{Short-time Dynamics}
575 +
576 + \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
577 + Evaluation of the long-time dynamics of charged systems was performed
578 + by considering the NaCl crystal system while using a subset of the
579 + best performing pairwise methods.  The NaCl crystal was chosen to
580 + avoid possible complications involving the propagation techniques of
581 + orientational motion in molecular systems.  To enhance the atomic
582 + motion, these crystals were equilibrated at 1000 K, near the
583 + experimental $T_m$ for NaCl.  Simulations were performed under the
584 + microcanonical ensemble, and velocity autocorrelation functions
585 + (Eq. \ref{eq:vCorr}) were computed for each of the trajectories,
586 + \begin{equation}
587 + C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
588 + \label{eq:vCorr}
589 + \end{equation}
590 + Velocity autocorrelation functions require detailed short time data
591 + and long trajectories for good statistics, thus velocity information
592 + was saved every 5 fs over 100 ps trajectories.  The power spectrum
593 + ($I(\omega)$) is obtained via Fourier transform of the autocorrelation
594 + function
595 + \begin{equation}
596 + I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
597 + \label{eq:powerSpec}
598 + \end{equation}
599 + where the frequency, $\omega=0,\ 1,\ ...,\ N-1$.
600 +
601 + \subsection{Representative Simulations}\label{sec:RepSims}
602 + A variety of common and representative simulations were analyzed to
603 + determine the relative effectiveness of the pairwise summation
604 + techniques in reproducing the energetics and dynamics exhibited by
605 + SPME.  The studied systems were as follows:
606 + \begin{enumerate}
607 + \item Liquid Water
608 + \item Crystalline Water (Ice I$_\textrm{c}$)
609 + \item NaCl Crystal
610 + \item NaCl Melt
611 + \item Low Ionic Strength Solution of NaCl in Water
612 + \item High Ionic Strength Solution of NaCl in Water
613 + \item 6 \AA\  Radius Sphere of Argon in Water
614 + \end{enumerate}
615 + By utilizing the pairwise techniques (outlined in section
616 + \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
617 + charged particles, and mixtures of the two, we can comment on possible
618 + system dependence and/or universal applicability of the techniques.
619 +
620 + Generation of the system configurations was dependent on the system
621 + type.  For the solid and liquid water configurations, configuration
622 + snapshots were taken at regular intervals from higher temperature 1000
623 + SPC/E water molecule trajectories and each equilibrated individually.
624 + The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
625 + ions and were selected and equilibrated in the same fashion as the
626 + water systems.  For the low and high ionic strength NaCl solutions, 4
627 + and 40 ions were first solvated in a 1000 water molecule boxes
628 + respectively.  Ion and water positions were then randomly swapped, and
629 + the resulting configurations were again equilibrated individually.
630 + Finally, for the Argon/Water "charge void" systems, the identities of
631 + all the SPC/E waters within 6 \AA\ of the center of the equilibrated
632 + water configurations were converted to argon
633 + (Fig. \ref{fig:argonSlice}).
634 +
635   \begin{figure}
636   \centering
637 < \includegraphics[width=3.25in]{./delEplot.pdf}
638 < \caption{The results from the statistical analysis of the $\Delta$E results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii.  Results close to a value of 1 (dashed line) indicate $\Delta E$ values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME.  Reaction Field results do not include NaCl crystal or melt configurations.}
639 < \label{delE}
637 > \includegraphics[width = \linewidth]{./slice.pdf}
638 > \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.}
639 > \label{fig:argonSlice}
640   \end{figure}
641  
642 < In figure \ref{delE}, it is apparent that it is unreasonable to expect realistic results using an unmodified cutoff.  This is not all that surprising since this results in large energy fluctuations as atoms move in and out of the cutoff radius.  These fluctuations can be alleviated to some degree by using group based cutoffs with a switching function.  The Group Switch Cutoff row doesn't show a significant improvement in this plot because the salt and salt solution systems contain non-neutral groups, see the accompanying supporting information for a comparison where all groups are neutral.  Correcting the resulting charged cutoff sphere is one of the purposes of the shifted potential proposed by Wolf \textit{et al.}, and this correction indeed improves the results as seen in the Shifted Potental rows.  While the undamped case of this method is a significant improvement over the pure cutoff, it still doesn't correlate that well with SPME.  Inclusion of potential damping improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows an excellent correlation and quality of fit with the SPME results, particularly with a cutoff radius greater than 12 \AA .  Use of a larger damping parameter is more helpful for the shortest cutoff shown, but it has a detrimental effect on simulations with larger cutoffs.  This trend is repeated in the Shifted Force rows, where increasing damping results in progressively poorer correlation; however, damping looks to be unnecessary with this method.  Overall, the undamped case is the best performing set, as the correlation and quality of fits are consistently superior regardless of the cutoff distance.  This result is beneficial in that the undamped case is less computationally prohibitive do to the lack of complimentary error function calculation when performing the electrostatic pair interaction.  The reaction field results illustrates some of that method's limitations, primarily that it was developed for use in homogenous systems; although it does provide results that are an improvement over those from an unmodified cutoff.
642 > \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
643 > Electrostatic summation method comparisons were performed using SPME,
644 > the {\sc sp} and {\sc sf} methods - both with damping
645 > parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
646 > moderate, and strong damping respectively), reaction field with an
647 > infinite dielectric constant, and an unmodified cutoff.  Group-based
648 > cutoffs with a fifth-order polynomial switching function were
649 > necessary for the reaction field simulations and were utilized in the
650 > SP, SF, and pure cutoff methods for comparison to the standard lack of
651 > group-based cutoffs with a hard truncation.  The SPME calculations
652 > were performed using the TINKER implementation of SPME,\cite{Ponder87}
653 > while all other method calculations were performed using the OOPSE
654 > molecular mechanics package.\cite{Meineke05}
655  
656 < \subsection{Force Magnitude Comparison}
656 > These methods were additionally evaluated with three different cutoff
657 > radii (9, 12, and 15 \AA) to investigate possible cutoff radius
658 > dependence.  It should be noted that the damping parameter chosen in
659 > SPME, or so called ``Ewald Coefficient", has a significant effect on
660 > the energies and forces calculated.  Typical molecular mechanics
661 > packages default this to a value dependent on the cutoff radius and a
662 > tolerance (typically less than $1 \times 10^{-4}$ kcal/mol).  Smaller
663 > tolerances are typically associated with increased accuracy, but this
664 > usually means more time spent calculating the reciprocal-space portion
665 > of the summation.\cite{Perram88,Essmann95} The default TINKER
666 > tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
667 > calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
668 > 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
669  
670 < While studying the energy differences provides insight into how comparable these methods are energetically, if we want to use these methods in Molecular Dynamics simulations, we also need to consider their effect on forces and torques.  Both the magnitude and the direction of the force and torque vectors of each of the bodies in the system can be compared to those observed while using SPME.  Analysis of the magnitude of these vectors can be performed in the manner described previously for comparing $\Delta E$ values, only instead of a single value between two system configurations, there is a value for each particle in each configuration.  For a system of 1000 water molecules and 40 ions, there are 1040 force vectors and 1000 torque vectors.  With 500 configurations, this results in excess of 500,000 data samples for each system type.  Figures \ref{frcMag} and \ref{trqMag} respectively show the force and torque vector magnitude results for the accumulated analysis over all the system types.
670 > \section{Results and Discussion}
671  
672 + \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
673 + In order to evaluate the performance of the pairwise electrostatic
674 + summation methods for Monte Carlo simulations, the energy differences
675 + between configurations were compared to the values obtained when using
676 + SPME.  The results for the subsequent regression analysis are shown in
677 + figure \ref{fig:delE}.
678 +
679   \begin{figure}
680   \centering
681 < \includegraphics[width=3.25in]{./frcMagplot.pdf}
682 < \caption{The results from the statistical analysis of the force vector magnitude results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii.  Results close to a value of 1 (dashed line) indicate force vector magnitude values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME.}
683 < \label{frcMag}
681 > \includegraphics[width=5.5in]{./delEplot.pdf}
682 > \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).}
683 > \label{fig:delE}
684   \end{figure}
685  
686 < The results in figure \ref{frcMag} for the most part parallel those seen in the previous look at the $\Delta E$ results.  The unmodified cutoff results are poor, but using group based cutoffs and a switching function provides a improvement much more significant than what was seen with $\Delta E$.  Looking at the Shifted Potential sets, the slope and R$^2$ improve with the use of damping to an optimal result of 0.2 \AA $^{-1}$ for the 12 and 15 \AA\ cutoffs.  Further increases in damping, while beneficial for simulations with a cutoff radius of 9 \AA\ , is detrimental to simulations with larger cutoff radii.  The undamped Shifted Force method gives forces in line with those obtained using SPME, and use of a damping function gives little to no gain.  The reaction field results are surprisingly good, considering the poor quality of the fits for the $\Delta E$ results.  There is still a considerable degree of scatter in the data, but it correlates well in general.
686 > In this figure, it is apparent that it is unreasonable to expect
687 > realistic results using an unmodified cutoff.  This is not all that
688 > surprising since this results in large energy fluctuations as atoms
689 > move in and out of the cutoff radius.  These fluctuations can be
690 > alleviated to some degree by using group based cutoffs with a
691 > switching function.\cite{Steinbach94} The Group Switch Cutoff row
692 > doesn't show a significant improvement in this plot because the salt
693 > and salt solution systems contain non-neutral groups, see the
694 > accompanying supporting information for a comparison where all groups
695 > are neutral.
696  
697 < \subsection{Torque Magnitude Comparison}
697 > Correcting the resulting charged cutoff sphere is one of the purposes
698 > of the damped Coulomb summation proposed by Wolf \textit{et
699 > al.},\cite{Wolf99} and this correction indeed improves the results as
700 > seen in the Shifted-Potental rows.  While the undamped case of this
701 > method is a significant improvement over the pure cutoff, it still
702 > doesn't correlate that well with SPME.  Inclusion of potential damping
703 > improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
704 > an excellent correlation and quality of fit with the SPME results,
705 > particularly with a cutoff radius greater than 12 \AA .  Use of a
706 > larger damping parameter is more helpful for the shortest cutoff
707 > shown, but it has a detrimental effect on simulations with larger
708 > cutoffs.  In the {\sc sf} sets, increasing damping results in
709 > progressively poorer correlation.  Overall, the undamped case is the
710 > best performing set, as the correlation and quality of fits are
711 > consistently superior regardless of the cutoff distance.  This result
712 > is beneficial in that the undamped case is less computationally
713 > prohibitive do to the lack of complimentary error function calculation
714 > when performing the electrostatic pair interaction.  The reaction
715 > field results illustrates some of that method's limitations, primarily
716 > that it was developed for use in homogenous systems; although it does
717 > provide results that are an improvement over those from an unmodified
718 > cutoff.
719  
720 + \subsection{Magnitudes of the Force and Torque Vectors}
721 +
722 + Evaluation of pairwise methods for use in Molecular Dynamics
723 + simulations requires consideration of effects on the forces and
724 + torques.  Investigation of the force and torque vector magnitudes
725 + provides a measure of the strength of these values relative to SPME.
726 + Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
727 + force and torque vector magnitude regression results for the
728 + accumulated analysis over all the system types.
729 +
730   \begin{figure}
731   \centering
732 < \includegraphics[width=3.25in]{./trqMagplot.pdf}
733 < \caption{The results from the statistical analysis of the torque vector magnitude results for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii.  Results close to a value of 1 (dashed line) indicate torque vector magnitude values from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME.  Torques are only accumulated on the rigid water molecules, so these results exclude NaCl the systems.}
734 < \label{trqMag}
732 > \includegraphics[width=5.5in]{./frcMagplot.pdf}
733 > \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).}
734 > \label{fig:frcMag}
735   \end{figure}
736  
737 < The torque vector magnitude results in figure \ref{trqMag} are similar to those seen for the forces, but more clearly show the improved behavior with increasing cutoff radius.  Moderate damping is beneficial to the Shifted Potential and unnecessary with the Shifted Force method, and they also show that over-damping adversely effects all cutoff radii rather than showing an improvement for systems with short cutoffs.  The reaction field method performs well when calculating the torques, better than the Shifted Force method over this limited data set.
737 > Figure \ref{fig:frcMag}, for the most part, parallels the results seen
738 > in the previous $\Delta E$ section.  The unmodified cutoff results are
739 > poor, but using group based cutoffs and a switching function provides
740 > a improvement much more significant than what was seen with $\Delta
741 > E$.  Looking at the {\sc sp} sets, the slope and $R^2$
742 > improve with the use of damping to an optimal result of 0.2 \AA
743 > $^{-1}$ for the 12 and 15 \AA\ cutoffs.  Further increases in damping,
744 > while beneficial for simulations with a cutoff radius of 9 \AA\ , is
745 > detrimental to simulations with larger cutoff radii.  The undamped
746 > {\sc sf} method gives forces in line with those obtained using
747 > SPME, and use of a damping function results in minor improvement.  The
748 > reaction field results are surprisingly good, considering the poor
749 > quality of the fits for the $\Delta E$ results.  There is still a
750 > considerable degree of scatter in the data, but it correlates well in
751 > general.  To be fair, we again note that the reaction field
752 > calculations do not encompass NaCl crystal and melt systems, so these
753 > results are partly biased towards conditions in which the method
754 > performs more favorably.
755  
124 \subsection{Force and Torque Direction Comparison}
125
126 Having force and torque vectors with magnitudes that are well correlated to SPME is good, but if they are not pointing in the proper direction the results will be incorrect.  These vector directions were investigated through measurement of the angle formed between them and those from SPME.  The dot product of these unit vectors provides a theta value that is accumulated in a distribution function, weighted by the area on the unit sphere.  Narrow distributions of theta values indicates similar to identical results between the tested method and SPME.  To measure the narrowness of the resulting distributions, non-linear Gaussian fits were performed.
127
756   \begin{figure}
757   \centering
758 < \includegraphics[width=3.25in]{./gaussFit.pdf}
759 < \caption{Example fitting of the angular distribution of the force vectors over all of the studied systems.  The solid and dotted lines show Gaussian and Voigt fits of the distribution data respectively.  Even though the Voigt profile make for a more accurate fit, the Gaussian was used due to more versatile statistical results.}
760 < \label{gaussian}
758 > \includegraphics[width=5.5in]{./trqMagplot.pdf}
759 > \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).}
760 > \label{fig:trqMag}
761   \end{figure}
762  
763 < Figure \ref{gaussian} shows an example distribution and the non-linear fit applied.  The solid line is a Gaussian profile, while the dotted line is a Voigt profile, a convolution of a Gaussian and a Lorentzian profile.  Since this distribution is a measure of angular error between two different electrostatic summation methods, there is particular reason for it to adhere to a particular shape.  Because of this and the Gaussian profile's more statistically meaningful properties, Gaussian fitting was used to compare all the methods considered in this study.  The results (Fig. \ref{frcTrqAng}) are compared through the variance ($\sigma^2$) of these non-linear fits.  
763 > To evaluate the torque vector magnitudes, the data set from which
764 > values are drawn is limited to rigid molecules in the systems
765 > (i.e. water molecules).  In spite of this smaller sampling pool, the
766 > torque vector magnitude results in figure \ref{fig:trqMag} are still
767 > similar to those seen for the forces; however, they more clearly show
768 > the improved behavior that comes with increasing the cutoff radius.
769 > Moderate damping is beneficial to the {\sc sp} and helpful
770 > yet possibly unnecessary with the {\sc sf} method, and they also
771 > show that over-damping adversely effects all cutoff radii rather than
772 > showing an improvement for systems with short cutoffs.  The reaction
773 > field method performs well when calculating the torques, better than
774 > the Shifted Force method over this limited data set.
775  
776 + \subsection{Directionality of the Force and Torque Vectors}
777 +
778 + Having force and torque vectors with magnitudes that are well
779 + correlated to SPME is good, but if they are not pointing in the proper
780 + direction the results will be incorrect.  These vector directions were
781 + investigated through measurement of the angle formed between them and
782 + those from SPME.  The results (Fig. \ref{fig:frcTrqAng}) are compared
783 + through the variance ($\sigma^2$) of the Gaussian fits of the angle
784 + error distributions of the combined set over all system types.
785 +
786   \begin{figure}
787   \centering
788 < \includegraphics[width=3.25in]{./frcTrqAngplot.pdf}
789 < \caption{The results from the statistical analysis of the force and torque vector angular distributions for all the system types at 9 \AA\ (${\bullet}$), 12 \AA\ ($\blacksquare$), and 15 \AA\ ($\blacktriangledown$) cutoff radii.  Plotted values are the variance ($\sigma^2$) of the Gaussian non-linear fits.  Results close to a value of 0 (dashed line) indicate force or torque vector directions from that particular method (listed on the left) are nearly indistinguishable from those obtained from SPME.  Torques are only accumulated on the rigid water molecules, so the torque vector angle results exclude NaCl the systems.}
790 < \label{frcTrqAng}
788 > \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
789 > \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).}
790 > \label{fig:frcTrqAng}
791   \end{figure}
792  
793 < Both the force and torque $\sigma^2$ results from the analysis of the total accumulated system data are tabulated in figure \ref{frcTrqAng}.  All of the sets, aside from the over-damped case show the improvement afforded by choosing a longer simulation cutoff.  Increasing the cutoff from 9 to 12 \AA\ typically results in a halving of $\sigma^2$, with a similar improvement going from 12 to 15 \AA .  The undamped Shifted Force, Group Based Cutoff, and Reaction Field methods all do equivalently well at capturing the direction of both the force and torque vectors.  Using damping improves the angular behavior significantly for the Shifted Potential and moderately for the Shifted Force methods.  Increasing the damping too far is destructive for both methods, particularly to the torque vectors.  Again it is important to recognize that the force vectors cover all particles in the systems, while torque vectors are only available for neutral molecular groups.  Damping appears to have a more beneficial non-neutral bodies, and this observation is investigated further in the accompanying supporting information.  
793 > Both the force and torque $\sigma^2$ results from the analysis of the
794 > total accumulated system data are tabulated in figure
795 > \ref{fig:frcTrqAng}.  All of the sets, aside from the over-damped case
796 > show the improvement afforded by choosing a longer simulation cutoff.
797 > Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
798 > of the distribution widths, with a similar improvement going from 12
799 > to 15 \AA .  The undamped {\sc sf}, Group Based Cutoff, and
800 > Reaction Field methods all do equivalently well at capturing the
801 > direction of both the force and torque vectors.  Using damping
802 > improves the angular behavior significantly for the {\sc sp}
803 > and moderately for the {\sc sf} methods.  Increasing the damping
804 > too far is destructive for both methods, particularly to the torque
805 > vectors.  Again it is important to recognize that the force vectors
806 > cover all particles in the systems, while torque vectors are only
807 > available for neutral molecular groups.  Damping appears to have a
808 > more beneficial effect on non-neutral bodies, and this observation is
809 > investigated further in the accompanying supporting information.
810  
811   \begin{table}[htbp]
812     \centering
# Line 173 | Line 838 | Both the force and torque $\sigma^2$ results from the
838  
839        \bottomrule
840     \end{tabular}
841 <   \label{groupAngle}
841 >   \label{tab:groupAngle}
842   \end{table}
843  
844 < Although not discussed previously, group based cutoffs can be applied to both the Shifted Potential and Force methods.  Use off a switching function corrects for the discontinuities that arise when atoms of a group exit the cutoff before the group's center of mass.  Though there are no significant benefit or drawbacks observed in $\Delta E$ and vector magnitude results when doing this, there is a measurable improvement in the vector angle results.  Table \ref{groupAngle} shows the angular variance values obtained using group based cutoffs and a switching function alongside the standard results seen in figure \ref{frcTrqAng} for comparison purposes.  The Shifted Potential shows much narrower angular distributions for both the force and torque vectors when using an $\alpha$ of 0.2 \AA$^{-1}$ or less, while Shifted Force shows improvements in the undamped and lightly damped cases.  Thus, by calculating the electrostatic interactions in terms of molecular pairs rather than atomic pairs, the direction of the force and torque vectors are determined more accurately.  
844 > Although not discussed previously, group based cutoffs can be applied
845 > to both the {\sc sp} and {\sc sf} methods.  Use off a
846 > switching function corrects for the discontinuities that arise when
847 > atoms of a group exit the cutoff before the group's center of mass.
848 > Though there are no significant benefit or drawbacks observed in
849 > $\Delta E$ and vector magnitude results when doing this, there is a
850 > measurable improvement in the vector angle results.  Table
851 > \ref{tab:groupAngle} shows the angular variance values obtained using
852 > group based cutoffs and a switching function alongside the standard
853 > results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
854 > The {\sc sp} shows much narrower angular distributions for
855 > both the force and torque vectors when using an $\alpha$ of 0.2
856 > \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
857 > undamped and lightly damped cases.  Thus, by calculating the
858 > electrostatic interactions in terms of molecular pairs rather than
859 > atomic pairs, the direction of the force and torque vectors are
860 > determined more accurately.
861  
862 < One additional trend to recognize in table \ref{groupAngle} is that the $\sigma^2$ values for both Shifted Potential and Shifted Force converge as $\alpha$ increases, something that is easier to see when using group based cutoffs.  Looking back on figures \ref{delE}, \ref{frcMag}, and \ref{trqMag}, show this behavior clearly at large $\alpha$ and cutoff values.  The reason for this is that the complimentary error function inserted into the potential weakens the electrostatic interaction as $\alpha$ increases.  Thus, at larger values of $\alpha$, both the summation method types progress toward non-interacting functions, so care is required in choosing large damping functions lest one generate an undesirable loss in the pair interaction.  Kast \textit{et al.}  developed a method for choosing appropriate $\alpha$ values for these types of electrostatic summation methods by fitting to $g(r)$ data, and their methods indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\ respectively.\cite{Kast03}  These appear to be reasonable choices to obtain proper MC behavior (Fig. \ref{delE}); however, based on these findings, choices this high would be introducing error in the molecular torques, particularly for the shorter cutoffs.  Based on the above findings, any empirical damping is arguably unnecessary with the choice of the Shifted Force method.
862 > One additional trend to recognize in table \ref{tab:groupAngle} is
863 > that the $\sigma^2$ values for both {\sc sp} and
864 > {\sc sf} converge as $\alpha$ increases, something that is easier
865 > to see when using group based cutoffs.  Looking back on figures
866 > \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
867 > behavior clearly at large $\alpha$ and cutoff values.  The reason for
868 > this is that the complimentary error function inserted into the
869 > potential weakens the electrostatic interaction as $\alpha$ increases.
870 > Thus, at larger values of $\alpha$, both the summation method types
871 > progress toward non-interacting functions, so care is required in
872 > choosing large damping functions lest one generate an undesirable loss
873 > in the pair interaction.  Kast \textit{et al.}  developed a method for
874 > choosing appropriate $\alpha$ values for these types of electrostatic
875 > summation methods by fitting to $g(r)$ data, and their methods
876 > indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
877 > values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
878 > to be reasonable choices to obtain proper MC behavior
879 > (Fig. \ref{fig:delE}); however, based on these findings, choices this
880 > high would introduce error in the molecular torques, particularly for
881 > the shorter cutoffs.  Based on the above findings, empirical damping
882 > up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
883 > unnecessary when using the {\sc sf} method.
884  
885 + \subsection{Collective Motion: Power Spectra of NaCl Crystals}
886 +
887 + In the previous studies using a {\sc sf} variant of the damped
888 + Wolf coulomb potential, the structure and dynamics of water were
889 + investigated rather extensively.\cite{Zahn02,Kast03} Their results
890 + indicated that the damped {\sc sf} method results in properties
891 + very similar to those obtained when using the Ewald summation.
892 + Considering the statistical results shown above, the good performance
893 + of this method is not that surprising.  Rather than consider the same
894 + systems and simply recapitulate their results, we decided to look at
895 + the solid state dynamical behavior obtained using the best performing
896 + summation methods from the above results.
897 +
898 + \begin{figure}
899 + \centering
900 + \includegraphics[width = \linewidth]{./spectraSquare.pdf}
901 + \caption{Power spectra obtained from the velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, {\sc sf} ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 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.}
902 + \label{fig:methodPS}
903 + \end{figure}
904 +
905 + Figure \ref{fig:methodPS} shows the power spectra for the NaCl
906 + crystals (from averaged Na and Cl ion velocity autocorrelation
907 + functions) using the stated electrostatic summation methods.  While
908 + high frequency peaks of all the spectra overlap, showing the same
909 + general features, the low frequency region shows how the summation
910 + methods differ.  Considering the low-frequency inset (expanded in the
911 + upper frame of figure \ref{fig:dampInc}), at frequencies below 100
912 + cm$^{-1}$, the correlated motions are blue-shifted when using undamped
913 + or weakly damped {\sc sf}.  When using moderate damping ($\alpha
914 + = 0.2$ \AA$^{-1}$) both the {\sc sf} and {\sc sp}
915 + methods give near identical correlated motion behavior as the Ewald
916 + method (which has a damping value of 0.3119).  The damping acts as a
917 + distance dependent Gaussian screening of the point charges for the
918 + pairwise summation methods.  This weakening of the electrostatic
919 + interaction with distance explains why the long-ranged correlated
920 + motions are at lower frequencies for the moderately damped methods
921 + than for undamped or weakly damped methods.  To see this effect more
922 + clearly, we show how damping strength affects a simple real-space
923 + electrostatic potential,
924 + \begin{equation}
925 + V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
926 + \end{equation}
927 + where $S(r)$ is a switching function that smoothly zeroes the
928 + potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how
929 + the low frequency motions are dependent on the damping used in the
930 + direct electrostatic sum.  As the damping increases, the peaks drop to
931 + lower frequencies.  Incidentally, use of an $\alpha$ of 0.25
932 + \AA$^{-1}$ on a simple electrostatic summation results in low
933 + frequency correlated dynamics equivalent to a simulation using SPME.
934 + When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
935 + shift to higher frequency in exponential fashion.  Though not shown,
936 + the spectrum for the simple undamped electrostatic potential is
937 + blue-shifted such that the lowest frequency peak resides near 325
938 + cm$^{-1}$.  In light of these results, the undamped {\sc sf}
939 + method producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is
940 + quite respectable; however, it appears as though moderate damping is
941 + required for accurate reproduction of crystal dynamics.
942 + \begin{figure}
943 + \centering
944 + \includegraphics[width = \linewidth]{./comboSquare.pdf}
945 + \caption{Regions of spectra showing the low-frequency correlated motions for NaCl crystals at 1000 K using various electrostatic summation methods.  The upper plot is a zoomed inset from figure \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential increases, the low-frequency peaks red-shift.  The lower plot is of spectra 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.}
946 + \label{fig:dampInc}
947 + \end{figure}
948 +
949   \section{Conclusions}
950  
951 < \section{Acknowledgments}
951 > This investigation of pairwise electrostatic summation techniques
952 > shows that there are viable and more computationally efficient
953 > electrostatic summation techniques than the Ewald summation, chiefly
954 > methods derived from the damped Coulombic sum originally proposed by
955 > Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
956 > {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
957 > shows a remarkable ability to reproduce the energetic and dynamic
958 > characteristics exhibited by simulations employing lattice summation
959 > techniques.  The cumulative energy difference results showed the
960 > undamped {\sc sf} and moderately damped {\sc sp} methods
961 > produced results nearly identical to SPME.  Similarly for the dynamic
962 > features, the undamped or moderately damped {\sc sf} and
963 > moderately damped {\sc sp} methods produce force and torque
964 > vector magnitude and directions very similar to the expected values.
965 > These results translate into long-time dynamic behavior equivalent to
966 > that produced in simulations using SPME.
967  
968 + Aside from the computational cost benefit, these techniques have
969 + applicability in situations where the use of the Ewald sum can prove
970 + problematic.  Primary among them is their use in interfacial systems,
971 + where the unmodified lattice sum techniques artificially accentuate
972 + the periodicity of the system in an undesirable manner.  There have
973 + been alterations to the standard Ewald techniques, via corrections and
974 + reformulations, to compensate for these systems; but the pairwise
975 + techniques discussed here require no modifications, making them
976 + natural tools to tackle these problems.  Additionally, this
977 + transferability gives them benefits over other pairwise methods, like
978 + reaction field, because estimations of physical properties (e.g. the
979 + dielectric constant) are unnecessary.
980 +
981 + We are not suggesting any flaw with the Ewald sum; in fact, it is the
982 + standard by which these simple pairwise sums are judged.  However,
983 + these results do suggest that in the typical simulations performed
984 + today, the Ewald summation may no longer be required to obtain the
985 + level of accuracy most researcher have come to expect
986 +
987 + \section{Acknowledgments}
988   \newpage
989  
990 < \bibliographystyle{achemso}
990 > \bibliographystyle{jcp2}
991   \bibliography{electrostaticMethods}
992  
993  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines