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 2629 by chrisfen, Thu Mar 16 03:48:32 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}
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 > blah blah blah Ewald Sum Important blah blah blah
85  
64 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.  
65
66 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}).
67
86   \begin{figure}
87   \centering
88 < \includegraphics[width=3.25in]{./slice.pdf}
89 < \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.}
90 < \label{argonSlice}
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 < 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.
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 < \section{Results and Discussion}
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 < \subsection{$\Delta E$ Comparison}
136 < 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}).
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 > using eq. (\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 {\sc sf} variant by including  the derivative
296 > term in eq. (\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 {\sc sf} 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 > In addition to the methods described above, we will consider some
337 > other techniques that commonly get used in molecular simulations.  The
338 > simplest of these is group-based cutoffs.  Though of little use for
339 > non-neutral molecules, collecting atoms into neutral groups takes
340 > advantage of the observation that the electrostatic interactions decay
341 > faster than those for monopolar pairs.\cite{Steinbach94} When
342 > considering these molecules as groups, an orientational aspect is
343 > introduced to the interactions.  Consequently, as these molecular
344 > particles move through $R_\textrm{c}$, the energy will drift upward
345 > due to the anisotropy of the net molecular dipole
346 > interactions.\cite{Rahman71} To maintain good energy conservation,
347 > both the potential and derivative need to be smoothly switched to zero
348 > at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
349 > switching function,
350 > \begin{equation}
351 > S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
352 > \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} \\
353 > 0 &\quad r>R_\textrm{c}
354 > \end{cases},
355 > \end{equation}
356 > where the above form is for a cubic function.  If a smooth second
357 > derivative is desired, a fifth (or higher) order polynomial can be
358 > used.\cite{Andrea83}
359 >
360 > Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
361 > and to incorporate their effect, a method like Reaction Field ({\sc
362 > rf}) can be used.  The orignal theory for {\sc rf} was originally
363 > developed by Onsager,\cite{Onsager36} and it was applied in
364 > simulations for the study of water by Barker and Watts.\cite{Barker73}
365 > In application, it is simply an extension of the group-based cutoff
366 > method where the net dipole within the cutoff sphere polarizes an
367 > external dielectric, which reacts back on the central dipole.  The
368 > same switching function considerations for group-based cutoffs need to
369 > made for {\sc rf}, with the additional prespecification of a
370 > dielectric constant.
371 >
372 > \section{Methods}
373 >
374 > In classical molecular mechanics simulations, there are two primary
375 > techniques utilized to obtain information about the system of
376 > interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
377 > techniques utilize pairwise summations of interactions between
378 > particle sites, but they use these summations in different ways.
379 >
380 > In MC, the potential energy difference between two subsequent
381 > configurations dictates the progression of MC sampling.  Going back to
382 > the origins of this method, the acceptance criterion for the canonical
383 > ensemble laid out by Metropolis \textit{et al.} states that a
384 > subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
385 > \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
386 > 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
387 > alternate method for handling the long-range electrostatics will
388 > ensure proper sampling from the ensemble.
389 >
390 > In MD, the derivative of the potential governs how the system will
391 > progress in time.  Consequently, the force and torque vectors on each
392 > body in the system dictate how the system evolves.  If the magnitude
393 > and direction of these vectors are similar when using alternate
394 > electrostatic summation techniques, the dynamics in the short term
395 > will be indistinguishable.  Because error in MD calculations is
396 > cumulative, one should expect greater deviation at longer times,
397 > although methods which have large differences in the force and torque
398 > vectors will diverge from each other more rapidly.
399 >
400 > \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
401 > The pairwise summation techniques (outlined in section
402 > \ref{sec:ESMethods}) were evaluated for use in MC simulations by
403 > studying the energy differences between conformations.  We took the
404 > SPME-computed energy difference between two conformations to be the
405 > correct behavior. An ideal performance by an alternative method would
406 > reproduce these energy differences exactly.  Since none of the methods
407 > provide exact energy differences, we used linear least squares
408 > regressions of the $\Delta E$ values between configurations using SPME
409 > against $\Delta E$ values using tested methods provides a quantitative
410 > comparison of this agreement.  Unitary results for both the
411 > correlation and correlation coefficient for these regressions indicate
412 > equivalent energetic results between the method under consideration
413 > and electrostatics handled using SPME.  Sample correlation plots for
414 > two alternate methods are shown in Fig. \ref{fig:linearFit}.
415 >
416 > \begin{figure}
417 > \centering
418 > \includegraphics[width = \linewidth]{./dualLinear.pdf}
419 > \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.}
420 > \label{fig:linearFit}
421 > \end{figure}
422 >
423 > Each system type (detailed in section \ref{sec:RepSims}) was
424 > represented using 500 independent configurations.  Additionally, we
425 > used seven different system types, so each of the alternate
426 > (non-Ewald) electrostatic summation methods was evaluated using
427 > 873,250 configurational energy differences.
428 >
429 > Results and discussion for the individual analysis of each of the
430 > system types appear in the supporting information, while the
431 > cumulative results over all the investigated systems appears below in
432 > section \ref{sec:EnergyResults}.
433 >
434 > \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
435 > We evaluated the pairwise methods (outlined in section
436 > \ref{sec:ESMethods}) for use in MD simulations by
437 > comparing the force and torque vectors with those obtained using the
438 > reference Ewald summation (SPME).  Both the magnitude and the
439 > direction of these vectors on each of the bodies in the system were
440 > analyzed.  For the magnitude of these vectors, linear least squares
441 > regression analyses were performed as described previously for
442 > comparing $\Delta E$ values.  Instead of a single energy difference
443 > between two system configurations, we compared the magnitudes of the
444 > forces (and torques) on each molecule in each configuration.  For a
445 > system of 1000 water molecules and 40 ions, there are 1040 force
446 > vectors and 1000 torque vectors.  With 500 configurations, this
447 > results in 520,000 force and 500,000 torque vector comparisons.
448 > Additionally, data from seven different system types was aggregated
449 > before the comparison was made.
450  
451 + The {\it directionality} of the force and torque vectors was
452 + investigated through measurement of the angle ($\theta$) formed
453 + between those computed from the particular method and those from SPME,
454 + \begin{equation}
455 + \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
456 + \end{equation}
457 + where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
458 + force vector computed using method $M$.  
459 +
460 + Each of these $\theta$ values was accumulated in a distribution
461 + function, weighted by the area on the unit sphere.  Non-linear
462 + Gaussian fits were used to measure the width of the resulting
463 + distributions.
464 +
465   \begin{figure}
466   \centering
467 < \includegraphics[width=3.25in]{./linearFit.pdf}
468 < \caption{Example least squares regression of the $\Delta E$ between configurations for the SF method against SPME in the pure water system.  }
469 < \label{linearFit}
467 > \includegraphics[width = \linewidth]{./gaussFit.pdf}
468 > \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.}
469 > \label{fig:gaussian}
470   \end{figure}
471  
472 < 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}.  
472 > Figure \ref{fig:gaussian} shows an example distribution with applied
473 > non-linear fits.  The solid line is a Gaussian profile, while the
474 > dotted line is a Voigt profile, a convolution of a Gaussian and a
475 > Lorentzian.  Since this distribution is a measure of angular error
476 > between two different electrostatic summation methods, there is no
477 > {\it a priori} reason for the profile to adhere to any specific shape.
478 > Gaussian fits was used to compare all the tested methods.  The
479 > variance ($\sigma^2$) was extracted from each of these fits and was
480 > used to compare distribution widths.  Values of $\sigma^2$ near zero
481 > indicate vector directions indistinguishable from those calculated
482 > when using the reference method (SPME).
483  
484 + \subsection{Short-time Dynamics}
485 +
486 + \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
487 + Evaluation of the long-time dynamics of charged systems was performed
488 + by considering the NaCl crystal system while using a subset of the
489 + best performing pairwise methods.  The NaCl crystal was chosen to
490 + avoid possible complications involving the propagation techniques of
491 + orientational motion in molecular systems.  To enhance the atomic
492 + motion, these crystals were equilibrated at 1000 K, near the
493 + experimental $T_m$ for NaCl.  Simulations were performed under the
494 + microcanonical ensemble, and velocity autocorrelation functions
495 + (Eq. \ref{eq:vCorr}) were computed for each of the trajectories,
496 + \begin{equation}
497 + C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
498 + \label{eq:vCorr}
499 + \end{equation}
500 + Velocity autocorrelation functions require detailed short time data
501 + and long trajectories for good statistics, thus velocity information
502 + was saved every 5 fs over 100 ps trajectories.  The power spectrum
503 + ($I(\omega)$) is obtained via Fourier transform of the autocorrelation
504 + function
505 + \begin{equation}
506 + I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
507 + \label{eq:powerSpec}
508 + \end{equation}
509 + where the frequency, $\omega=0,\ 1,\ ...,\ N-1$.
510 +
511 + \subsection{Representative Simulations}\label{sec:RepSims}
512 + A variety of common and representative simulations were analyzed to
513 + determine the relative effectiveness of the pairwise summation
514 + techniques in reproducing the energetics and dynamics exhibited by
515 + SPME.  The studied systems were as follows:
516 + \begin{enumerate}
517 + \item Liquid Water
518 + \item Crystalline Water (Ice I$_\textrm{c}$)
519 + \item NaCl Crystal
520 + \item NaCl Melt
521 + \item Low Ionic Strength Solution of NaCl in Water
522 + \item High Ionic Strength Solution of NaCl in Water
523 + \item 6 \AA\  Radius Sphere of Argon in Water
524 + \end{enumerate}
525 + By utilizing the pairwise techniques (outlined in section
526 + \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
527 + charged particles, and mixtures of the two, we can comment on possible
528 + system dependence and/or universal applicability of the techniques.
529 +
530 + Generation of the system configurations was dependent on the system
531 + type.  For the solid and liquid water configurations, configuration
532 + snapshots were taken at regular intervals from higher temperature 1000
533 + SPC/E water molecule trajectories and each equilibrated individually.
534 + The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
535 + ions and were selected and equilibrated in the same fashion as the
536 + water systems.  For the low and high ionic strength NaCl solutions, 4
537 + and 40 ions were first solvated in a 1000 water molecule boxes
538 + respectively.  Ion and water positions were then randomly swapped, and
539 + the resulting configurations were again equilibrated individually.
540 + Finally, for the Argon/Water "charge void" systems, the identities of
541 + all the SPC/E waters within 6 \AA\ of the center of the equilibrated
542 + water configurations were converted to argon
543 + (Fig. \ref{fig:argonSlice}).
544 +
545   \begin{figure}
546   \centering
547 < \includegraphics[width=3.25in]{./delEplot.pdf}
548 < \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.}
549 < \label{delE}
547 > \includegraphics[width = \linewidth]{./slice.pdf}
548 > \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.}
549 > \label{fig:argonSlice}
550   \end{figure}
551  
552 < 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.
552 > \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
553 > Electrostatic summation method comparisons were performed using SPME,
554 > the {\sc sp} and {\sc sf} methods - both with damping
555 > parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
556 > moderate, and strong damping respectively), reaction field with an
557 > infinite dielectric constant, and an unmodified cutoff.  Group-based
558 > cutoffs with a fifth-order polynomial switching function were
559 > necessary for the reaction field simulations and were utilized in the
560 > SP, SF, and pure cutoff methods for comparison to the standard lack of
561 > group-based cutoffs with a hard truncation.  The SPME calculations
562 > were performed using the TINKER implementation of SPME,\cite{Ponder87}
563 > while all other method calculations were performed using the OOPSE
564 > molecular mechanics package.\cite{Meineke05}
565  
566 < \subsection{Force Magnitude Comparison}
566 > These methods were additionally evaluated with three different cutoff
567 > radii (9, 12, and 15 \AA) to investigate possible cutoff radius
568 > dependence.  It should be noted that the damping parameter chosen in
569 > SPME, or so called ``Ewald Coefficient", has a significant effect on
570 > the energies and forces calculated.  Typical molecular mechanics
571 > packages default this to a value dependent on the cutoff radius and a
572 > tolerance (typically less than $1 \times 10^{-4}$ kcal/mol).  Smaller
573 > tolerances are typically associated with increased accuracy in the
574 > real-space portion of the summation.\cite{Essmann95} The default
575 > TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
576 > calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
577 > 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
578  
579 < 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.
579 > \section{Results and Discussion}
580  
581 + \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
582 + In order to evaluate the performance of the pairwise electrostatic
583 + summation methods for Monte Carlo simulations, the energy differences
584 + between configurations were compared to the values obtained when using
585 + SPME.  The results for the subsequent regression analysis are shown in
586 + figure \ref{fig:delE}.
587 +
588   \begin{figure}
589   \centering
590 < \includegraphics[width=3.25in]{./frcMagplot.pdf}
591 < \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.}
592 < \label{frcMag}
590 > \includegraphics[width=5.5in]{./delEplot.pdf}
591 > \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).}
592 > \label{fig:delE}
593   \end{figure}
594  
595 < 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.
595 > In this figure, it is apparent that it is unreasonable to expect
596 > realistic results using an unmodified cutoff.  This is not all that
597 > surprising since this results in large energy fluctuations as atoms
598 > move in and out of the cutoff radius.  These fluctuations can be
599 > alleviated to some degree by using group based cutoffs with a
600 > switching function.\cite{Steinbach94} The Group Switch Cutoff row
601 > doesn't show a significant improvement in this plot because the salt
602 > and salt solution systems contain non-neutral groups, see the
603 > accompanying supporting information for a comparison where all groups
604 > are neutral.
605  
606 < \subsection{Torque Magnitude Comparison}
606 > Correcting the resulting charged cutoff sphere is one of the purposes
607 > of the damped Coulomb summation proposed by Wolf \textit{et
608 > al.},\cite{Wolf99} and this correction indeed improves the results as
609 > seen in the Shifted-Potental rows.  While the undamped case of this
610 > method is a significant improvement over the pure cutoff, it still
611 > doesn't correlate that well with SPME.  Inclusion of potential damping
612 > improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
613 > an excellent correlation and quality of fit with the SPME results,
614 > particularly with a cutoff radius greater than 12 \AA .  Use of a
615 > larger damping parameter is more helpful for the shortest cutoff
616 > shown, but it has a detrimental effect on simulations with larger
617 > cutoffs.  In the {\sc sf} sets, increasing damping results in
618 > progressively poorer correlation.  Overall, the undamped case is the
619 > best performing set, as the correlation and quality of fits are
620 > consistently superior regardless of the cutoff distance.  This result
621 > is beneficial in that the undamped case is less computationally
622 > prohibitive do to the lack of complimentary error function calculation
623 > when performing the electrostatic pair interaction.  The reaction
624 > field results illustrates some of that method's limitations, primarily
625 > that it was developed for use in homogenous systems; although it does
626 > provide results that are an improvement over those from an unmodified
627 > cutoff.
628  
629 + \subsection{Magnitudes of the Force and Torque Vectors}
630 +
631 + Evaluation of pairwise methods for use in Molecular Dynamics
632 + simulations requires consideration of effects on the forces and
633 + torques.  Investigation of the force and torque vector magnitudes
634 + provides a measure of the strength of these values relative to SPME.
635 + Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
636 + force and torque vector magnitude regression results for the
637 + accumulated analysis over all the system types.
638 +
639   \begin{figure}
640   \centering
641 < \includegraphics[width=3.25in]{./trqMagplot.pdf}
642 < \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.}
643 < \label{trqMag}
641 > \includegraphics[width=5.5in]{./frcMagplot.pdf}
642 > \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).}
643 > \label{fig:frcMag}
644   \end{figure}
645  
646 < 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.
647 <
648 < \subsection{Force and Torque Direction Comparison}
649 <
650 < 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.
646 > Figure \ref{fig:frcMag}, for the most part, parallels the results seen
647 > in the previous $\Delta E$ section.  The unmodified cutoff results are
648 > poor, but using group based cutoffs and a switching function provides
649 > a improvement much more significant than what was seen with $\Delta
650 > E$.  Looking at the {\sc sp} sets, the slope and $R^2$
651 > improve with the use of damping to an optimal result of 0.2 \AA
652 > $^{-1}$ for the 12 and 15 \AA\ cutoffs.  Further increases in damping,
653 > while beneficial for simulations with a cutoff radius of 9 \AA\ , is
654 > detrimental to simulations with larger cutoff radii.  The undamped
655 > {\sc sf} method gives forces in line with those obtained using
656 > SPME, and use of a damping function results in minor improvement.  The
657 > reaction field results are surprisingly good, considering the poor
658 > quality of the fits for the $\Delta E$ results.  There is still a
659 > considerable degree of scatter in the data, but it correlates well in
660 > general.  To be fair, we again note that the reaction field
661 > calculations do not encompass NaCl crystal and melt systems, so these
662 > results are partly biased towards conditions in which the method
663 > performs more favorably.
664  
665   \begin{figure}
666   \centering
667 < \includegraphics[width=3.25in]{./gaussFit.pdf}
668 < \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.}
669 < \label{gaussian}
667 > \includegraphics[width=5.5in]{./trqMagplot.pdf}
668 > \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).}
669 > \label{fig:trqMag}
670   \end{figure}
671  
672 < 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.  
672 > To evaluate the torque vector magnitudes, the data set from which
673 > values are drawn is limited to rigid molecules in the systems
674 > (i.e. water molecules).  In spite of this smaller sampling pool, the
675 > torque vector magnitude results in figure \ref{fig:trqMag} are still
676 > similar to those seen for the forces; however, they more clearly show
677 > the improved behavior that comes with increasing the cutoff radius.
678 > Moderate damping is beneficial to the {\sc sp} and helpful
679 > yet possibly unnecessary with the {\sc sf} method, and they also
680 > show that over-damping adversely effects all cutoff radii rather than
681 > showing an improvement for systems with short cutoffs.  The reaction
682 > field method performs well when calculating the torques, better than
683 > the Shifted Force method over this limited data set.
684  
685 + \subsection{Directionality of the Force and Torque Vectors}
686 +
687 + Having force and torque vectors with magnitudes that are well
688 + correlated to SPME is good, but if they are not pointing in the proper
689 + direction the results will be incorrect.  These vector directions were
690 + investigated through measurement of the angle formed between them and
691 + those from SPME.  The results (Fig. \ref{fig:frcTrqAng}) are compared
692 + through the variance ($\sigma^2$) of the Gaussian fits of the angle
693 + error distributions of the combined set over all system types.
694 +
695   \begin{figure}
696   \centering
697 < \includegraphics[width=3.25in]{./frcTrqAngplot.pdf}
698 < \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.}
699 < \label{frcTrqAng}
697 > \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
698 > \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).}
699 > \label{fig:frcTrqAng}
700   \end{figure}
701  
702 < 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.  
702 > Both the force and torque $\sigma^2$ results from the analysis of the
703 > total accumulated system data are tabulated in figure
704 > \ref{fig:frcTrqAng}.  All of the sets, aside from the over-damped case
705 > show the improvement afforded by choosing a longer simulation cutoff.
706 > Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
707 > of the distribution widths, with a similar improvement going from 12
708 > to 15 \AA .  The undamped {\sc sf}, Group Based Cutoff, and
709 > Reaction Field methods all do equivalently well at capturing the
710 > direction of both the force and torque vectors.  Using damping
711 > improves the angular behavior significantly for the {\sc sp}
712 > and moderately for the {\sc sf} methods.  Increasing the damping
713 > too far is destructive for both methods, particularly to the torque
714 > vectors.  Again it is important to recognize that the force vectors
715 > cover all particles in the systems, while torque vectors are only
716 > available for neutral molecular groups.  Damping appears to have a
717 > more beneficial effect on non-neutral bodies, and this observation is
718 > investigated further in the accompanying supporting information.
719  
720   \begin{table}[htbp]
721     \centering
# Line 173 | Line 747 | Both the force and torque $\sigma^2$ results from the
747  
748        \bottomrule
749     \end{tabular}
750 <   \label{groupAngle}
750 >   \label{tab:groupAngle}
751   \end{table}
752  
753 < 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.  
753 > Although not discussed previously, group based cutoffs can be applied
754 > to both the {\sc sp} and {\sc sf} methods.  Use off a
755 > switching function corrects for the discontinuities that arise when
756 > atoms of a group exit the cutoff before the group's center of mass.
757 > Though there are no significant benefit or drawbacks observed in
758 > $\Delta E$ and vector magnitude results when doing this, there is a
759 > measurable improvement in the vector angle results.  Table
760 > \ref{tab:groupAngle} shows the angular variance values obtained using
761 > group based cutoffs and a switching function alongside the standard
762 > results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
763 > The {\sc sp} shows much narrower angular distributions for
764 > both the force and torque vectors when using an $\alpha$ of 0.2
765 > \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
766 > undamped and lightly damped cases.  Thus, by calculating the
767 > electrostatic interactions in terms of molecular pairs rather than
768 > atomic pairs, the direction of the force and torque vectors are
769 > determined more accurately.
770  
771 < 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.
771 > One additional trend to recognize in table \ref{tab:groupAngle} is
772 > that the $\sigma^2$ values for both {\sc sp} and
773 > {\sc sf} converge as $\alpha$ increases, something that is easier
774 > to see when using group based cutoffs.  Looking back on figures
775 > \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
776 > behavior clearly at large $\alpha$ and cutoff values.  The reason for
777 > this is that the complimentary error function inserted into the
778 > potential weakens the electrostatic interaction as $\alpha$ increases.
779 > Thus, at larger values of $\alpha$, both the summation method types
780 > progress toward non-interacting functions, so care is required in
781 > choosing large damping functions lest one generate an undesirable loss
782 > in the pair interaction.  Kast \textit{et al.}  developed a method for
783 > choosing appropriate $\alpha$ values for these types of electrostatic
784 > summation methods by fitting to $g(r)$ data, and their methods
785 > indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
786 > values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
787 > to be reasonable choices to obtain proper MC behavior
788 > (Fig. \ref{fig:delE}); however, based on these findings, choices this
789 > high would introduce error in the molecular torques, particularly for
790 > the shorter cutoffs.  Based on the above findings, empirical damping
791 > up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
792 > unnecessary when using the {\sc sf} method.
793  
794 + \subsection{Collective Motion: Power Spectra of NaCl Crystals}
795 +
796 + In the previous studies using a {\sc sf} variant of the damped
797 + Wolf coulomb potential, the structure and dynamics of water were
798 + investigated rather extensively.\cite{Zahn02,Kast03} Their results
799 + indicated that the damped {\sc sf} method results in properties
800 + very similar to those obtained when using the Ewald summation.
801 + Considering the statistical results shown above, the good performance
802 + of this method is not that surprising.  Rather than consider the same
803 + systems and simply recapitulate their results, we decided to look at
804 + the solid state dynamical behavior obtained using the best performing
805 + summation methods from the above results.
806 +
807 + \begin{figure}
808 + \centering
809 + \includegraphics[width = \linewidth]{./spectraSquare.pdf}
810 + \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.}
811 + \label{fig:methodPS}
812 + \end{figure}
813 +
814 + Figure \ref{fig:methodPS} shows the power spectra for the NaCl
815 + crystals (from averaged Na and Cl ion velocity autocorrelation
816 + functions) using the stated electrostatic summation methods.  While
817 + high frequency peaks of all the spectra overlap, showing the same
818 + general features, the low frequency region shows how the summation
819 + methods differ.  Considering the low-frequency inset (expanded in the
820 + upper frame of figure \ref{fig:dampInc}), at frequencies below 100
821 + cm$^{-1}$, the correlated motions are blue-shifted when using undamped
822 + or weakly damped {\sc sf}.  When using moderate damping ($\alpha
823 + = 0.2$ \AA$^{-1}$) both the {\sc sf} and {\sc sp}
824 + methods give near identical correlated motion behavior as the Ewald
825 + method (which has a damping value of 0.3119).  The damping acts as a
826 + distance dependent Gaussian screening of the point charges for the
827 + pairwise summation methods.  This weakening of the electrostatic
828 + interaction with distance explains why the long-ranged correlated
829 + motions are at lower frequencies for the moderately damped methods
830 + than for undamped or weakly damped methods.  To see this effect more
831 + clearly, we show how damping strength affects a simple real-space
832 + electrostatic potential,
833 + \begin{equation}
834 + V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
835 + \end{equation}
836 + where $S(r)$ is a switching function that smoothly zeroes the
837 + potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how
838 + the low frequency motions are dependent on the damping used in the
839 + direct electrostatic sum.  As the damping increases, the peaks drop to
840 + lower frequencies.  Incidentally, use of an $\alpha$ of 0.25
841 + \AA$^{-1}$ on a simple electrostatic summation results in low
842 + frequency correlated dynamics equivalent to a simulation using SPME.
843 + When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
844 + shift to higher frequency in exponential fashion.  Though not shown,
845 + the spectrum for the simple undamped electrostatic potential is
846 + blue-shifted such that the lowest frequency peak resides near 325
847 + cm$^{-1}$.  In light of these results, the undamped {\sc sf}
848 + method producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is
849 + quite respectable; however, it appears as though moderate damping is
850 + required for accurate reproduction of crystal dynamics.
851 + \begin{figure}
852 + \centering
853 + \includegraphics[width = \linewidth]{./comboSquare.pdf}
854 + \caption{Upper: Zoomed inset from figure \ref{fig:methodPS}.  As the damping value for the {\sc sf} 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.}
855 + \label{fig:dampInc}
856 + \end{figure}
857 +
858   \section{Conclusions}
859  
860 < \section{Acknowledgments}
860 > This investigation of pairwise electrostatic summation techniques
861 > shows that there are viable and more computationally efficient
862 > electrostatic summation techniques than the Ewald summation, chiefly
863 > methods derived from the damped Coulombic sum originally proposed by
864 > Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
865 > {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
866 > shows a remarkable ability to reproduce the energetic and dynamic
867 > characteristics exhibited by simulations employing lattice summation
868 > techniques.  The cumulative energy difference results showed the
869 > undamped {\sc sf} and moderately damped {\sc sp} methods
870 > produced results nearly identical to SPME.  Similarly for the dynamic
871 > features, the undamped or moderately damped {\sc sf} and
872 > moderately damped {\sc sp} methods produce force and torque
873 > vector magnitude and directions very similar to the expected values.
874 > These results translate into long-time dynamic behavior equivalent to
875 > that produced in simulations using SPME.
876  
877 + Aside from the computational cost benefit, these techniques have
878 + applicability in situations where the use of the Ewald sum can prove
879 + problematic.  Primary among them is their use in interfacial systems,
880 + where the unmodified lattice sum techniques artificially accentuate
881 + the periodicity of the system in an undesirable manner.  There have
882 + been alterations to the standard Ewald techniques, via corrections and
883 + reformulations, to compensate for these systems; but the pairwise
884 + techniques discussed here require no modifications, making them
885 + natural tools to tackle these problems.  Additionally, this
886 + transferability gives them benefits over other pairwise methods, like
887 + reaction field, because estimations of physical properties (e.g. the
888 + dielectric constant) are unnecessary.
889 +
890 + We are not suggesting any flaw with the Ewald sum; in fact, it is the
891 + standard by which these simple pairwise sums are judged.  However,
892 + these results do suggest that in the typical simulations performed
893 + today, the Ewald summation may no longer be required to obtain the
894 + level of accuracy most researcher have come to expect
895 +
896 + \section{Acknowledgments}
897   \newpage
898  
899 < \bibliographystyle{achemso}
899 > \bibliographystyle{jcp2}
900   \bibliography{electrostaticMethods}
901  
902  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines