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 2652 by chrisfen, Tue Mar 21 19:26:59 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 essential and is one of the most
69 > computationally-demanding tasks.  The common molecular mechanics force
70 > fields represent atomic sites with full or partial charges protected
71 > by Lennard-Jones (short range) interactions.  This means that nearly
72 > every pair interaction involves a calculation of charge-charge forces.
73 > Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
74 > interactions quickly become the most expensive part of molecular
75 > simulations.  Historically, the electrostatic pair interaction would
76 > not have decayed appreciably within the typical box lengths that could
77 > be feasibly simulated.  In the larger systems that are more typical of
78 > modern simulations, large cutoffs should be used to incorporate
79 > electrostatics correctly.
80  
81 < \section{Methods}
81 > There have been many efforts to address the proper and practical
82 > handling of electrostatic interactions, and these have resulted in a
83 > variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
84 > typically classified as implicit methods (i.e., continuum dielectrics,
85 > static dipolar fields),\cite{Born20,Grossfield00} explicit methods
86 > (i.e., Ewald summations, interaction shifting or
87 > truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
88 > reaction field type methods, fast multipole
89 > methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
90 > often preferred because they physically incorporate solvent molecules
91 > in the system of interest, but these methods are sometimes difficult
92 > to utilize because of their high computational cost.\cite{Roux99} In
93 > addition to the computational cost, there have been some questions
94 > regarding possible artifacts caused by the inherent periodicity of the
95 > explicit Ewald summation.\cite{Tobias01}
96  
97 < 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.  
97 > In this paper, we focus on a new set of shifted methods devised by
98 > Wolf {\it et al.},\cite{Wolf99} which we further extend.  These
99 > methods along with a few other mixed methods (i.e. reaction field) are
100 > compared with the smooth particle mesh Ewald
101 > sum,\cite{Onsager36,Essmann99} which is our reference method for
102 > handling long-range electrostatic interactions. The new methods for
103 > handling electrostatics have the potential to scale linearly with
104 > increasing system size since they involve only a simple modification
105 > to the direct pairwise sum.  They also lack the added periodicity of
106 > the Ewald sum, so they can be used for systems which are non-periodic
107 > or which have one- or two-dimensional periodicity.  Below, these
108 > methods are evaluated using a variety of model systems to establish
109 > their usability in molecular simulations.
110  
111 < 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}).
111 > \subsection{The Ewald Sum}
112 > The complete accumulation electrostatic interactions in a system with
113 > periodic boundary conditions (PBC) requires the consideration of the
114 > effect of all charges within a (cubic) simulation box as well as those
115 > in the periodic replicas,
116 > \begin{equation}
117 > 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],
118 > \label{eq:PBCSum}
119 > \end{equation}
120 > where the sum over $\mathbf{n}$ is a sum over all periodic box
121 > replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
122 > prime indicates $i = j$ are neglected for $\mathbf{n} =
123 > 0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
124 > particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
125 > the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
126 > $j$, and $\phi$ is the solution to Poisson's equation
127 > ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
128 > charge-charge interactions). In the case of monopole electrostatics,
129 > eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
130 > non-neutral systems.
131  
132 + The electrostatic summation problem was originally studied by Ewald
133 + for the case of an infinite crystal.\cite{Ewald21}. The approach he
134 + took was to convert this conditionally convergent sum into two
135 + absolutely convergent summations: a short-ranged real-space summation
136 + and a long-ranged reciprocal-space summation,
137 + \begin{equation}
138 + \begin{split}
139 + 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,
140 + \end{split}
141 + \label{eq:EwaldSum}
142 + \end{equation}
143 + where $\alpha$ is the damping or convergence parameter with units of
144 + \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
145 + $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
146 + constant of the surrounding medium. The final two terms of
147 + eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
148 + for interacting with a surrounding dielectric.\cite{Allen87} This
149 + dipolar term was neglected in early applications in molecular
150 + simulations,\cite{Brush66,Woodcock71} until it was introduced by de
151 + Leeuw {\it et al.} to address situations where the unit cell has a
152 + dipole moment which is magnified through replication of the periodic
153 + images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
154 + system is said to be using conducting (or ``tin-foil'') boundary
155 + conditions, $\epsilon_{\rm S} = \infty$. Figure
156 + \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
157 + time.  Initially, due to the small sizes of the systems that could be
158 + feasibly simulated, the entire simulation box was replicated to
159 + convergence.  In more modern simulations, the simulation boxes have
160 + grown large enough that a real-space cutoff could potentially give
161 + convergent behavior.  Indeed, it has often been observed that the
162 + reciprocal-space portion of the Ewald sum can be small and rapidly
163 + convergent compared to the real-space portion with the choice of small
164 + $\alpha$.\cite{Karasawa89,Kolafa92}
165 +
166   \begin{figure}
167   \centering
168 < \includegraphics[width=3.25in]{./slice.pdf}
169 < \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.}
170 < \label{argonSlice}
168 > \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
169 > \caption{How the application of the Ewald summation has changed with
170 > the increase in computer power.  Initially, only small numbers of
171 > particles could be studied, and the Ewald sum acted to replicate the
172 > unit cell charge distribution out to convergence.  Now, much larger
173 > systems of charges are investigated with fixed distance cutoffs.  The
174 > calculated structure factor is used to sum out to great distance, and
175 > a surrounding dielectric term is included.}
176 > \label{fig:ewaldTime}
177   \end{figure}
178  
179 < 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.
179 > The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The
180 > convergence parameter $(\alpha)$ plays an important role in balancing
181 > the computational cost between the direct and reciprocal-space
182 > portions of the summation.  The choice of this value allows one to
183 > select whether the real-space or reciprocal space portion of the
184 > summation is an $\mathscr{O}(N^2)$ calculation (with the other being
185 > $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
186 > $\alpha$ and thoughtful algorithm development, this cost can be
187 > reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
188 > taken to reduce the cost of the Ewald summation even further is to set
189 > $\alpha$ such that the real-space interactions decay rapidly, allowing
190 > for a short spherical cutoff. Then the reciprocal space summation is
191 > optimized.  These optimizations usually involve utilization of the
192 > fast Fourier transform (FFT),\cite{Hockney81} leading to the
193 > particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
194 > methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
195 > methods, the cost of the reciprocal-space portion of the Ewald
196 > summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
197 > \log N)$.
198  
199 < \section{Results and Discussion}
199 > These developments and optimizations have made the use of the Ewald
200 > summation routine in simulations with periodic boundary
201 > conditions. However, in certain systems, such as vapor-liquid
202 > interfaces and membranes, the intrinsic three-dimensional periodicity
203 > can prove problematic.  The Ewald sum has been reformulated to handle
204 > 2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
205 > new methods are computationally expensive.\cite{Spohr97,Yeh99}
206 > Inclusion of a correction term in the Ewald summation is a possible
207 > direction for handling 2D systems while still enabling the use of the
208 > modern optimizations.\cite{Yeh99}
209  
210 < \subsection{$\Delta E$ Comparison}
211 < 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}).
210 > Several studies have recognized that the inherent periodicity in the
211 > Ewald sum can also have an effect on three-dimensional
212 > systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
213 > Solvated proteins are essentially kept at high concentration due to
214 > the periodicity of the electrostatic summation method.  In these
215 > systems, the more compact folded states of a protein can be
216 > artificially stabilized by the periodic replicas introduced by the
217 > Ewald summation.\cite{Weber00} Thus, care must be taken when
218 > considering the use of the Ewald summation where the assumed
219 > periodicity would introduce spurious effects in the system dynamics.
220  
221 < \begin{figure}
222 < \centering
223 < \includegraphics[width=3.25in]{./linearFit.pdf}
224 < \caption{Example least squares regression of the $\Delta E$ between configurations for the SF method against SPME in the pure water system.  }
225 < \label{linearFit}
226 < \end{figure}
221 > \subsection{The Wolf and Zahn Methods}
222 > In a recent paper by Wolf \textit{et al.}, a procedure was outlined
223 > for the accurate accumulation of electrostatic interactions in an
224 > efficient pairwise fashion.  This procedure lacks the inherent
225 > periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
226 > observed that the electrostatic interaction is effectively
227 > short-ranged in condensed phase systems and that neutralization of the
228 > charge contained within the cutoff radius is crucial for potential
229 > stability. They devised a pairwise summation method that ensures
230 > charge neutrality and gives results similar to those obtained with the
231 > Ewald summation.  The resulting shifted Coulomb potential
232 > (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
233 > placement on the cutoff sphere and a distance-dependent damping
234 > function (identical to that seen in the real-space portion of the
235 > Ewald sum) to aid convergence
236 > \begin{equation}
237 > V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
238 > \label{eq:WolfPot}
239 > \end{equation}
240 > Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
241 > potential.  However, neutralizing the charge contained within each
242 > cutoff sphere requires the placement of a self-image charge on the
243 > surface of the cutoff sphere.  This additional self-term in the total
244 > potential enabled Wolf {\it et al.}  to obtain excellent estimates of
245 > Madelung energies for many crystals.
246  
247 < 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}.  
247 > In order to use their charge-neutralized potential in molecular
248 > dynamics simulations, Wolf \textit{et al.} suggested taking the
249 > derivative of this potential prior to evaluation of the limit.  This
250 > procedure gives an expression for the forces,
251 > \begin{equation}
252 > F_{\textrm{Wolf}}(r_{ij}) = q_i q_j\left\{\left[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}\right]-\left[\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right]\right\},
253 > \label{eq:WolfForces}
254 > \end{equation}
255 > that incorporates both image charges and damping of the electrostatic
256 > interaction.
257 >
258 > More recently, Zahn \textit{et al.} investigated these potential and
259 > force expressions for use in simulations involving water.\cite{Zahn02}
260 > In their work, they pointed out that the forces and derivative of
261 > the potential are not commensurate.  Attempts to use both
262 > eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
263 > to poor energy conservation.  They correctly observed that taking the
264 > limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
265 > derivatives gives forces for a different potential energy function
266 > than the one shown in eq. (\ref{eq:WolfPot}).
267 >
268 > Zahn \textit{et al.} introduced a modified form of this summation
269 > method as a way to use the technique in Molecular Dynamics
270 > simulations.  They proposed a new damped Coulomb potential,
271 > \begin{equation}
272 > V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\},
273 > \label{eq:ZahnPot}
274 > \end{equation}
275 > and showed that this potential does fairly well at capturing the
276 > structural and dynamic properties of water compared the same
277 > properties obtained using the Ewald sum.
278 >
279 > \subsection{Simple Forms for Pairwise Electrostatics}
280 >
281 > The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
282 > al.} are constructed using two different (and separable) computational
283 > tricks: \begin{enumerate}
284 > \item shifting through the use of image charges, and
285 > \item damping the electrostatic interaction.
286 > \end{enumerate}  Wolf \textit{et al.} treated the
287 > development of their summation method as a progressive application of
288 > these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
289 > their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
290 > post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
291 > both techniques.  It is possible, however, to separate these
292 > tricks and study their effects independently.
293  
294 + Starting with the original observation that the effective range of the
295 + electrostatic interaction in condensed phases is considerably less
296 + than $r^{-1}$, either the cutoff sphere neutralization or the
297 + distance-dependent damping technique could be used as a foundation for
298 + a new pairwise summation method.  Wolf \textit{et al.} made the
299 + observation that charge neutralization within the cutoff sphere plays
300 + a significant role in energy convergence; therefore we will begin our
301 + analysis with the various shifted forms that maintain this charge
302 + neutralization.  We can evaluate the methods of Wolf
303 + \textit{et al.}  and Zahn \textit{et al.} by considering the standard
304 + shifted potential,
305 + \begin{equation}
306 + V_\textrm{SP}(r) =      \begin{cases}
307 + v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
308 + R_\textrm{c}  
309 + \end{cases},
310 + \label{eq:shiftingPotForm}
311 + \end{equation}
312 + and shifted force,
313 + \begin{equation}
314 + V_\textrm{SF}(r) =      \begin{cases}
315 + v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
316 + &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
317 +                                                \end{cases},
318 + \label{eq:shiftingForm}
319 + \end{equation}
320 + functions where $v(r)$ is the unshifted form of the potential, and
321 + $v_c$ is $v(R_\textrm{c})$.  The Shifted Force ({\sc sf}) form ensures
322 + that both the potential and the forces goes to zero at the cutoff
323 + radius, while the Shifted Potential ({\sc sp}) form only ensures the
324 + potential is smooth at the cutoff radius
325 + ($R_\textrm{c}$).\cite{Allen87}
326 +
327 + The forces associated with the shifted potential are simply the forces
328 + of the unshifted potential itself (when inside the cutoff sphere),
329 + \begin{equation}
330 + F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
331 + \end{equation}
332 + and are zero outside.  Inside the cutoff sphere, the forces associated
333 + with the shifted force form can be written,
334 + \begin{equation}
335 + F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
336 + v(r)}{dr} \right)_{r=R_\textrm{c}}.
337 + \end{equation}
338 +
339 + If the potential, $v(r)$, is taken to be the normal Coulomb potential,
340 + \begin{equation}
341 + v(r) = \frac{q_i q_j}{r},
342 + \label{eq:Coulomb}
343 + \end{equation}
344 + then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
345 + al.}'s undamped prescription:
346 + \begin{equation}
347 + V_\textrm{SP}(r) =
348 + q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
349 + r\leqslant R_\textrm{c},
350 + \label{eq:SPPot}
351 + \end{equation}
352 + with associated forces,
353 + \begin{equation}
354 + F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
355 + \label{eq:SPForces}
356 + \end{equation}
357 + These forces are identical to the forces of the standard Coulomb
358 + interaction, and cutting these off at $R_c$ was addressed by Wolf
359 + \textit{et al.} as undesirable.  They pointed out that the effect of
360 + the image charges is neglected in the forces when this form is
361 + used,\cite{Wolf99} thereby eliminating any benefit from the method in
362 + molecular dynamics.  Additionally, there is a discontinuity in the
363 + forces at the cutoff radius which results in energy drift during MD
364 + simulations.
365 +
366 + The shifted force ({\sc sf}) form using the normal Coulomb potential
367 + will give,
368 + \begin{equation}
369 + V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
370 + \label{eq:SFPot}
371 + \end{equation}
372 + with associated forces,
373 + \begin{equation}
374 + F_\textrm{SF}(r) =  q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
375 + \label{eq:SFForces}
376 + \end{equation}
377 + This formulation has the benefits that there are no discontinuities at
378 + the cutoff radius, while the neutralizing image charges are present in
379 + both the energy and force expressions.  It would be simple to add the
380 + self-neutralizing term back when computing the total energy of the
381 + system, thereby maintaining the agreement with the Madelung energies.
382 + A side effect of this treatment is the alteration in the shape of the
383 + potential that comes from the derivative term.  Thus, a degree of
384 + clarity about agreement with the empirical potential is lost in order
385 + to gain functionality in dynamics simulations.
386 +
387 + Wolf \textit{et al.} originally discussed the energetics of the
388 + shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
389 + insufficient for accurate determination of the energy with reasonable
390 + cutoff distances.  The calculated Madelung energies fluctuated around
391 + the expected value as the cutoff radius was increased, but the
392 + oscillations converged toward the correct value.\cite{Wolf99} A
393 + damping function was incorporated to accelerate the convergence; and
394 + though alternative forms for the damping function could be
395 + used,\cite{Jones56,Heyes81} the complimentary error function was
396 + chosen to mirror the effective screening used in the Ewald summation.
397 + Incorporating this error function damping into the simple Coulomb
398 + potential,
399 + \begin{equation}
400 + v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
401 + \label{eq:dampCoulomb}
402 + \end{equation}
403 + the shifted potential (eq. (\ref{eq:SPPot})) becomes
404 + \begin{equation}
405 + V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
406 + \label{eq:DSPPot}
407 + \end{equation}
408 + with associated forces,
409 + \begin{equation}
410 + F_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
411 + \label{eq:DSPForces}
412 + \end{equation}
413 + Again, this damped shifted potential suffers from a
414 + force-discontinuity at the cutoff radius, and the image charges play
415 + no role in the forces.  To remedy these concerns, one may derive a
416 + {\sc sf} variant by including the derivative term in
417 + eq. (\ref{eq:shiftingForm}),
418 + \begin{equation}
419 + \begin{split}
420 + V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}.
421 + \label{eq:DSFPot}
422 + \end{split}
423 + \end{equation}
424 + The derivative of the above potential will lead to the following forces,
425 + \begin{equation}
426 + \begin{split}
427 + F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
428 + \label{eq:DSFForces}
429 + \end{split}
430 + \end{equation}
431 + If the damping parameter $(\alpha)$ is set to zero, the undamped case,
432 + eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
433 + recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
434 +
435 + This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
436 + derived by Zahn \textit{et al.}; however, there are two important
437 + differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
438 + eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
439 + with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present
440 + in the Zahn potential, resulting in a potential discontinuity as
441 + particles cross $R_\textrm{c}$.  Second, the sign of the derivative
442 + portion is different.  The missing $v_\textrm{c}$ term would not
443 + affect molecular dynamics simulations (although the computed energy
444 + would be expected to have sudden jumps as particle distances crossed
445 + $R_c$).  The sign problem is a potential source of errors, however.
446 + In fact, it introduces a discontinuity in the forces at the cutoff,
447 + because the force function is shifted in the wrong direction and
448 + doesn't cross zero at $R_\textrm{c}$.
449 +
450 + Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
451 + electrostatic summation method in which the potential and forces are
452 + continuous at the cutoff radius and which incorporates the damping
453 + function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
454 + this paper, we will evaluate exactly how good these methods ({\sc sp},
455 + {\sc sf}, damping) are at reproducing the correct electrostatic
456 + summation performed by the Ewald sum.
457 +
458 + \subsection{Other alternatives}
459 + In addition to the methods described above, we considered some other
460 + techniques that are commonly used in molecular simulations.  The
461 + simplest of these is group-based cutoffs.  Though of little use for
462 + charged molecules, collecting atoms into neutral groups takes
463 + advantage of the observation that the electrostatic interactions decay
464 + faster than those for monopolar pairs.\cite{Steinbach94} When
465 + considering these molecules as neutral groups, the relative
466 + orientations of the molecules control the strength of the interactions
467 + at the cutoff radius.  Consequently, as these molecular particles move
468 + through $R_\textrm{c}$, the energy will drift upward due to the
469 + anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
470 + maintain good energy conservation, both the potential and derivative
471 + need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
472 + This is accomplished using a standard switching function.  If a smooth
473 + second derivative is desired, a fifth (or higher) order polynomial can
474 + be used.\cite{Andrea83}
475 +
476 + Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
477 + and to incorporate the effects of the surroundings, a method like
478 + Reaction Field ({\sc rf}) can be used.  The original theory for {\sc
479 + rf} was originally developed by Onsager,\cite{Onsager36} and it was
480 + applied in simulations for the study of water by Barker and
481 + Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
482 + an extension of the group-based cutoff method where the net dipole
483 + within the cutoff sphere polarizes an external dielectric, which
484 + reacts back on the central dipole.  The same switching function
485 + considerations for group-based cutoffs need to made for {\sc rf}, with
486 + the additional pre-specification of a dielectric constant.
487 +
488 + \section{Methods}
489 +
490 + In classical molecular mechanics simulations, there are two primary
491 + techniques utilized to obtain information about the system of
492 + interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
493 + techniques utilize pairwise summations of interactions between
494 + particle sites, but they use these summations in different ways.
495 +
496 + In MC, the potential energy difference between configurations dictates
497 + the progression of MC sampling.  Going back to the origins of this
498 + method, the acceptance criterion for the canonical ensemble laid out
499 + by Metropolis \textit{et al.} states that a subsequent configuration
500 + is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
501 + $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
502 + Maintaining the correct $\Delta E$ when using an alternate method for
503 + handling the long-range electrostatics will ensure proper sampling
504 + from the ensemble.
505 +
506 + In MD, the derivative of the potential governs how the system will
507 + progress in time.  Consequently, the force and torque vectors on each
508 + body in the system dictate how the system evolves.  If the magnitude
509 + and direction of these vectors are similar when using alternate
510 + electrostatic summation techniques, the dynamics in the short term
511 + will be indistinguishable.  Because error in MD calculations is
512 + cumulative, one should expect greater deviation at longer times,
513 + although methods which have large differences in the force and torque
514 + vectors will diverge from each other more rapidly.
515 +
516 + \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
517 +
518 + The pairwise summation techniques (outlined in section
519 + \ref{sec:ESMethods}) were evaluated for use in MC simulations by
520 + studying the energy differences between conformations.  We took the
521 + SPME-computed energy difference between two conformations to be the
522 + correct behavior. An ideal performance by an alternative method would
523 + reproduce these energy differences exactly (even if the absolute
524 + energies calculated by the methods are different).  Since none of the
525 + methods provide exact energy differences, we used linear least squares
526 + regressions of energy gap data to evaluate how closely the methods
527 + mimicked the Ewald energy gaps.  Unitary results for both the
528 + correlation (slope) and correlation coefficient for these regressions
529 + indicate perfect agreement between the alternative method and SPME.
530 + Sample correlation plots for two alternate methods are shown in
531 + Fig. \ref{fig:linearFit}.
532 +
533   \begin{figure}
534   \centering
535 < \includegraphics[width=3.25in]{./delEplot.pdf}
536 < \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.}
537 < \label{delE}
535 > \includegraphics[width = \linewidth]{./dualLinear.pdf}
536 > \caption{Example least squares regressions of the configuration energy
537 > differences for SPC/E water systems. The upper plot shows a data set
538 > with a poor correlation coefficient ($R^2$), while the lower plot
539 > shows a data set with a good correlation coefficient.}
540 > \label{fig:linearFit}
541   \end{figure}
542  
543 < 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.
543 > Each system type (detailed in section \ref{sec:RepSims}) was
544 > represented using 500 independent configurations.  Additionally, we
545 > used seven different system types, so each of the alternative
546 > (non-Ewald) electrostatic summation methods was evaluated using
547 > 873,250 configurational energy differences.
548  
549 < \subsection{Force Magnitude Comparison}
549 > Results and discussion for the individual analysis of each of the
550 > system types appear in the supporting information, while the
551 > cumulative results over all the investigated systems appears below in
552 > section \ref{sec:EnergyResults}.
553  
554 < 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.
554 > \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
555 > We evaluated the pairwise methods (outlined in section
556 > \ref{sec:ESMethods}) for use in MD simulations by
557 > comparing the force and torque vectors with those obtained using the
558 > reference Ewald summation (SPME).  Both the magnitude and the
559 > direction of these vectors on each of the bodies in the system were
560 > analyzed.  For the magnitude of these vectors, linear least squares
561 > regression analyses were performed as described previously for
562 > comparing $\Delta E$ values.  Instead of a single energy difference
563 > between two system configurations, we compared the magnitudes of the
564 > forces (and torques) on each molecule in each configuration.  For a
565 > system of 1000 water molecules and 40 ions, there are 1040 force
566 > vectors and 1000 torque vectors.  With 500 configurations, this
567 > results in 520,000 force and 500,000 torque vector comparisons.
568 > Additionally, data from seven different system types was aggregated
569 > before the comparison was made.
570  
571 + The {\it directionality} of the force and torque vectors was
572 + investigated through measurement of the angle ($\theta$) formed
573 + between those computed from the particular method and those from SPME,
574 + \begin{equation}
575 + \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
576 + \end{equation}
577 + where $\hat{f}_\textrm{M}$ is the unit vector pointing along the force
578 + vector computed using method M.  Each of these $\theta$ values was
579 + accumulated in a distribution function and weighted by the area on the
580 + unit sphere.  Since this distribution is a measure of angular error
581 + between two different electrostatic summation methods, there is no
582 + {\it a priori} reason for the profile to adhere to any specific
583 + shape. Thus, gaussian fits were used to measure the width of the
584 + resulting distributions.
585 + %
586 + %\begin{figure}
587 + %\centering
588 + %\includegraphics[width = \linewidth]{./gaussFit.pdf}
589 + %\caption{Sample fit of the angular distribution of the force vectors
590 + %accumulated using all of the studied systems.  Gaussian fits were used
591 + %to obtain values for the variance in force and torque vectors.}
592 + %\label{fig:gaussian}
593 + %\end{figure}
594 + %
595 + %Figure \ref{fig:gaussian} shows an example distribution with applied
596 + %non-linear fits.  The solid line is a Gaussian profile, while the
597 + %dotted line is a Voigt profile, a convolution of a Gaussian and a
598 + %Lorentzian.  
599 + %Since this distribution is a measure of angular error between two
600 + %different electrostatic summation methods, there is no {\it a priori}
601 + %reason for the profile to adhere to any specific shape.
602 + %Gaussian fits was used to compare all the tested methods.  
603 + The variance ($\sigma^2$) was extracted from each of these fits and
604 + was used to compare distribution widths.  Values of $\sigma^2$ near
605 + zero indicate vector directions indistinguishable from those
606 + calculated when using the reference method (SPME).
607 +
608 + \subsection{Short-time Dynamics}
609 +
610 + The effects of the alternative electrostatic summation methods on the
611 + short-time dynamics of charged systems were evaluated by considering a
612 + NaCl crystal at a temperature of 1000 K.  A subset of the best
613 + performing pairwise methods was used in this comparison.  The NaCl
614 + crystal was chosen to avoid possible complications from the treatment
615 + of orientational motion in molecular systems.  All systems were
616 + started with the same initial positions and velocities.  Simulations
617 + were performed under the microcanonical ensemble, and velocity
618 + autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
619 + of the trajectories,
620 + \begin{equation}
621 + C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
622 + \label{eq:vCorr}
623 + \end{equation}
624 + Velocity autocorrelation functions require detailed short time data,
625 + thus velocity information was saved every 2 fs over 10 ps
626 + trajectories. Because the NaCl crystal is composed of two different
627 + atom types, the average of the two resulting velocity autocorrelation
628 + functions was used for comparisons.
629 +
630 + \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
631 +
632 + The effects of the same subset of alternative electrostatic methods on
633 + the {\it long-time} dynamics of charged systems were evaluated using
634 + the same model system (NaCl crystals at 1000K).  The power spectrum
635 + ($I(\omega)$) was obtained via Fourier transform of the velocity
636 + autocorrelation function, \begin{equation} I(\omega) =
637 + \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
638 + \label{eq:powerSpec}
639 + \end{equation}
640 + where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
641 + NaCl crystal is composed of two different atom types, the average of
642 + the two resulting power spectra was used for comparisons. Simulations
643 + were performed under the microcanonical ensemble, and velocity
644 + information was saved every 5 fs over 100 ps trajectories.
645 +
646 + \subsection{Representative Simulations}\label{sec:RepSims}
647 + A variety of representative simulations were analyzed to determine the
648 + relative effectiveness of the pairwise summation techniques in
649 + reproducing the energetics and dynamics exhibited by SPME.  We wanted
650 + to span the space of modern simulations (i.e. from liquids of neutral
651 + molecules to ionic crystals), so the systems studied were:
652 + \begin{enumerate}
653 + \item liquid water (SPC/E),\cite{Berendsen87}
654 + \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
655 + \item NaCl crystals,
656 + \item NaCl melts,
657 + \item a low ionic strength solution of NaCl in water (0.11 M),
658 + \item a high ionic strength solution of NaCl in water (1.1 M), and
659 + \item a 6 \AA\  radius sphere of Argon in water.
660 + \end{enumerate}
661 + By utilizing the pairwise techniques (outlined in section
662 + \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
663 + charged particles, and mixtures of the two, we hope to discern under
664 + which conditions it will be possible to use one of the alternative
665 + summation methodologies instead of the Ewald sum.
666 +
667 + For the solid and liquid water configurations, configurations were
668 + taken at regular intervals from high temperature trajectories of 1000
669 + SPC/E water molecules.  Each configuration was equilibrated
670 + independently at a lower temperature (300~K for the liquid, 200~K for
671 + the crystal).  The solid and liquid NaCl systems consisted of 500
672 + $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions.  Configurations for
673 + these systems were selected and equilibrated in the same manner as the
674 + water systems.  The equilibrated temperatures were 1000~K for the NaCl
675 + crystal and 7000~K for the liquid. The ionic solutions were made by
676 + solvating 4 (or 40) ions in a periodic box containing 1000 SPC/E water
677 + molecules.  Ion and water positions were then randomly swapped, and
678 + the resulting configurations were again equilibrated individually.
679 + Finally, for the Argon / Water ``charge void'' systems, the identities
680 + of all the SPC/E waters within 6 \AA\ of the center of the
681 + equilibrated water configurations were converted to argon.
682 + %(Fig. \ref{fig:argonSlice}).
683 +
684 + These procedures guaranteed us a set of representative configurations
685 + from chemically-relevant systems sampled from an appropriate
686 + ensemble. Force field parameters for the ions and Argon were taken
687 + from the force field utilized by {\sc oopse}.\cite{Meineke05}
688 +
689 + %\begin{figure}
690 + %\centering
691 + %\includegraphics[width = \linewidth]{./slice.pdf}
692 + %\caption{A slice from the center of a water box used in a charge void
693 + %simulation.  The darkened region represents the boundary sphere within
694 + %which the water molecules were converted to argon atoms.}
695 + %\label{fig:argonSlice}
696 + %\end{figure}
697 +
698 + \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
699 + We compared the following alternative summation methods with results
700 + from the reference method (SPME):
701 + \begin{itemize}
702 + \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
703 + and 0.3 \AA$^{-1}$,
704 + \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
705 + and 0.3 \AA$^{-1}$,
706 + \item reaction field with an infinite dielectric constant, and
707 + \item an unmodified cutoff.
708 + \end{itemize}
709 + Group-based cutoffs with a fifth-order polynomial switching function
710 + were utilized for the reaction field simulations.  Additionally, we
711 + investigated the use of these cutoffs with the SP, SF, and pure
712 + cutoff.  The SPME electrostatics were performed using the TINKER
713 + implementation of SPME,\cite{Ponder87} while all other method
714 + calculations were performed using the OOPSE molecular mechanics
715 + package.\cite{Meineke05} All other portions of the energy calculation
716 + (i.e. Lennard-Jones interactions) were handled in exactly the same
717 + manner across all systems and configurations.
718 +
719 + The althernative methods were also evaluated with three different
720 + cutoff radii (9, 12, and 15 \AA).  As noted previously, the
721 + convergence parameter ($\alpha$) plays a role in the balance of the
722 + real-space and reciprocal-space portions of the Ewald calculation.
723 + Typical molecular mechanics packages set this to a value dependent on
724 + the cutoff radius and a tolerance (typically less than $1 \times
725 + 10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with
726 + increased accuracy at the expense of increased time spent calculating
727 + the reciprocal-space portion of the
728 + summation.\cite{Perram88,Essmann95} The default TINKER tolerance of $1
729 + \times 10^{-8}$ kcal/mol was used in all SPME calculations, resulting
730 + in Ewald Coefficients of 0.4200, 0.3119, and 0.2476 \AA$^{-1}$ for
731 + cutoff radii of 9, 12, and 15 \AA\ respectively.
732 +
733 + \section{Results and Discussion}
734 +
735 + \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
736 + In order to evaluate the performance of the pairwise electrostatic
737 + summation methods for Monte Carlo simulations, the energy differences
738 + between configurations were compared to the values obtained when using
739 + SPME.  The results for the subsequent regression analysis are shown in
740 + figure \ref{fig:delE}.
741 +
742   \begin{figure}
743   \centering
744 < \includegraphics[width=3.25in]{./frcMagplot.pdf}
745 < \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.}
746 < \label{frcMag}
744 > \includegraphics[width=5.5in]{./delEplot.pdf}
745 > \caption{Statistical analysis of the quality of configurational energy
746 > differences for a given electrostatic method compared with the
747 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
748 > indicate $\Delta E$ values indistinguishable from those obtained using
749 > SPME.  Different values of the cutoff radius are indicated with
750 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
751 > inverted triangles).}
752 > \label{fig:delE}
753   \end{figure}
754  
755 < 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.
755 > The most striking feature of this plot is how well the Shifted Force
756 > ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
757 > differences.  For the undamped {\sc sf} method, and the
758 > moderately-damped {\sc sp} methods, the results are nearly
759 > indistinguishable from the Ewald results.  The other common methods do
760 > significantly less well.  
761  
762 < \subsection{Torque Magnitude Comparison}
762 > The unmodified cutoff method is essentially unusable.  This is not
763 > surprising since hard cutoffs give large energy fluctuations as atoms
764 > or molecules move in and out of the cutoff
765 > radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
766 > some degree by using group based cutoffs with a switching
767 > function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
768 > significant improvement using the group-switched cutoff because the
769 > salt and salt solution systems contain non-neutral groups.  Interested
770 > readers can consult the accompanying supporting information for a
771 > comparison where all groups are neutral.
772  
773 + For the {\sc sp} method, inclusion of potential damping improves the
774 + agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
775 + an excellent correlation and quality of fit with the SPME results,
776 + particularly with a cutoff radius greater than 12
777 + \AA .  Use of a larger damping parameter is more helpful for the
778 + shortest cutoff shown, but it has a detrimental effect on simulations
779 + with larger cutoffs.  
780 +
781 + In the {\sc sf} sets, increasing damping results in progressively
782 + worse correlation with Ewald.  Overall, the undamped case is the best
783 + performing set, as the correlation and quality of fits are
784 + consistently superior regardless of the cutoff distance.  The undamped
785 + case is also less computationally demanding (because no evaluation of
786 + the complementary error function is required).
787 +
788 + The reaction field results illustrates some of that method's
789 + limitations, primarily that it was developed for use in homogenous
790 + systems; although it does provide results that are an improvement over
791 + those from an unmodified cutoff.
792 +
793 + \subsection{Magnitudes of the Force and Torque Vectors}
794 +
795 + Evaluation of pairwise methods for use in Molecular Dynamics
796 + simulations requires consideration of effects on the forces and
797 + torques.  Investigation of the force and torque vector magnitudes
798 + provides a measure of the strength of these values relative to SPME.
799 + Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
800 + force and torque vector magnitude regression results for the
801 + accumulated analysis over all the system types.
802 +
803   \begin{figure}
804   \centering
805 < \includegraphics[width=3.25in]{./trqMagplot.pdf}
806 < \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.}
807 < \label{trqMag}
805 > \includegraphics[width=5.5in]{./frcMagplot.pdf}
806 > \caption{Statistical analysis of the quality of the force vector
807 > magnitudes for a given electrostatic method compared with the
808 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
809 > indicate force magnitude values indistinguishable from those obtained
810 > using SPME.  Different values of the cutoff radius are indicated with
811 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
812 > inverted triangles).}
813 > \label{fig:frcMag}
814   \end{figure}
815  
816 < 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.
816 > Figure \ref{fig:frcMag}, for the most part, parallels the results seen
817 > in the previous $\Delta E$ section.  The unmodified cutoff results are
818 > poor, but using group based cutoffs and a switching function provides
819 > a improvement much more significant than what was seen with $\Delta
820 > E$.  Looking at the {\sc sp} sets, the slope and $R^2$
821 > improve with the use of damping to an optimal result of 0.2 \AA
822 > $^{-1}$ for the 12 and 15 \AA\ cutoffs.  Further increases in damping,
823 > while beneficial for simulations with a cutoff radius of 9 \AA\ , is
824 > detrimental to simulations with larger cutoff radii.  The undamped
825 > {\sc sf} method gives forces in line with those obtained using
826 > SPME, and use of a damping function results in minor improvement.  The
827 > reaction field results are surprisingly good, considering the poor
828 > quality of the fits for the $\Delta E$ results.  There is still a
829 > considerable degree of scatter in the data, but it correlates well in
830 > general.  To be fair, we again note that the reaction field
831 > calculations do not encompass NaCl crystal and melt systems, so these
832 > results are partly biased towards conditions in which the method
833 > performs more favorably.
834  
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
835   \begin{figure}
836   \centering
837 < \includegraphics[width=3.25in]{./gaussFit.pdf}
838 < \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.}
839 < \label{gaussian}
837 > \includegraphics[width=5.5in]{./trqMagplot.pdf}
838 > \caption{Statistical analysis of the quality of the torque vector
839 > magnitudes for a given electrostatic method compared with the
840 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
841 > indicate torque magnitude values indistinguishable from those obtained
842 > using SPME.  Different values of the cutoff radius are indicated with
843 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
844 > inverted triangles).}
845 > \label{fig:trqMag}
846   \end{figure}
847  
848 < 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.  
848 > To evaluate the torque vector magnitudes, the data set from which
849 > values are drawn is limited to rigid molecules in the systems
850 > (i.e. water molecules).  In spite of this smaller sampling pool, the
851 > torque vector magnitude results in figure \ref{fig:trqMag} are still
852 > similar to those seen for the forces; however, they more clearly show
853 > the improved behavior that comes with increasing the cutoff radius.
854 > Moderate damping is beneficial to the {\sc sp} and helpful
855 > yet possibly unnecessary with the {\sc sf} method, and they also
856 > show that over-damping adversely effects all cutoff radii rather than
857 > showing an improvement for systems with short cutoffs.  The reaction
858 > field method performs well when calculating the torques, better than
859 > the Shifted Force method over this limited data set.
860  
861 + \subsection{Directionality of the Force and Torque Vectors}
862 +
863 + Having force and torque vectors with magnitudes that are well
864 + correlated to SPME is good, but if they are not pointing in the proper
865 + direction the results will be incorrect.  These vector directions were
866 + investigated through measurement of the angle formed between them and
867 + those from SPME.  The results (Fig. \ref{fig:frcTrqAng}) are compared
868 + through the variance ($\sigma^2$) of the Gaussian fits of the angle
869 + error distributions of the combined set over all system types.
870 +
871   \begin{figure}
872   \centering
873 < \includegraphics[width=3.25in]{./frcTrqAngplot.pdf}
874 < \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.}
875 < \label{frcTrqAng}
873 > \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
874 > \caption{Statistical analysis of the quality of the Gaussian fit of
875 > the force and torque vector angular distributions for a given
876 > electrostatic method compared with the reference Ewald sum.  Results
877 > with a variance ($\sigma^2$) equal to zero (dashed line) indicate
878 > force and torque directions indistinguishable from those obtained
879 > using SPME.  Different values of the cutoff radius are indicated with
880 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
881 > inverted triangles).}
882 > \label{fig:frcTrqAng}
883   \end{figure}
884  
885 < 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.  
885 > Both the force and torque $\sigma^2$ results from the analysis of the
886 > total accumulated system data are tabulated in figure
887 > \ref{fig:frcTrqAng}.  All of the sets, aside from the over-damped case
888 > show the improvement afforded by choosing a longer simulation cutoff.
889 > Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
890 > of the distribution widths, with a similar improvement going from 12
891 > to 15 \AA .  The undamped {\sc sf}, Group Based Cutoff, and
892 > Reaction Field methods all do equivalently well at capturing the
893 > direction of both the force and torque vectors.  Using damping
894 > improves the angular behavior significantly for the {\sc sp}
895 > and moderately for the {\sc sf} methods.  Increasing the damping
896 > too far is destructive for both methods, particularly to the torque
897 > vectors.  Again it is important to recognize that the force vectors
898 > cover all particles in the systems, while torque vectors are only
899 > available for neutral molecular groups.  Damping appears to have a
900 > more beneficial effect on non-neutral bodies, and this observation is
901 > investigated further in the accompanying supporting information.
902  
903   \begin{table}[htbp]
904     \centering
905 <   \caption{Variance ($\sigma^2$) of the force (top set) and torque (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods.  Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function.  The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}  
905 >   \caption{Variance ($\sigma^2$) of the force (top set) and torque
906 > (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods.  Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function.  The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}      
907     \begin{tabular}{@{} ccrrrrrrrr @{}}
908        \\
909        \toprule
# Line 173 | Line 931 | Both the force and torque $\sigma^2$ results from the
931  
932        \bottomrule
933     \end{tabular}
934 <   \label{groupAngle}
934 >   \label{tab:groupAngle}
935   \end{table}
936  
937 < 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.  
937 > Although not discussed previously, group based cutoffs can be applied
938 > to both the {\sc sp} and {\sc sf} methods.  Use off a
939 > switching function corrects for the discontinuities that arise when
940 > atoms of a group exit the cutoff before the group's center of mass.
941 > Though there are no significant benefit or drawbacks observed in
942 > $\Delta E$ and vector magnitude results when doing this, there is a
943 > measurable improvement in the vector angle results.  Table
944 > \ref{tab:groupAngle} shows the angular variance values obtained using
945 > group based cutoffs and a switching function alongside the standard
946 > results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
947 > The {\sc sp} shows much narrower angular distributions for
948 > both the force and torque vectors when using an $\alpha$ of 0.2
949 > \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
950 > undamped and lightly damped cases.  Thus, by calculating the
951 > electrostatic interactions in terms of molecular pairs rather than
952 > atomic pairs, the direction of the force and torque vectors are
953 > determined more accurately.
954  
955 < 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.
955 > One additional trend to recognize in table \ref{tab:groupAngle} is
956 > that the $\sigma^2$ values for both {\sc sp} and
957 > {\sc sf} converge as $\alpha$ increases, something that is easier
958 > to see when using group based cutoffs.  Looking back on figures
959 > \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
960 > behavior clearly at large $\alpha$ and cutoff values.  The reason for
961 > this is that the complimentary error function inserted into the
962 > potential weakens the electrostatic interaction as $\alpha$ increases.
963 > Thus, at larger values of $\alpha$, both the summation method types
964 > progress toward non-interacting functions, so care is required in
965 > choosing large damping functions lest one generate an undesirable loss
966 > in the pair interaction.  Kast \textit{et al.}  developed a method for
967 > choosing appropriate $\alpha$ values for these types of electrostatic
968 > summation methods by fitting to $g(r)$ data, and their methods
969 > indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
970 > values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
971 > to be reasonable choices to obtain proper MC behavior
972 > (Fig. \ref{fig:delE}); however, based on these findings, choices this
973 > high would introduce error in the molecular torques, particularly for
974 > the shorter cutoffs.  Based on the above findings, empirical damping
975 > up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
976 > unnecessary when using the {\sc sf} method.
977  
978 + \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
979 +
980 + In the previous studies using a {\sc sf} variant of the damped
981 + Wolf coulomb potential, the structure and dynamics of water were
982 + investigated rather extensively.\cite{Zahn02,Kast03} Their results
983 + indicated that the damped {\sc sf} method results in properties
984 + very similar to those obtained when using the Ewald summation.
985 + Considering the statistical results shown above, the good performance
986 + of this method is not that surprising.  Rather than consider the same
987 + systems and simply recapitulate their results, we decided to look at
988 + the solid state dynamical behavior obtained using the best performing
989 + summation methods from the above results.
990 +
991 + \begin{figure}
992 + \centering
993 + \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
994 + \caption{Velocity auto-correlation functions of NaCl crystals at
995 + 1000 K while using SPME, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and
996 + {\sc sp} ($\alpha$ = 0.2). The inset is a magnification of the first
997 + trough. The times to first collision are nearly identical, but the
998 + differences can be seen in the peaks and troughs, where the undamped
999 + to weakly damped methods are stiffer than the moderately damped and
1000 + SPME methods.}
1001 + \label{fig:vCorrPlot}
1002 + \end{figure}
1003 +
1004 + The short-time decays through the first collision are nearly identical
1005 + in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
1006 + functions show how the methods differ.  The undamped {\sc sf} method
1007 + has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
1008 + peaks than any of the other methods.  As the damping function is
1009 + increased, these peaks are smoothed out, and approach the SPME
1010 + curve. The damping acts as a distance dependent Gaussian screening of
1011 + the point charges for the pairwise summation methods; thus, the
1012 + collisions are more elastic in the undamped {\sc sf} potential, and the
1013 + stiffness of the potential is diminished as the electrostatic
1014 + interactions are softened by the damping function.  With $\alpha$
1015 + values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
1016 + nearly identical and track the SPME features quite well.  This is not
1017 + too surprising in that the differences between the {\sc sf} and {\sc
1018 + sp} potentials are mitigated with increased damping.  However, this
1019 + appears to indicate that once damping is utilized, the form of the
1020 + potential seems to play a lesser role in the crystal dynamics.
1021 +
1022 + \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1023 +
1024 + The short time dynamics were extended to evaluate how the differences
1025 + between the methods affect the collective long-time motion.  The same
1026 + electrostatic summation methods were used as in the short time
1027 + velocity autocorrelation function evaluation, but the trajectories
1028 + were sampled over a much longer time. The power spectra of the
1029 + resulting velocity autocorrelation functions were calculated and are
1030 + displayed in figure \ref{fig:methodPS}.
1031 +
1032 + \begin{figure}
1033 + \centering
1034 + \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1035 + \caption{Power spectra obtained from the velocity auto-correlation
1036 + functions of NaCl crystals at 1000 K while using SPME, {\sc sf}
1037 + ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).
1038 + Apodization of the correlation functions via a cubic switching
1039 + function between 40 and 50 ps was used to clear up the spectral noise
1040 + resulting from data truncation, and had no noticeable effect on peak
1041 + location or magnitude.  The inset shows the frequency region below 100
1042 + cm$^{-1}$ to highlight where the spectra begin to differ.}
1043 + \label{fig:methodPS}
1044 + \end{figure}
1045 +
1046 + While high frequency peaks of the spectra in this figure overlap,
1047 + showing the same general features, the low frequency region shows how
1048 + the summation methods differ.  Considering the low-frequency inset
1049 + (expanded in the upper frame of figure \ref{fig:dampInc}), at
1050 + frequencies below 100 cm$^{-1}$, the correlated motions are
1051 + blue-shifted when using undamped or weakly damped {\sc sf}.  When
1052 + using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
1053 + and {\sc sp} methods give near identical correlated motion behavior as
1054 + the Ewald method (which has a damping value of 0.3119).  This
1055 + weakening of the electrostatic interaction with increased damping
1056 + explains why the long-ranged correlated motions are at lower
1057 + frequencies for the moderately damped methods than for undamped or
1058 + weakly damped methods.  To see this effect more clearly, we show how
1059 + damping strength alone affects a simple real-space electrostatic
1060 + potential,
1061 + \begin{equation}
1062 + V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
1063 + \end{equation}
1064 + where $S(r)$ is a switching function that smoothly zeroes the
1065 + potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how
1066 + the low frequency motions are dependent on the damping used in the
1067 + direct electrostatic sum.  As the damping increases, the peaks drop to
1068 + lower frequencies.  Incidentally, use of an $\alpha$ of 0.25
1069 + \AA$^{-1}$ on a simple electrostatic summation results in low
1070 + frequency correlated dynamics equivalent to a simulation using SPME.
1071 + When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1072 + shift to higher frequency in exponential fashion.  Though not shown,
1073 + the spectrum for the simple undamped electrostatic potential is
1074 + blue-shifted such that the lowest frequency peak resides near 325
1075 + cm$^{-1}$.  In light of these results, the undamped {\sc sf} method
1076 + producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1077 + respectable and shows that the shifted force procedure accounts for
1078 + most of the effect afforded through use of the Ewald summation.
1079 + However, it appears as though moderate damping is required for
1080 + accurate reproduction of crystal dynamics.
1081 + \begin{figure}
1082 + \centering
1083 + \includegraphics[width = \linewidth]{./comboSquare.pdf}
1084 + \caption{Regions of spectra showing the low-frequency correlated
1085 + motions for NaCl crystals at 1000 K using various electrostatic
1086 + summation methods.  The upper plot is a zoomed inset from figure
1087 + \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential
1088 + increases, the low-frequency peaks red-shift.  The lower plot is of
1089 + spectra when using SPME and a simple damped Coulombic sum with damping
1090 + coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$.  As
1091 + $\alpha$ increases, the peaks are red-shifted toward and eventually
1092 + beyond the values given by SPME.  The larger $\alpha$ values weaken
1093 + the real-space electrostatics, explaining this shift towards less
1094 + strongly correlated motions in the crystal.}
1095 + \label{fig:dampInc}
1096 + \end{figure}
1097 +
1098   \section{Conclusions}
1099  
1100 < \section{Acknowledgments}
1100 > This investigation of pairwise electrostatic summation techniques
1101 > shows that there are viable and more computationally efficient
1102 > electrostatic summation techniques than the Ewald summation, chiefly
1103 > methods derived from the damped Coulombic sum originally proposed by
1104 > Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1105 > {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1106 > shows a remarkable ability to reproduce the energetic and dynamic
1107 > characteristics exhibited by simulations employing lattice summation
1108 > techniques.  The cumulative energy difference results showed the
1109 > undamped {\sc sf} and moderately damped {\sc sp} methods
1110 > produced results nearly identical to SPME.  Similarly for the dynamic
1111 > features, the undamped or moderately damped {\sc sf} and
1112 > moderately damped {\sc sp} methods produce force and torque
1113 > vector magnitude and directions very similar to the expected values.
1114 > These results translate into long-time dynamic behavior equivalent to
1115 > that produced in simulations using SPME.
1116  
1117 + Aside from the computational cost benefit, these techniques have
1118 + applicability in situations where the use of the Ewald sum can prove
1119 + problematic.  Primary among them is their use in interfacial systems,
1120 + where the unmodified lattice sum techniques artificially accentuate
1121 + the periodicity of the system in an undesirable manner.  There have
1122 + been alterations to the standard Ewald techniques, via corrections and
1123 + reformulations, to compensate for these systems; but the pairwise
1124 + techniques discussed here require no modifications, making them
1125 + natural tools to tackle these problems.  Additionally, this
1126 + transferability gives them benefits over other pairwise methods, like
1127 + reaction field, because estimations of physical properties (e.g. the
1128 + dielectric constant) are unnecessary.
1129 +
1130 + We are not suggesting any flaw with the Ewald sum; in fact, it is the
1131 + standard by which these simple pairwise sums are judged.  However,
1132 + these results do suggest that in the typical simulations performed
1133 + today, the Ewald summation may no longer be required to obtain the
1134 + level of accuracy most researchers have come to expect
1135 +
1136 + \section{Acknowledgments}
1137   \newpage
1138  
1139 < \bibliographystyle{achemso}
1139 > \bibliographystyle{jcp2}
1140   \bibliography{electrostaticMethods}
1141  
1142  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines