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 2575 by chrisfen, Sat Jan 28 22:42:46 2006 UTC vs.
Revision 2741 by chrisfen, Wed Apr 26 15:22:09 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}
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{bibentry}
15 + \usepackage{mathrsfs}
16   \usepackage[ref]{overcite}
17   \pagestyle{plain}
18   \pagenumbering{arabic}
# Line 20 | Line 25
25  
26   \begin{document}
27  
28 < \title{On the necessity of the Ewald Summation in molecular simulations.}
28 > \title{Is the Ewald summation still necessary? \\
29 > Pairwise alternatives to the accepted standard for \\
30 > long-range electrostatics in molecular simulations}
31  
32 < \author{Christopher J. Fennell and J. Daniel Gezelter \\
32 > \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
33 > gezelter@nd.edu} \\
34   Department of Chemistry and Biochemistry\\
35   University of Notre Dame\\
36   Notre Dame, Indiana 46556}
# Line 30 | Line 38 | Notre Dame, Indiana 46556}
38   \date{\today}
39  
40   \maketitle
41 < %\doublespacing
41 > \doublespacing
42  
43   \begin{abstract}
44 + We investigate pairwise electrostatic interaction methods and show
45 + that there are viable and computationally efficient $(\mathscr{O}(N))$
46 + alternatives to the Ewald summation for typical modern molecular
47 + simulations.  These methods are extended from the damped and
48 + cutoff-neutralized Coulombic sum originally proposed by
49 + [D. Wolf, P. Keblinski, S.~R. Phillpot, and J. Eggebrecht, {\it J. Chem. Phys.} {\bf 110}, 8255 (1999)] One of these, the damped shifted force method, shows
50 + a remarkable ability to reproduce the energetic and dynamic
51 + characteristics exhibited by simulations employing lattice summation
52 + techniques.  Comparisons were performed with this and other pairwise
53 + methods against the smooth particle mesh Ewald ({\sc spme}) summation
54 + to see how well they reproduce the energetics and dynamics of a
55 + variety of simulation types.
56   \end{abstract}
57  
58 + \newpage
59 +
60   %\narrowtext
61  
62 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63   %                              BODY OF TEXT
64 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65  
66   \section{Introduction}
67 +
68 + In molecular simulations, proper accumulation of the electrostatic
69 + interactions is essential and is one of the most
70 + computationally-demanding tasks.  The common molecular mechanics force
71 + fields represent atomic sites with full or partial charges protected
72 + by Lennard-Jones (short range) interactions.  This means that nearly
73 + every pair interaction involves a calculation of charge-charge forces.
74 + Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
75 + interactions quickly become the most expensive part of molecular
76 + simulations.  Historically, the electrostatic pair interaction would
77 + not have decayed appreciably within the typical box lengths that could
78 + be feasibly simulated.  In the larger systems that are more typical of
79 + modern simulations, large cutoffs should be used to incorporate
80 + electrostatics correctly.
81 +
82 + There have been many efforts to address the proper and practical
83 + handling of electrostatic interactions, and these have resulted in a
84 + variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
85 + typically classified as implicit methods (i.e., continuum dielectrics,
86 + static dipolar fields),\cite{Born20,Grossfield00} explicit methods
87 + (i.e., Ewald summations, interaction shifting or
88 + truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
89 + reaction field type methods, fast multipole
90 + methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
91 + often preferred because they physically incorporate solvent molecules
92 + in the system of interest, but these methods are sometimes difficult
93 + to utilize because of their high computational cost.\cite{Roux99} In
94 + addition to the computational cost, there have been some questions
95 + regarding possible artifacts caused by the inherent periodicity of the
96 + explicit Ewald summation.\cite{Tobias01}
97 +
98 + In this paper, we focus on a new set of pairwise methods devised by
99 + Wolf {\it et al.},\cite{Wolf99} which we further extend.  These
100 + methods along with a few other mixed methods (i.e. reaction field) are
101 + compared with the smooth particle mesh Ewald
102 + sum,\cite{Onsager36,Essmann99} which is our reference method for
103 + handling long-range electrostatic interactions. The new methods for
104 + handling electrostatics have the potential to scale linearly with
105 + increasing system size since they involve only a simple modification
106 + to the direct pairwise sum.  They also lack the added periodicity of
107 + the Ewald sum, so they can be used for systems which are non-periodic
108 + or which have one- or two-dimensional periodicity.  Below, these
109 + methods are evaluated using a variety of model systems to
110 + establish their usability in molecular simulations.
111 +
112 + \subsection{The Ewald Sum}
113 + The complete accumulation of the electrostatic interactions in a system with
114 + periodic boundary conditions (PBC) requires the consideration of the
115 + effect of all charges within a (cubic) simulation box as well as those
116 + in the periodic replicas,
117 + \begin{equation}
118 + 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],
119 + \label{eq:PBCSum}
120 + \end{equation}
121 + where the sum over $\mathbf{n}$ is a sum over all periodic box
122 + replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
123 + prime indicates $i = j$ are neglected for $\mathbf{n} =
124 + 0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
125 + particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
126 + the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
127 + $j$, and $\phi$ is the solution to Poisson's equation
128 + ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
129 + charge-charge interactions). In the case of monopole electrostatics,
130 + eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
131 + non-neutral systems.
132 +
133 + The electrostatic summation problem was originally studied by Ewald
134 + for the case of an infinite crystal.\cite{Ewald21}. The approach he
135 + took was to convert this conditionally convergent sum into two
136 + absolutely convergent summations: a short-ranged real-space summation
137 + and a long-ranged reciprocal-space summation,
138 + \begin{equation}
139 + \begin{split}
140 + 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,
141 + \end{split}
142 + \label{eq:EwaldSum}
143 + \end{equation}
144 + where $\alpha$ is the damping or convergence parameter with units of
145 + \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
146 + $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
147 + constant of the surrounding medium. The final two terms of
148 + eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
149 + for interacting with a surrounding dielectric.\cite{Allen87} This
150 + dipolar term was neglected in early applications in molecular
151 + simulations,\cite{Brush66,Woodcock71} until it was introduced by de
152 + Leeuw {\it et al.} to address situations where the unit cell has a
153 + dipole moment which is magnified through replication of the periodic
154 + images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
155 + system is said to be using conducting (or ``tin-foil'') boundary
156 + conditions, $\epsilon_{\rm S} = \infty$. Figure
157 + \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
158 + time.  Initially, due to the small system sizes that could be
159 + simulated feasibly, the entire simulation box was replicated to
160 + convergence.  In more modern simulations, the systems have grown large
161 + enough that a real-space cutoff could potentially give convergent
162 + behavior.  Indeed, it has been observed that with the choice of a
163 + small $\alpha$, the reciprocal-space portion of the Ewald sum can be
164 + rapidly convergent and small relative to the real-space
165 + portion.\cite{Karasawa89,Kolafa92}
166 +
167 + \begin{figure}
168 + \centering
169 + \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
170 + \caption{The change in the need for the Ewald sum with
171 + increasing computational power.  A:~Initially, only small systems
172 + could be studied, and the Ewald sum replicated the simulation box to
173 + convergence.  B:~Now, radial cutoff methods should be able to reach
174 + convergence for the larger systems of charges that are common today.}
175 + \label{fig:ewaldTime}
176 + \end{figure}
177 +
178 + The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The
179 + convergence parameter $(\alpha)$ plays an important role in balancing
180 + the computational cost between the direct and reciprocal-space
181 + portions of the summation.  The choice of this value allows one to
182 + select whether the real-space or reciprocal space portion of the
183 + summation is an $\mathscr{O}(N^2)$ calculation (with the other being
184 + $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
185 + $\alpha$ and thoughtful algorithm development, this cost can be
186 + reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
187 + taken to reduce the cost of the Ewald summation even further is to set
188 + $\alpha$ such that the real-space interactions decay rapidly, allowing
189 + for a short spherical cutoff. Then the reciprocal space summation is
190 + optimized.  These optimizations usually involve utilization of the
191 + fast Fourier transform (FFT),\cite{Hockney81} leading to the
192 + particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
193 + methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
194 + methods, the cost of the reciprocal-space portion of the Ewald
195 + summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
196 + \log N)$.
197 +
198 + These developments and optimizations have made the use of the Ewald
199 + summation routine in simulations with periodic boundary
200 + conditions. However, in certain systems, such as vapor-liquid
201 + interfaces and membranes, the intrinsic three-dimensional periodicity
202 + can prove problematic.  The Ewald sum has been reformulated to handle
203 + 2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
204 + new methods are computationally expensive.\cite{Spohr97,Yeh99} More
205 + recently, there have been several successful efforts toward reducing
206 + the computational cost of 2D lattice summations, often enabling the
207 + use of the mentioned
208 + optimizations.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
209 +
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 + \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 includes
232 + image-charges subtracted out through placement on the cutoff sphere
233 + and a distance-dependent damping function (identical to that seen in
234 + the real-space portion of the Ewald sum) to aid convergence
235 + \begin{equation}
236 + 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\}.
237 + \label{eq:WolfPot}
238 + \end{equation}
239 + Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
240 + potential.  However, neutralizing the charge contained within each
241 + cutoff sphere requires the placement of a self-image charge on the
242 + surface of the cutoff sphere.  This additional self-term in the total
243 + potential enabled Wolf {\it et al.}  to obtain excellent estimates of
244 + Madelung energies for many crystals.
245 +
246 + In order to use their charge-neutralized potential in molecular
247 + dynamics simulations, Wolf \textit{et al.} suggested taking the
248 + derivative of this potential prior to evaluation of the limit.  This
249 + procedure gives an expression for the forces,
250 + \begin{equation}
251 + 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\},
252 + \label{eq:WolfForces}
253 + \end{equation}
254 + that incorporates both image charges and damping of the electrostatic
255 + interaction.
256 +
257 + More recently, Zahn \textit{et al.} investigated these potential and
258 + force expressions for use in simulations involving water.\cite{Zahn02}
259 + In their work, they pointed out that the forces and derivative of
260 + the potential are not commensurate.  Attempts to use both
261 + eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
262 + to poor energy conservation.  They correctly observed that taking the
263 + limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
264 + derivatives gives forces for a different potential energy function
265 + than the one shown in eq. (\ref{eq:WolfPot}).
266 +
267 + Zahn \textit{et al.} introduced a modified form of this summation
268 + method as a way to use the technique in Molecular Dynamics
269 + simulations.  They proposed a new damped Coulomb potential,
270 + \begin{equation}
271 + 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\},
272 + \label{eq:ZahnPot}
273 + \end{equation}
274 + and showed that this potential does fairly well at capturing the
275 + structural and dynamic properties of water compared the same
276 + properties obtained using the Ewald sum.
277 +
278 + \subsection{Simple Forms for Pairwise Electrostatics}
279 +
280 + The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
281 + al.} are constructed using two different (and separable) computational
282 + tricks: \begin{enumerate}
283 + \item shifting through the use of image charges, and
284 + \item damping the electrostatic interaction.
285 + \end{enumerate}  Wolf \textit{et al.} treated the
286 + development of their summation method as a progressive application of
287 + these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
288 + their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
289 + post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
290 + both techniques.  It is possible, however, to separate these
291 + tricks and study their effects independently.
292 +
293 + Starting with the original observation that the effective range of the
294 + electrostatic interaction in condensed phases is considerably less
295 + than $r^{-1}$, either the cutoff sphere neutralization or the
296 + distance-dependent damping technique could be used as a foundation for
297 + a new pairwise summation method.  Wolf \textit{et al.} made the
298 + observation that charge neutralization within the cutoff sphere plays
299 + a significant role in energy convergence; therefore we will begin our
300 + analysis with the various shifted forms that maintain this charge
301 + neutralization.  We can evaluate the methods of Wolf
302 + \textit{et al.}  and Zahn \textit{et al.} by considering the standard
303 + shifted potential,
304 + \begin{equation}
305 + V_\textrm{SP}(r) =      \begin{cases}
306 + v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
307 + R_\textrm{c}  
308 + \end{cases},
309 + \label{eq:shiftingPotForm}
310 + \end{equation}
311 + and shifted force,
312 + \begin{equation}
313 + V_\textrm{SF}(r) =      \begin{cases}
314 + v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
315 + &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
316 +                                                \end{cases},
317 + \label{eq:shiftingForm}
318 + \end{equation}
319 + functions where $v(r)$ is the unshifted form of the potential, and
320 + $v_c$ is $v(R_\textrm{c})$.  The Shifted Force ({\sc sf}) form ensures
321 + that both the potential and the forces goes to zero at the cutoff
322 + radius, while the Shifted Potential ({\sc sp}) form only ensures the
323 + potential is smooth at the cutoff radius
324 + ($R_\textrm{c}$).\cite{Allen87}
325  
326 + The forces associated with the shifted potential are simply the forces
327 + of the unshifted potential itself (when inside the cutoff sphere),
328 + \begin{equation}
329 + F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
330 + \end{equation}
331 + and are zero outside.  Inside the cutoff sphere, the forces associated
332 + with the shifted force form can be written,
333 + \begin{equation}
334 + F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
335 + v(r)}{dr} \right)_{r=R_\textrm{c}}.
336 + \end{equation}
337 +
338 + If the potential, $v(r)$, is taken to be the normal Coulomb potential,
339 + \begin{equation}
340 + v(r) = \frac{q_i q_j}{r},
341 + \label{eq:Coulomb}
342 + \end{equation}
343 + then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
344 + al.}'s undamped prescription:
345 + \begin{equation}
346 + V_\textrm{SP}(r) =
347 + q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
348 + r\leqslant R_\textrm{c},
349 + \label{eq:SPPot}
350 + \end{equation}
351 + with associated forces,
352 + \begin{equation}
353 + F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
354 + \label{eq:SPForces}
355 + \end{equation}
356 + These forces are identical to the forces of the standard Coulomb
357 + interaction, and cutting these off at $R_c$ was addressed by Wolf
358 + \textit{et al.} as undesirable.  They pointed out that the effect of
359 + the image charges is neglected in the forces when this form is
360 + used,\cite{Wolf99} thereby eliminating any benefit from the method in
361 + molecular dynamics.  Additionally, there is a discontinuity in the
362 + forces at the cutoff radius which results in energy drift during MD
363 + simulations.
364 +
365 + The shifted force ({\sc sf}) form using the normal Coulomb potential
366 + will give,
367 + \begin{equation}
368 + 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}.
369 + \label{eq:SFPot}
370 + \end{equation}
371 + with associated forces,
372 + \begin{equation}
373 + 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}.
374 + \label{eq:SFForces}
375 + \end{equation}
376 + This formulation has the benefits that there are no discontinuities at
377 + the cutoff radius, while the neutralizing image charges are present in
378 + both the energy and force expressions.  It would be simple to add the
379 + self-neutralizing term back when computing the total energy of the
380 + system, thereby maintaining the agreement with the Madelung energies.
381 + A side effect of this treatment is the alteration in the shape of the
382 + potential that comes from the derivative term.  Thus, a degree of
383 + clarity about agreement with the empirical potential is lost in order
384 + to gain functionality in dynamics simulations.
385 +
386 + Wolf \textit{et al.} originally discussed the energetics of the
387 + shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
388 + insufficient for accurate determination of the energy with reasonable
389 + cutoff distances.  The calculated Madelung energies fluctuated around
390 + the expected value as the cutoff radius was increased, but the
391 + oscillations converged toward the correct value.\cite{Wolf99} A
392 + damping function was incorporated to accelerate the convergence; and
393 + though alternative forms for the damping function could be
394 + used,\cite{Jones56,Heyes81} the complimentary error function was
395 + chosen to mirror the effective screening used in the Ewald summation.
396 + Incorporating this error function damping into the simple Coulomb
397 + potential,
398 + \begin{equation}
399 + v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
400 + \label{eq:dampCoulomb}
401 + \end{equation}
402 + the shifted potential (eq. (\ref{eq:SPPot})) becomes
403 + \begin{equation}
404 + 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},
405 + \label{eq:DSPPot}
406 + \end{equation}
407 + with associated forces,
408 + \begin{equation}
409 + 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}.
410 + \label{eq:DSPForces}
411 + \end{equation}
412 + Again, this damped shifted potential suffers from a
413 + force-discontinuity at the cutoff radius, and the image charges play
414 + no role in the forces.  To remedy these concerns, one may derive a
415 + {\sc sf} variant by including the derivative term in
416 + eq. (\ref{eq:shiftingForm}),
417 + \begin{equation}
418 + \begin{split}
419 + 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}.
420 + \label{eq:DSFPot}
421 + \end{split}
422 + \end{equation}
423 + The derivative of the above potential will lead to the following forces,
424 + \begin{equation}
425 + \begin{split}
426 + 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}.
427 + \label{eq:DSFForces}
428 + \end{split}
429 + \end{equation}
430 + If the damping parameter $(\alpha)$ is set to zero, the undamped case,
431 + eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
432 + recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
433 +
434 + This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
435 + derived by Zahn \textit{et al.}; however, there are two important
436 + differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
437 + eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
438 + with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present
439 + in the Zahn potential, resulting in a potential discontinuity as
440 + particles cross $R_\textrm{c}$.  Second, the sign of the derivative
441 + portion is different.  The missing $v_\textrm{c}$ term would not
442 + affect molecular dynamics simulations (although the computed energy
443 + would be expected to have sudden jumps as particle distances crossed
444 + $R_c$).  The sign problem is a potential source of errors, however.
445 + In fact, it introduces a discontinuity in the forces at the cutoff,
446 + because the force function is shifted in the wrong direction and
447 + doesn't cross zero at $R_\textrm{c}$.
448 +
449 + Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
450 + electrostatic summation method in which the potential and forces are
451 + continuous at the cutoff radius and which incorporates the damping
452 + function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
453 + this paper, we will evaluate exactly how good these methods ({\sc sp},
454 + {\sc sf}, damping) are at reproducing the correct electrostatic
455 + summation performed by the Ewald sum.
456 +
457 + \subsection{Other alternatives}
458 + In addition to the methods described above, we considered some other
459 + techniques that are commonly used in molecular simulations.  The
460 + simplest of these is group-based cutoffs.  Though of little use for
461 + charged molecules, collecting atoms into neutral groups takes
462 + advantage of the observation that the electrostatic interactions decay
463 + faster than those for monopolar pairs.\cite{Steinbach94} When
464 + considering these molecules as neutral groups, the relative
465 + orientations of the molecules control the strength of the interactions
466 + at the cutoff radius.  Consequently, as these molecular particles move
467 + through $R_\textrm{c}$, the energy will drift upward due to the
468 + anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
469 + maintain good energy conservation, both the potential and derivative
470 + need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
471 + This is accomplished using a standard switching function.  If a smooth
472 + second derivative is desired, a fifth (or higher) order polynomial can
473 + be used.\cite{Andrea83}
474 +
475 + Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
476 + and to incorporate the effects of the surroundings, a method like
477 + Reaction Field ({\sc rf}) can be used.  The original theory for {\sc
478 + rf} was originally developed by Onsager,\cite{Onsager36} and it was
479 + applied in simulations for the study of water by Barker and
480 + Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
481 + an extension of the group-based cutoff method where the net dipole
482 + within the cutoff sphere polarizes an external dielectric, which
483 + reacts back on the central dipole.  The same switching function
484 + considerations for group-based cutoffs need to made for {\sc rf}, with
485 + the additional pre-specification of a dielectric constant.
486 +
487   \section{Methods}
488 +
489 + In classical molecular mechanics simulations, there are two primary
490 + techniques utilized to obtain information about the system of
491 + interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
492 + techniques utilize pairwise summations of interactions between
493 + particle sites, but they use these summations in different ways.
494 +
495 + In MC, the potential energy difference between configurations dictates
496 + the progression of MC sampling.  Going back to the origins of this
497 + method, the acceptance criterion for the canonical ensemble laid out
498 + by Metropolis \textit{et al.} states that a subsequent configuration
499 + is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
500 + $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
501 + Maintaining the correct $\Delta E$ when using an alternate method for
502 + handling the long-range electrostatics will ensure proper sampling
503 + from the ensemble.
504 +
505 + In MD, the derivative of the potential governs how the system will
506 + progress in time.  Consequently, the force and torque vectors on each
507 + body in the system dictate how the system evolves.  If the magnitude
508 + and direction of these vectors are similar when using alternate
509 + electrostatic summation techniques, the dynamics in the short term
510 + will be indistinguishable.  Because error in MD calculations is
511 + cumulative, one should expect greater deviation at longer times,
512 + although methods which have large differences in the force and torque
513 + vectors will diverge from each other more rapidly.
514 +
515 + \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
516 +
517 + The pairwise summation techniques (outlined in section
518 + \ref{sec:ESMethods}) were evaluated for use in MC simulations by
519 + studying the energy differences between conformations.  We took the
520 + {\sc spme}-computed energy difference between two conformations to be the
521 + correct behavior. An ideal performance by an alternative method would
522 + reproduce these energy differences exactly (even if the absolute
523 + energies calculated by the methods are different).  Since none of the
524 + methods provide exact energy differences, we used linear least squares
525 + regressions of energy gap data to evaluate how closely the methods
526 + mimicked the Ewald energy gaps.  Unitary results for both the
527 + correlation (slope) and correlation coefficient for these regressions
528 + indicate perfect agreement between the alternative method and {\sc spme}.
529 + Sample correlation plots for two alternate methods are shown in
530 + Fig. \ref{fig:linearFit}.
531 +
532 + \begin{figure}
533 + \centering
534 + \includegraphics[width = \linewidth]{./dualLinear.pdf}
535 + \caption{Example least squares regressions of the configuration energy
536 + differences for SPC/E water systems. The upper plot shows a data set
537 + with a poor correlation coefficient ($R^2$), while the lower plot
538 + shows a data set with a good correlation coefficient.}
539 + \label{fig:linearFit}
540 + \end{figure}
541 +
542 + Each of the seven system types (detailed in section \ref{sec:RepSims})
543 + were represented using 500 independent configurations.  Thus, each of
544 + the alternative (non-Ewald) electrostatic summation methods was
545 + evaluated using an accumulated 873,250 configurational energy
546 + differences.
547 +
548 + Results and discussion for the individual analysis of each of the
549 + system types appear in the supporting information, while the
550 + cumulative results over all the investigated systems appears below in
551 + section \ref{sec:EnergyResults}.
552 +
553 + \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
554 + We evaluated the pairwise methods (outlined in section
555 + \ref{sec:ESMethods}) for use in MD simulations by
556 + comparing the force and torque vectors with those obtained using the
557 + reference Ewald summation ({\sc spme}).  Both the magnitude and the
558 + direction of these vectors on each of the bodies in the system were
559 + analyzed.  For the magnitude of these vectors, linear least squares
560 + regression analyses were performed as described previously for
561 + comparing $\Delta E$ values.  Instead of a single energy difference
562 + between two system configurations, we compared the magnitudes of the
563 + forces (and torques) on each molecule in each configuration.  For a
564 + system of 1000 water molecules and 40 ions, there are 1040 force
565 + vectors and 1000 torque vectors.  With 500 configurations, this
566 + results in 520,000 force and 500,000 torque vector comparisons.
567 + Additionally, data from seven different system types was aggregated
568 + before the comparison was made.
569 +
570 + The {\it directionality} of the force and torque vectors was
571 + investigated through measurement of the angle ($\theta$) formed
572 + between those computed from the particular method and those from {\sc spme},
573 + \begin{equation}
574 + \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
575 + \end{equation}
576 + where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
577 + vector computed using method M.  Each of these $\theta$ values was
578 + accumulated in a distribution function and weighted by the area on the
579 + unit sphere.  Since this distribution is a measure of angular error
580 + between two different electrostatic summation methods, there is no
581 + {\it a priori} reason for the profile to adhere to any specific
582 + shape. Thus, gaussian fits were used to measure the width of the
583 + resulting distributions. The variance ($\sigma^2$) was extracted from
584 + each of these fits and was used to compare distribution widths.
585 + Values of $\sigma^2$ near zero indicate vector directions
586 + indistinguishable from those calculated when using the reference
587 + method ({\sc spme}).
588  
589 + \subsection{Short-time Dynamics}
590 +
591 + The effects of the alternative electrostatic summation methods on the
592 + short-time dynamics of charged systems were evaluated by considering a
593 + NaCl crystal at a temperature of 1000 K.  A subset of the best
594 + performing pairwise methods was used in this comparison.  The NaCl
595 + crystal was chosen to avoid possible complications from the treatment
596 + of orientational motion in molecular systems.  All systems were
597 + started with the same initial positions and velocities.  Simulations
598 + were performed under the microcanonical ensemble, and velocity
599 + autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
600 + of the trajectories,
601 + \begin{equation}
602 + C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
603 + \label{eq:vCorr}
604 + \end{equation}
605 + Velocity autocorrelation functions require detailed short time data,
606 + thus velocity information was saved every 2 fs over 10 ps
607 + trajectories. Because the NaCl crystal is composed of two different
608 + atom types, the average of the two resulting velocity autocorrelation
609 + functions was used for comparisons.
610 +
611 + \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
612 +
613 + The effects of the same subset of alternative electrostatic methods on
614 + the {\it long-time} dynamics of charged systems were evaluated using
615 + the same model system (NaCl crystals at 1000~K).  The power spectrum
616 + ($I(\omega)$) was obtained via Fourier transform of the velocity
617 + autocorrelation function, \begin{equation} I(\omega) =
618 + \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
619 + \label{eq:powerSpec}
620 + \end{equation}
621 + where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
622 + NaCl crystal is composed of two different atom types, the average of
623 + the two resulting power spectra was used for comparisons. Simulations
624 + were performed under the microcanonical ensemble, and velocity
625 + information was saved every 5~fs over 100~ps trajectories.
626 +
627 + \subsection{Representative Simulations}\label{sec:RepSims}
628 + A variety of representative molecular simulations were analyzed to
629 + determine the relative effectiveness of the pairwise summation
630 + techniques in reproducing the energetics and dynamics exhibited by
631 + {\sc spme}.  We wanted to span the space of typical molecular
632 + simulations (i.e. from liquids of neutral molecules to ionic
633 + crystals), so the systems studied were:
634 + \begin{enumerate}
635 + \item liquid water (SPC/E),\cite{Berendsen87}
636 + \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
637 + \item NaCl crystals,
638 + \item NaCl melts,
639 + \item a low ionic strength solution of NaCl in water (0.11 M),
640 + \item a high ionic strength solution of NaCl in water (1.1 M), and
641 + \item a 6 \AA\  radius sphere of Argon in water.
642 + \end{enumerate}
643 + By utilizing the pairwise techniques (outlined in section
644 + \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
645 + charged particles, and mixtures of the two, we hope to discern under
646 + which conditions it will be possible to use one of the alternative
647 + summation methodologies instead of the Ewald sum.
648 +
649 + For the solid and liquid water configurations, configurations were
650 + taken at regular intervals from high temperature trajectories of 1000
651 + SPC/E water molecules.  Each configuration was equilibrated
652 + independently at a lower temperature (300~K for the liquid, 200~K for
653 + the crystal).  The solid and liquid NaCl systems consisted of 500
654 + $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions.  Configurations for
655 + these systems were selected and equilibrated in the same manner as the
656 + water systems. In order to introduce measurable fluctuations in the
657 + configuration energy differences, the crystalline simulations were
658 + equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid
659 + NaCl configurations needed to represent a fully disordered array of
660 + point charges, so the high temperature of 7000~K was selected for
661 + equilibration. The ionic solutions were made by solvating 4 (or 40)
662 + ions in a periodic box containing 1000 SPC/E water molecules.  Ion and
663 + water positions were then randomly swapped, and the resulting
664 + configurations were again equilibrated individually.  Finally, for the
665 + Argon / Water ``charge void'' systems, the identities of all the SPC/E
666 + waters within 6 \AA\ of the center of the equilibrated water
667 + configurations were converted to argon.
668 +
669 + These procedures guaranteed us a set of representative configurations
670 + from chemically-relevant systems sampled from appropriate
671 + ensembles. Force field parameters for the ions and Argon were taken
672 + from the force field utilized by {\sc oopse}.\cite{Meineke05}
673 +
674 + \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
675 + We compared the following alternative summation methods with results
676 + from the reference method ({\sc spme}):
677 + \begin{itemize}
678 + \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
679 + and 0.3 \AA$^{-1}$,
680 + \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
681 + and 0.3 \AA$^{-1}$,
682 + \item reaction field with an infinite dielectric constant, and
683 + \item an unmodified cutoff.
684 + \end{itemize}
685 + Group-based cutoffs with a fifth-order polynomial switching function
686 + were utilized for the reaction field simulations.  Additionally, we
687 + investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
688 + cutoff.  The {\sc spme} electrostatics were performed using the {\sc tinker}
689 + implementation of {\sc spme},\cite{Ponder87} while all other calculations
690 + were performed using the {\sc oopse} molecular mechanics
691 + package.\cite{Meineke05} All other portions of the energy calculation
692 + (i.e. Lennard-Jones interactions) were handled in exactly the same
693 + manner across all systems and configurations.
694 +
695 + The alternative methods were also evaluated with three different
696 + cutoff radii (9, 12, and 15 \AA).  As noted previously, the
697 + convergence parameter ($\alpha$) plays a role in the balance of the
698 + real-space and reciprocal-space portions of the Ewald calculation.
699 + Typical molecular mechanics packages set this to a value dependent on
700 + the cutoff radius and a tolerance (typically less than $1 \times
701 + 10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with
702 + increasing accuracy at the expense of computational time spent on the
703 + reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
704 + The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
705 + in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
706 + 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
707 + respectively.
708 +
709   \section{Results and Discussion}
710  
711 + \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
712 + In order to evaluate the performance of the pairwise electrostatic
713 + summation methods for Monte Carlo simulations, the energy differences
714 + between configurations were compared to the values obtained when using
715 + {\sc spme}.  The results for the subsequent regression analysis are shown in
716 + figure \ref{fig:delE}.
717 +
718 + \begin{figure}
719 + \centering
720 + \includegraphics[width=5.5in]{./delEplot.pdf}
721 + \caption{Statistical analysis of the quality of configurational energy
722 + differences for a given electrostatic method compared with the
723 + reference Ewald sum.  Results with a value equal to 1 (dashed line)
724 + indicate $\Delta E$ values indistinguishable from those obtained using
725 + {\sc spme}.  Different values of the cutoff radius are indicated with
726 + different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
727 + inverted triangles).}
728 + \label{fig:delE}
729 + \end{figure}
730 +
731 + The most striking feature of this plot is how well the Shifted Force
732 + ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
733 + differences.  For the undamped {\sc sf} method, and the
734 + moderately-damped {\sc sp} methods, the results are nearly
735 + indistinguishable from the Ewald results.  The other common methods do
736 + significantly less well.  
737 +
738 + The unmodified cutoff method is essentially unusable.  This is not
739 + surprising since hard cutoffs give large energy fluctuations as atoms
740 + or molecules move in and out of the cutoff
741 + radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
742 + some degree by using group based cutoffs with a switching
743 + function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
744 + significant improvement using the group-switched cutoff because the
745 + salt and salt solution systems contain non-neutral groups.  Interested
746 + readers can consult the accompanying supporting information for a
747 + comparison where all groups are neutral.
748 +
749 + For the {\sc sp} method, inclusion of electrostatic damping improves
750 + the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
751 + shows an excellent correlation and quality of fit with the {\sc spme}
752 + results, particularly with a cutoff radius greater than 12
753 + \AA .  Use of a larger damping parameter is more helpful for the
754 + shortest cutoff shown, but it has a detrimental effect on simulations
755 + with larger cutoffs.  
756 +
757 + In the {\sc sf} sets, increasing damping results in progressively {\it
758 + worse} correlation with Ewald.  Overall, the undamped case is the best
759 + performing set, as the correlation and quality of fits are
760 + consistently superior regardless of the cutoff distance.  The undamped
761 + case is also less computationally demanding (because no evaluation of
762 + the complementary error function is required).
763 +
764 + The reaction field results illustrates some of that method's
765 + limitations, primarily that it was developed for use in homogenous
766 + systems; although it does provide results that are an improvement over
767 + those from an unmodified cutoff.
768 +
769 + \subsection{Magnitudes of the Force and Torque Vectors}
770 +
771 + Evaluation of pairwise methods for use in Molecular Dynamics
772 + simulations requires consideration of effects on the forces and
773 + torques.  Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
774 + regression results for the force and torque vector magnitudes,
775 + respectively.  The data in these figures was generated from an
776 + accumulation of the statistics from all of the system types.
777 +
778 + \begin{figure}
779 + \centering
780 + \includegraphics[width=5.5in]{./frcMagplot.pdf}
781 + \caption{Statistical analysis of the quality of the force vector
782 + magnitudes for a given electrostatic method compared with the
783 + reference Ewald sum.  Results with a value equal to 1 (dashed line)
784 + indicate force magnitude values indistinguishable from those obtained
785 + using {\sc spme}.  Different values of the cutoff radius are indicated with
786 + different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
787 + inverted triangles).}
788 + \label{fig:frcMag}
789 + \end{figure}
790 +
791 + Again, it is striking how well the Shifted Potential and Shifted Force
792 + methods are doing at reproducing the {\sc spme} forces.  The undamped and
793 + weakly-damped {\sc sf} method gives the best agreement with Ewald.
794 + This is perhaps expected because this method explicitly incorporates a
795 + smooth transition in the forces at the cutoff radius as well as the
796 + neutralizing image charges.
797 +
798 + Figure \ref{fig:frcMag}, for the most part, parallels the results seen
799 + in the previous $\Delta E$ section.  The unmodified cutoff results are
800 + poor, but using group based cutoffs and a switching function provides
801 + an improvement much more significant than what was seen with $\Delta
802 + E$.
803 +
804 + With moderate damping and a large enough cutoff radius, the {\sc sp}
805 + method is generating usable forces.  Further increases in damping,
806 + while beneficial for simulations with a cutoff radius of 9 \AA\ , is
807 + detrimental to simulations with larger cutoff radii.
808 +
809 + The reaction field results are surprisingly good, considering the poor
810 + quality of the fits for the $\Delta E$ results.  There is still a
811 + considerable degree of scatter in the data, but the forces correlate
812 + well with the Ewald forces in general.  We note that the reaction
813 + field calculations do not include the pure NaCl systems, so these
814 + results are partly biased towards conditions in which the method
815 + performs more favorably.
816 +
817 + \begin{figure}
818 + \centering
819 + \includegraphics[width=5.5in]{./trqMagplot.pdf}
820 + \caption{Statistical analysis of the quality of the torque vector
821 + magnitudes for a given electrostatic method compared with the
822 + reference Ewald sum.  Results with a value equal to 1 (dashed line)
823 + indicate torque magnitude values indistinguishable from those obtained
824 + using {\sc spme}.  Different values of the cutoff radius are indicated with
825 + different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
826 + inverted triangles).}
827 + \label{fig:trqMag}
828 + \end{figure}
829 +
830 + Molecular torques were only available from the systems which contained
831 + rigid molecules (i.e. the systems containing water).  The data in
832 + fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
833 +
834 + Torques appear to be much more sensitive to charges at a longer
835 + distance.   The striking feature in comparing the new electrostatic
836 + methods with {\sc spme} is how much the agreement improves with increasing
837 + cutoff radius.  Again, the weakly damped and undamped {\sc sf} method
838 + appears to be reproducing the {\sc spme} torques most accurately.  
839 +
840 + Water molecules are dipolar, and the reaction field method reproduces
841 + the effect of the surrounding polarized medium on each of the
842 + molecular bodies. Therefore it is not surprising that reaction field
843 + performs best of all of the methods on molecular torques.
844 +
845 + \subsection{Directionality of the Force and Torque Vectors}
846 +
847 + It is clearly important that a new electrostatic method can reproduce
848 + the magnitudes of the force and torque vectors obtained via the Ewald
849 + sum. However, the {\it directionality} of these vectors will also be
850 + vital in calculating dynamical quantities accurately.  Force and
851 + torque directionalities were investigated by measuring the angles
852 + formed between these vectors and the same vectors calculated using
853 + {\sc spme}.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the
854 + variance ($\sigma^2$) of the Gaussian fits of the angle error
855 + distributions of the combined set over all system types.
856 +
857 + \begin{figure}
858 + \centering
859 + \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
860 + \caption{Statistical analysis of the width of the angular distribution
861 + that the force and torque vectors from a given electrostatic method
862 + make with their counterparts obtained using the reference Ewald sum.
863 + Results with a variance ($\sigma^2$) equal to zero (dashed line)
864 + indicate force and torque directions indistinguishable from those
865 + obtained using {\sc spme}.  Different values of the cutoff radius are
866 + indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
867 + and 15\AA\ = inverted triangles).}
868 + \label{fig:frcTrqAng}
869 + \end{figure}
870 +
871 + Both the force and torque $\sigma^2$ results from the analysis of the
872 + total accumulated system data are tabulated in figure
873 + \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
874 + sp}) method would be essentially unusable for molecular dynamics
875 + unless the damping function is added.  The Shifted Force ({\sc sf})
876 + method, however, is generating force and torque vectors which are
877 + within a few degrees of the Ewald results even with weak (or no)
878 + damping.
879 +
880 + All of the sets (aside from the over-damped case) show the improvement
881 + afforded by choosing a larger cutoff radius.  Increasing the cutoff
882 + from 9 to 12 \AA\ typically results in a halving of the width of the
883 + distribution, with a similar improvement when going from 12 to 15
884 + \AA .
885 +
886 + The undamped {\sc sf}, group-based cutoff, and reaction field methods
887 + all do equivalently well at capturing the direction of both the force
888 + and torque vectors.  Using the electrostatic damping improves the
889 + angular behavior significantly for the {\sc sp} and moderately for the
890 + {\sc sf} methods.  Overdamping is detrimental to both methods.  Again
891 + it is important to recognize that the force vectors cover all
892 + particles in all seven systems, while torque vectors are only
893 + available for neutral molecular groups.  Damping is more beneficial to
894 + charged bodies, and this observation is investigated further in the
895 + accompanying supporting information.
896 +
897 + Although not discussed previously, group based cutoffs can be applied
898 + to both the {\sc sp} and {\sc sf} methods.  The group-based cutoffs
899 + will reintroduce small discontinuities at the cutoff radius, but the
900 + effects of these can be minimized by utilizing a switching function.
901 + Though there are no significant benefits or drawbacks observed in
902 + $\Delta E$ and the force and torque magnitudes when doing this, there
903 + is a measurable improvement in the directionality of the forces and
904 + torques. Table \ref{tab:groupAngle} shows the angular variances
905 + obtained using group based cutoffs along with the results seen in
906 + figure \ref{fig:frcTrqAng}.  The {\sc sp} (with an $\alpha$ of 0.2
907 + \AA$^{-1}$ or smaller) shows much narrower angular distributions when
908 + using group-based cutoffs. The {\sc sf} method likewise shows
909 + improvement in the undamped and lightly damped cases.
910 +
911 + \begin{table}[htbp]
912 +   \centering
913 +   \caption{Statistical analysis of the angular
914 +   distributions that the force (upper) and torque (lower) vectors
915 +   from a given electrostatic method make with their counterparts
916 +   obtained using the reference Ewald sum.  Calculations were
917 +   performed both with (Y) and without (N) group based cutoffs and a
918 +   switching function.  The $\alpha$ values have units of \AA$^{-1}$
919 +   and the variance values have units of degrees$^2$.}
920 +
921 +   \begin{tabular}{@{} ccrrrrrrrr @{}}
922 +      \\
923 +      \toprule
924 +      & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
925 +      \cmidrule(lr){3-6}
926 +      \cmidrule(l){7-10}
927 +            $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
928 +      \midrule
929 +    
930 + 9 \AA   & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
931 +        & \textbf{Y} & \textbf{2.486} & \textbf{2.160} & \textbf{0.667} & \textbf{0.608} & \textbf{1.768} & \textbf{1.766} & \textbf{0.676} & \textbf{0.609} \\
932 + 12 \AA  & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
933 +        & \textbf{Y} & \textbf{0.515} & \textbf{0.288} & \textbf{0.127} & \textbf{0.586} & \textbf{0.308} & \textbf{0.249} & \textbf{0.127} & \textbf{0.586} \\
934 + 15 \AA  & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
935 +        & \textbf{Y} & \textbf{0.228} & \textbf{0.099} & \textbf{0.121} & \textbf{0.598} & \textbf{0.144} & \textbf{0.090} & \textbf{0.121} & \textbf{0.598} \\      
936 +
937 +      \midrule
938 +      
939 + 9 \AA   & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
940 +        & \textbf{Y} & \textbf{2.115} & \textbf{1.914} & \textbf{1.878} & \textbf{5.142} & \textbf{2.076} & \textbf{2.039} & \textbf{1.972} & \textbf{5.146} \\
941 + 12 \AA  & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
942 +        & \textbf{Y} & \textbf{0.810} & \textbf{0.685} & \textbf{1.352} & \textbf{5.082} & \textbf{0.765} & \textbf{0.714} & \textbf{1.360} & \textbf{5.082} \\
943 + 15 \AA  & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
944 +        & \textbf{Y} & \textbf{0.282} & \textbf{0.294} & \textbf{1.272} & \textbf{4.999} & \textbf{0.324} & \textbf{0.318} & \textbf{1.272} & \textbf{4.999} \\
945 +
946 +      \bottomrule
947 +   \end{tabular}
948 +   \label{tab:groupAngle}
949 + \end{table}
950 +
951 + One additional trend in table \ref{tab:groupAngle} is that the
952 + $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
953 + increases, something that is more obvious with group-based cutoffs.
954 + The complimentary error function inserted into the potential weakens
955 + the electrostatic interaction as the value of $\alpha$ is increased.
956 + However, at larger values of $\alpha$, it is possible to overdamp the
957 + electrostatic interaction and to remove it completely.  Kast
958 + \textit{et al.}  developed a method for choosing appropriate $\alpha$
959 + values for these types of electrostatic summation methods by fitting
960 + to $g(r)$ data, and their methods indicate optimal values of 0.34,
961 + 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
962 + respectively.\cite{Kast03} These appear to be reasonable choices to
963 + obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
964 + these findings, choices this high would introduce error in the
965 + molecular torques, particularly for the shorter cutoffs.  Based on our
966 + observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
967 + but damping may be unnecessary when using the {\sc sf} method.
968 +
969 + \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
970 +
971 + Zahn {\it et al.} investigated the structure and dynamics of water
972 + using eqs. (\ref{eq:ZahnPot}) and
973 + (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
974 + that a method similar (but not identical with) the damped {\sc sf}
975 + method resulted in properties very similar to those obtained when
976 + using the Ewald summation.  The properties they studied (pair
977 + distribution functions, diffusion constants, and velocity and
978 + orientational correlation functions) may not be particularly sensitive
979 + to the long-range and collective behavior that governs the
980 + low-frequency behavior in crystalline systems.  Additionally, the
981 + ionic crystals are the worst case scenario for the pairwise methods
982 + because they lack the reciprocal space contribution contained in the
983 + Ewald summation.  
984 +
985 + We are using two separate measures to probe the effects of these
986 + alternative electrostatic methods on the dynamics in crystalline
987 + materials.  For short- and intermediate-time dynamics, we are
988 + computing the velocity autocorrelation function, and for long-time
989 + and large length-scale collective motions, we are looking at the
990 + low-frequency portion of the power spectrum.
991 +
992 + \begin{figure}
993 + \centering
994 + \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
995 + \caption{Velocity autocorrelation functions of NaCl crystals at
996 + 1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
997 + sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
998 + the first minimum.  The times to first collision are nearly identical,
999 + but differences can be seen in the peaks and troughs, where the
1000 + undamped and weakly damped methods are stiffer than the moderately
1001 + damped and {\sc spme} methods.}
1002 + \label{fig:vCorrPlot}
1003 + \end{figure}
1004 +
1005 + The short-time decay of the velocity autocorrelation function through
1006 + the first collision are nearly identical in figure
1007 + \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1008 + how the methods differ.  The undamped {\sc sf} method has deeper
1009 + troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1010 + any of the other methods.  As the damping parameter ($\alpha$) is
1011 + increased, these peaks are smoothed out, and the {\sc sf} method
1012 + approaches the {\sc spme} results.  With $\alpha$ values of 0.2 \AA$^{-1}$,
1013 + the {\sc sf} and {\sc sp} functions are nearly identical and track the
1014 + {\sc spme} features quite well.  This is not surprising because the {\sc sf}
1015 + and {\sc sp} potentials become nearly identical with increased
1016 + damping.  However, this appears to indicate that once damping is
1017 + utilized, the details of the form of the potential (and forces)
1018 + constructed out of the damped electrostatic interaction are less
1019 + important.
1020 +
1021 + \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1022 +
1023 + To evaluate how the differences between the methods affect the
1024 + collective long-time motion, we computed power spectra from long-time
1025 + traces of the velocity autocorrelation function. The power spectra for
1026 + the best-performing alternative methods are shown in
1027 + fig. \ref{fig:methodPS}.  Apodization of the correlation functions via
1028 + a cubic switching function between 40 and 50 ps was used to reduce the
1029 + ringing resulting from data truncation.  This procedure had no
1030 + noticeable effect on peak location or magnitude.
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 {\sc spme}, {\sc sf}
1037 + ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).  The inset
1038 + shows the frequency region below 100 cm$^{-1}$ to highlight where the
1039 + spectra differ.}
1040 + \label{fig:methodPS}
1041 + \end{figure}
1042 +
1043 + While the high frequency regions of the power spectra for the
1044 + alternative methods are quantitatively identical with Ewald spectrum,
1045 + the low frequency region shows how the summation methods differ.
1046 + Considering the low-frequency inset (expanded in the upper frame of
1047 + figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1048 + correlated motions are blue-shifted when using undamped or weakly
1049 + damped {\sc sf}.  When using moderate damping ($\alpha = 0.2$
1050 + \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1051 + correlated motion to the Ewald method (which has a convergence
1052 + parameter of 0.3119 \AA$^{-1}$).  This weakening of the electrostatic
1053 + interaction with increased damping explains why the long-ranged
1054 + correlated motions are at lower frequencies for the moderately damped
1055 + methods than for undamped or weakly damped methods.
1056 +
1057 + To isolate the role of the damping constant, we have computed the
1058 + spectra for a single method ({\sc sf}) with a range of damping
1059 + constants and compared this with the {\sc spme} spectrum.
1060 + Fig. \ref{fig:dampInc} shows more clearly that increasing the
1061 + electrostatic damping red-shifts the lowest frequency phonon modes.
1062 + However, even without any electrostatic damping, the {\sc sf} method
1063 + has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1064 + Without the {\sc sf} modifications, an undamped (pure cutoff) method
1065 + would predict the lowest frequency peak near 325 cm$^{-1}$.  {\it
1066 + Most} of the collective behavior in the crystal is accurately captured
1067 + using the {\sc sf} method.  Quantitative agreement with Ewald can be
1068 + obtained using moderate damping in addition to the shifting at the
1069 + cutoff distance.
1070 +
1071 + \begin{figure}
1072 + \centering
1073 + \includegraphics[width = \linewidth]{./increasedDamping.pdf}
1074 + \caption{Effect of damping on the two lowest-frequency phonon modes in
1075 + the NaCl crystal at 1000~K.  The undamped shifted force ({\sc sf})
1076 + method is off by less than 10 cm$^{-1}$, and increasing the
1077 + electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1078 + with the power spectrum obtained using the Ewald sum.  Overdamping can
1079 + result in underestimates of frequencies of the long-wavelength
1080 + motions.}
1081 + \label{fig:dampInc}
1082 + \end{figure}
1083 +
1084   \section{Conclusions}
1085  
1086 + This investigation of pairwise electrostatic summation techniques
1087 + shows that there are viable and computationally efficient alternatives
1088 + to the Ewald summation.  These methods are derived from the damped and
1089 + cutoff-neutralized Coulombic sum originally proposed by Wolf
1090 + \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1091 + method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1092 + (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1093 + energetic and dynamic characteristics exhibited by simulations
1094 + employing lattice summation techniques.  The cumulative energy
1095 + difference results showed the undamped {\sc sf} and moderately damped
1096 + {\sc sp} methods produced results nearly identical to {\sc spme}.  Similarly
1097 + for the dynamic features, the undamped or moderately damped {\sc sf}
1098 + and moderately damped {\sc sp} methods produce force and torque vector
1099 + magnitude and directions very similar to the expected values.  These
1100 + results translate into long-time dynamic behavior equivalent to that
1101 + produced in simulations using {\sc spme}.
1102 +
1103 + As in all purely-pairwise cutoff methods, these methods are expected
1104 + to scale approximately {\it linearly} with system size, and they are
1105 + easily parallelizable.  This should result in substantial reductions
1106 + in the computational cost of performing large simulations.
1107 +
1108 + Aside from the computational cost benefit, these techniques have
1109 + applicability in situations where the use of the Ewald sum can prove
1110 + problematic.  Of greatest interest is their potential use in
1111 + interfacial systems, where the unmodified lattice sum techniques
1112 + artificially accentuate the periodicity of the system in an
1113 + undesirable manner.  There have been alterations to the standard Ewald
1114 + techniques, via corrections and reformulations, to compensate for
1115 + these systems; but the pairwise techniques discussed here require no
1116 + modifications, making them natural tools to tackle these problems.
1117 + Additionally, this transferability gives them benefits over other
1118 + pairwise methods, like reaction field, because estimations of physical
1119 + properties (e.g. the dielectric constant) are unnecessary.
1120 +
1121 + If a researcher is using Monte Carlo simulations of large chemical
1122 + systems containing point charges, most structural features will be
1123 + accurately captured using the undamped {\sc sf} method or the {\sc sp}
1124 + method with an electrostatic damping of 0.2 \AA$^{-1}$.  These methods
1125 + would also be appropriate for molecular dynamics simulations where the
1126 + data of interest is either structural or short-time dynamical
1127 + quantities.  For long-time dynamics and collective motions, the safest
1128 + pairwise method we have evaluated is the {\sc sf} method with an
1129 + electrostatic damping between 0.2 and 0.25
1130 + \AA$^{-1}$.
1131 +
1132 + We are not suggesting that there is any flaw with the Ewald sum; in
1133 + fact, it is the standard by which these simple pairwise sums have been
1134 + judged.  However, these results do suggest that in the typical
1135 + simulations performed today, the Ewald summation may no longer be
1136 + required to obtain the level of accuracy most researchers have come to
1137 + expect.
1138 +
1139   \section{Acknowledgments}
1140 + Support for this project was provided by the National Science
1141 + Foundation under grant CHE-0134881.  The authors would like to thank
1142 + Steve Corcelli and Ed Maginn for helpful discussions and comments.
1143  
1144 < \newpage
1144 > \newpage
1145  
1146 < \bibliographystyle{achemso}
1146 > \bibliographystyle{jcp2}
1147   \bibliography{electrostaticMethods}
1148  
1149  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines