ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2667
Committed: Fri Mar 24 02:39:59 2006 UTC (18 years, 3 months ago) by chrisfen
Content type: application/x-tex
File size: 60661 byte(s)
Log Message:
Figure adjusted and Steve's recommendations incorporated

File Contents

# Content
1 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 %\documentclass[aps,prb,preprint]{revtex4}
3 \documentclass[11pt]{article}
4 \usepackage{endfloat}
5 \usepackage{amsmath,bm}
6 \usepackage{amssymb}
7 \usepackage{epsf}
8 \usepackage{times}
9 \usepackage{mathptmx}
10 \usepackage{setspace}
11 \usepackage{tabularx}
12 \usepackage{graphicx}
13 \usepackage{booktabs}
14 \usepackage{bibentry}
15 \usepackage{mathrsfs}
16 \usepackage[ref]{overcite}
17 \pagestyle{plain}
18 \pagenumbering{arabic}
19 \oddsidemargin 0.0cm \evensidemargin 0.0cm
20 \topmargin -21pt \headsep 10pt
21 \textheight 9.0in \textwidth 6.5in
22 \brokenpenalty=10000
23 \renewcommand{\baselinestretch}{1.2}
24 \renewcommand\citemid{\ } % no comma in optional reference note
25
26 \begin{document}
27
28 \title{Is the Ewald summation still necessary? \\
29 Pairwise alternatives to the accepted standard for \\
30 long-range electrostatics}
31
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}
37
38 \date{\today}
39
40 \maketitle
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 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63 % BODY OF TEXT
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 establish
110 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 application of the Ewald sum with
171 increasing computational power. A:~Initially, only small systems could
172 be studied, and the Ewald sum replicated the simulation box to
173 convergence. B:~Now, much larger systems of charges can be
174 investigated with fixed-distance cutoffs.}
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}
205 Inclusion of a correction term in the Ewald summation is a possible
206 direction for handling 2D systems while still enabling the use of the
207 modern optimizations.\cite{Yeh99}
208
209 Several studies have recognized that the inherent periodicity in the
210 Ewald sum can also have an effect on three-dimensional
211 systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
212 Solvated proteins are essentially kept at high concentration due to
213 the periodicity of the electrostatic summation method. In these
214 systems, the more compact folded states of a protein can be
215 artificially stabilized by the periodic replicas introduced by the
216 Ewald summation.\cite{Weber00} Thus, care must be taken when
217 considering the use of the Ewald summation where the assumed
218 periodicity would introduce spurious effects in the system dynamics.
219
220 \subsection{The Wolf and Zahn Methods}
221 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
222 for the accurate accumulation of electrostatic interactions in an
223 efficient pairwise fashion. This procedure lacks the inherent
224 periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
225 observed that the electrostatic interaction is effectively
226 short-ranged in condensed phase systems and that neutralization of the
227 charge contained within the cutoff radius is crucial for potential
228 stability. They devised a pairwise summation method that ensures
229 charge neutrality and gives results similar to those obtained with the
230 Ewald summation. The resulting shifted Coulomb potential includes
231 image-charges subtracted out through placement on the cutoff sphere
232 and a distance-dependent damping function (identical to that seen in
233 the real-space portion of the Ewald sum) to aid convergence
234 \begin{equation}
235 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\}.
236 \label{eq:WolfPot}
237 \end{equation}
238 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
239 potential. However, neutralizing the charge contained within each
240 cutoff sphere requires the placement of a self-image charge on the
241 surface of the cutoff sphere. This additional self-term in the total
242 potential enabled Wolf {\it et al.} to obtain excellent estimates of
243 Madelung energies for many crystals.
244
245 In order to use their charge-neutralized potential in molecular
246 dynamics simulations, Wolf \textit{et al.} suggested taking the
247 derivative of this potential prior to evaluation of the limit. This
248 procedure gives an expression for the forces,
249 \begin{equation}
250 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\},
251 \label{eq:WolfForces}
252 \end{equation}
253 that incorporates both image charges and damping of the electrostatic
254 interaction.
255
256 More recently, Zahn \textit{et al.} investigated these potential and
257 force expressions for use in simulations involving water.\cite{Zahn02}
258 In their work, they pointed out that the forces and derivative of
259 the potential are not commensurate. Attempts to use both
260 eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
261 to poor energy conservation. They correctly observed that taking the
262 limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
263 derivatives gives forces for a different potential energy function
264 than the one shown in eq. (\ref{eq:WolfPot}).
265
266 Zahn \textit{et al.} introduced a modified form of this summation
267 method as a way to use the technique in Molecular Dynamics
268 simulations. They proposed a new damped Coulomb potential,
269 \begin{equation}
270 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\},
271 \label{eq:ZahnPot}
272 \end{equation}
273 and showed that this potential does fairly well at capturing the
274 structural and dynamic properties of water compared the same
275 properties obtained using the Ewald sum.
276
277 \subsection{Simple Forms for Pairwise Electrostatics}
278
279 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
280 al.} are constructed using two different (and separable) computational
281 tricks: \begin{enumerate}
282 \item shifting through the use of image charges, and
283 \item damping the electrostatic interaction.
284 \end{enumerate} Wolf \textit{et al.} treated the
285 development of their summation method as a progressive application of
286 these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
287 their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
288 post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
289 both techniques. It is possible, however, to separate these
290 tricks and study their effects independently.
291
292 Starting with the original observation that the effective range of the
293 electrostatic interaction in condensed phases is considerably less
294 than $r^{-1}$, either the cutoff sphere neutralization or the
295 distance-dependent damping technique could be used as a foundation for
296 a new pairwise summation method. Wolf \textit{et al.} made the
297 observation that charge neutralization within the cutoff sphere plays
298 a significant role in energy convergence; therefore we will begin our
299 analysis with the various shifted forms that maintain this charge
300 neutralization. We can evaluate the methods of Wolf
301 \textit{et al.} and Zahn \textit{et al.} by considering the standard
302 shifted potential,
303 \begin{equation}
304 V_\textrm{SP}(r) = \begin{cases}
305 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
306 R_\textrm{c}
307 \end{cases},
308 \label{eq:shiftingPotForm}
309 \end{equation}
310 and shifted force,
311 \begin{equation}
312 V_\textrm{SF}(r) = \begin{cases}
313 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
314 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
315 \end{cases},
316 \label{eq:shiftingForm}
317 \end{equation}
318 functions where $v(r)$ is the unshifted form of the potential, and
319 $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
320 that both the potential and the forces goes to zero at the cutoff
321 radius, while the Shifted Potential ({\sc sp}) form only ensures the
322 potential is smooth at the cutoff radius
323 ($R_\textrm{c}$).\cite{Allen87}
324
325 The forces associated with the shifted potential are simply the forces
326 of the unshifted potential itself (when inside the cutoff sphere),
327 \begin{equation}
328 F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
329 \end{equation}
330 and are zero outside. Inside the cutoff sphere, the forces associated
331 with the shifted force form can be written,
332 \begin{equation}
333 F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
334 v(r)}{dr} \right)_{r=R_\textrm{c}}.
335 \end{equation}
336
337 If the potential, $v(r)$, is taken to be the normal Coulomb potential,
338 \begin{equation}
339 v(r) = \frac{q_i q_j}{r},
340 \label{eq:Coulomb}
341 \end{equation}
342 then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
343 al.}'s undamped prescription:
344 \begin{equation}
345 V_\textrm{SP}(r) =
346 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
347 r\leqslant R_\textrm{c},
348 \label{eq:SPPot}
349 \end{equation}
350 with associated forces,
351 \begin{equation}
352 F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
353 \label{eq:SPForces}
354 \end{equation}
355 These forces are identical to the forces of the standard Coulomb
356 interaction, and cutting these off at $R_c$ was addressed by Wolf
357 \textit{et al.} as undesirable. They pointed out that the effect of
358 the image charges is neglected in the forces when this form is
359 used,\cite{Wolf99} thereby eliminating any benefit from the method in
360 molecular dynamics. Additionally, there is a discontinuity in the
361 forces at the cutoff radius which results in energy drift during MD
362 simulations.
363
364 The shifted force ({\sc sf}) form using the normal Coulomb potential
365 will give,
366 \begin{equation}
367 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}.
368 \label{eq:SFPot}
369 \end{equation}
370 with associated forces,
371 \begin{equation}
372 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}.
373 \label{eq:SFForces}
374 \end{equation}
375 This formulation has the benefits that there are no discontinuities at
376 the cutoff radius, while the neutralizing image charges are present in
377 both the energy and force expressions. It would be simple to add the
378 self-neutralizing term back when computing the total energy of the
379 system, thereby maintaining the agreement with the Madelung energies.
380 A side effect of this treatment is the alteration in the shape of the
381 potential that comes from the derivative term. Thus, a degree of
382 clarity about agreement with the empirical potential is lost in order
383 to gain functionality in dynamics simulations.
384
385 Wolf \textit{et al.} originally discussed the energetics of the
386 shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
387 insufficient for accurate determination of the energy with reasonable
388 cutoff distances. The calculated Madelung energies fluctuated around
389 the expected value as the cutoff radius was increased, but the
390 oscillations converged toward the correct value.\cite{Wolf99} A
391 damping function was incorporated to accelerate the convergence; and
392 though alternative forms for the damping function could be
393 used,\cite{Jones56,Heyes81} the complimentary error function was
394 chosen to mirror the effective screening used in the Ewald summation.
395 Incorporating this error function damping into the simple Coulomb
396 potential,
397 \begin{equation}
398 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
399 \label{eq:dampCoulomb}
400 \end{equation}
401 the shifted potential (eq. (\ref{eq:SPPot})) becomes
402 \begin{equation}
403 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},
404 \label{eq:DSPPot}
405 \end{equation}
406 with associated forces,
407 \begin{equation}
408 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}.
409 \label{eq:DSPForces}
410 \end{equation}
411 Again, this damped shifted potential suffers from a
412 force-discontinuity at the cutoff radius, and the image charges play
413 no role in the forces. To remedy these concerns, one may derive a
414 {\sc sf} variant by including the derivative term in
415 eq. (\ref{eq:shiftingForm}),
416 \begin{equation}
417 \begin{split}
418 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}.
419 \label{eq:DSFPot}
420 \end{split}
421 \end{equation}
422 The derivative of the above potential will lead to the following forces,
423 \begin{equation}
424 \begin{split}
425 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}.
426 \label{eq:DSFForces}
427 \end{split}
428 \end{equation}
429 If the damping parameter $(\alpha)$ is set to zero, the undamped case,
430 eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
431 recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
432
433 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
434 derived by Zahn \textit{et al.}; however, there are two important
435 differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
436 eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
437 with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
438 in the Zahn potential, resulting in a potential discontinuity as
439 particles cross $R_\textrm{c}$. Second, the sign of the derivative
440 portion is different. The missing $v_\textrm{c}$ term would not
441 affect molecular dynamics simulations (although the computed energy
442 would be expected to have sudden jumps as particle distances crossed
443 $R_c$). The sign problem is a potential source of errors, however.
444 In fact, it introduces a discontinuity in the forces at the cutoff,
445 because the force function is shifted in the wrong direction and
446 doesn't cross zero at $R_\textrm{c}$.
447
448 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
449 electrostatic summation method in which the potential and forces are
450 continuous at the cutoff radius and which incorporates the damping
451 function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
452 this paper, we will evaluate exactly how good these methods ({\sc sp},
453 {\sc sf}, damping) are at reproducing the correct electrostatic
454 summation performed by the Ewald sum.
455
456 \subsection{Other alternatives}
457 In addition to the methods described above, we considered some other
458 techniques that are commonly used in molecular simulations. The
459 simplest of these is group-based cutoffs. Though of little use for
460 charged molecules, collecting atoms into neutral groups takes
461 advantage of the observation that the electrostatic interactions decay
462 faster than those for monopolar pairs.\cite{Steinbach94} When
463 considering these molecules as neutral groups, the relative
464 orientations of the molecules control the strength of the interactions
465 at the cutoff radius. Consequently, as these molecular particles move
466 through $R_\textrm{c}$, the energy will drift upward due to the
467 anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
468 maintain good energy conservation, both the potential and derivative
469 need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
470 This is accomplished using a standard switching function. If a smooth
471 second derivative is desired, a fifth (or higher) order polynomial can
472 be used.\cite{Andrea83}
473
474 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
475 and to incorporate the effects of the surroundings, a method like
476 Reaction Field ({\sc rf}) can be used. The original theory for {\sc
477 rf} was originally developed by Onsager,\cite{Onsager36} and it was
478 applied in simulations for the study of water by Barker and
479 Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
480 an extension of the group-based cutoff method where the net dipole
481 within the cutoff sphere polarizes an external dielectric, which
482 reacts back on the central dipole. The same switching function
483 considerations for group-based cutoffs need to made for {\sc rf}, with
484 the additional pre-specification of a dielectric constant.
485
486 \section{Methods}
487
488 In classical molecular mechanics simulations, there are two primary
489 techniques utilized to obtain information about the system of
490 interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
491 techniques utilize pairwise summations of interactions between
492 particle sites, but they use these summations in different ways.
493
494 In MC, the potential energy difference between configurations dictates
495 the progression of MC sampling. Going back to the origins of this
496 method, the acceptance criterion for the canonical ensemble laid out
497 by Metropolis \textit{et al.} states that a subsequent configuration
498 is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
499 $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
500 Maintaining the correct $\Delta E$ when using an alternate method for
501 handling the long-range electrostatics will ensure proper sampling
502 from the ensemble.
503
504 In MD, the derivative of the potential governs how the system will
505 progress in time. Consequently, the force and torque vectors on each
506 body in the system dictate how the system evolves. If the magnitude
507 and direction of these vectors are similar when using alternate
508 electrostatic summation techniques, the dynamics in the short term
509 will be indistinguishable. Because error in MD calculations is
510 cumulative, one should expect greater deviation at longer times,
511 although methods which have large differences in the force and torque
512 vectors will diverge from each other more rapidly.
513
514 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
515
516 The pairwise summation techniques (outlined in section
517 \ref{sec:ESMethods}) were evaluated for use in MC simulations by
518 studying the energy differences between conformations. We took the
519 {\sc spme}-computed energy difference between two conformations to be the
520 correct behavior. An ideal performance by an alternative method would
521 reproduce these energy differences exactly (even if the absolute
522 energies calculated by the methods are different). Since none of the
523 methods provide exact energy differences, we used linear least squares
524 regressions of energy gap data to evaluate how closely the methods
525 mimicked the Ewald energy gaps. Unitary results for both the
526 correlation (slope) and correlation coefficient for these regressions
527 indicate perfect agreement between the alternative method and {\sc spme}.
528 Sample correlation plots for two alternate methods are shown in
529 Fig. \ref{fig:linearFit}.
530
531 \begin{figure}
532 \centering
533 \includegraphics[width = \linewidth]{./dualLinear.pdf}
534 \caption{Example least squares regressions of the configuration energy
535 differences for SPC/E water systems. The upper plot shows a data set
536 with a poor correlation coefficient ($R^2$), while the lower plot
537 shows a data set with a good correlation coefficient.}
538 \label{fig:linearFit}
539 \end{figure}
540
541 Each system type (detailed in section \ref{sec:RepSims}) was
542 represented using 500 independent configurations. Additionally, we
543 used seven different system types, so each of the alternative
544 (non-Ewald) electrostatic summation methods was evaluated using
545 873,250 configurational energy differences.
546
547 Results and discussion for the individual analysis of each of the
548 system types appear in the supporting information, while the
549 cumulative results over all the investigated systems appears below in
550 section \ref{sec:EnergyResults}.
551
552 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
553 We evaluated the pairwise methods (outlined in section
554 \ref{sec:ESMethods}) for use in MD simulations by
555 comparing the force and torque vectors with those obtained using the
556 reference Ewald summation ({\sc spme}). Both the magnitude and the
557 direction of these vectors on each of the bodies in the system were
558 analyzed. For the magnitude of these vectors, linear least squares
559 regression analyses were performed as described previously for
560 comparing $\Delta E$ values. Instead of a single energy difference
561 between two system configurations, we compared the magnitudes of the
562 forces (and torques) on each molecule in each configuration. For a
563 system of 1000 water molecules and 40 ions, there are 1040 force
564 vectors and 1000 torque vectors. With 500 configurations, this
565 results in 520,000 force and 500,000 torque vector comparisons.
566 Additionally, data from seven different system types was aggregated
567 before the comparison was made.
568
569 The {\it directionality} of the force and torque vectors was
570 investigated through measurement of the angle ($\theta$) formed
571 between those computed from the particular method and those from {\sc spme},
572 \begin{equation}
573 \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
574 \end{equation}
575 where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
576 vector computed using method M. Each of these $\theta$ values was
577 accumulated in a distribution function and weighted by the area on the
578 unit sphere. Since this distribution is a measure of angular error
579 between two different electrostatic summation methods, there is no
580 {\it a priori} reason for the profile to adhere to any specific
581 shape. Thus, gaussian fits were used to measure the width of the
582 resulting distributions. The variance ($\sigma^2$) was extracted from
583 each of these fits and was used to compare distribution widths.
584 Values of $\sigma^2$ near zero indicate vector directions
585 indistinguishable from those calculated when using the reference
586 method ({\sc spme}).
587
588 \subsection{Short-time Dynamics}
589
590 The effects of the alternative electrostatic summation methods on the
591 short-time dynamics of charged systems were evaluated by considering a
592 NaCl crystal at a temperature of 1000 K. A subset of the best
593 performing pairwise methods was used in this comparison. The NaCl
594 crystal was chosen to avoid possible complications from the treatment
595 of orientational motion in molecular systems. All systems were
596 started with the same initial positions and velocities. Simulations
597 were performed under the microcanonical ensemble, and velocity
598 autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
599 of the trajectories,
600 \begin{equation}
601 C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
602 \label{eq:vCorr}
603 \end{equation}
604 Velocity autocorrelation functions require detailed short time data,
605 thus velocity information was saved every 2 fs over 10 ps
606 trajectories. Because the NaCl crystal is composed of two different
607 atom types, the average of the two resulting velocity autocorrelation
608 functions was used for comparisons.
609
610 \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
611
612 The effects of the same subset of alternative electrostatic methods on
613 the {\it long-time} dynamics of charged systems were evaluated using
614 the same model system (NaCl crystals at 1000~K). The power spectrum
615 ($I(\omega)$) was obtained via Fourier transform of the velocity
616 autocorrelation function, \begin{equation} I(\omega) =
617 \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
618 \label{eq:powerSpec}
619 \end{equation}
620 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
621 NaCl crystal is composed of two different atom types, the average of
622 the two resulting power spectra was used for comparisons. Simulations
623 were performed under the microcanonical ensemble, and velocity
624 information was saved every 5 fs over 100 ps trajectories.
625
626 \subsection{Representative Simulations}\label{sec:RepSims}
627 A variety of representative simulations were analyzed to determine the
628 relative effectiveness of the pairwise summation techniques in
629 reproducing the energetics and dynamics exhibited by {\sc spme}. We wanted
630 to span the space of modern simulations (i.e. from liquids of neutral
631 molecules to ionic crystals), so the systems studied were:
632 \begin{enumerate}
633 \item liquid water (SPC/E),\cite{Berendsen87}
634 \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
635 \item NaCl crystals,
636 \item NaCl melts,
637 \item a low ionic strength solution of NaCl in water (0.11 M),
638 \item a high ionic strength solution of NaCl in water (1.1 M), and
639 \item a 6 \AA\ radius sphere of Argon in water.
640 \end{enumerate}
641 By utilizing the pairwise techniques (outlined in section
642 \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
643 charged particles, and mixtures of the two, we hope to discern under
644 which conditions it will be possible to use one of the alternative
645 summation methodologies instead of the Ewald sum.
646
647 For the solid and liquid water configurations, configurations were
648 taken at regular intervals from high temperature trajectories of 1000
649 SPC/E water molecules. Each configuration was equilibrated
650 independently at a lower temperature (300~K for the liquid, 200~K for
651 the crystal). The solid and liquid NaCl systems consisted of 500
652 $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
653 these systems were selected and equilibrated in the same manner as the
654 water systems. In order to introduce measurable fluctuations in the
655 configuration energy differences, the crystalline simulations were
656 equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid
657 NaCl configurations needed to represent a fully disordered array of
658 point charges, so the high temperature of 7000~K was selected for
659 equilibration. The ionic solutions were made by solvating 4 (or 40)
660 ions in a periodic box containing 1000 SPC/E water molecules. Ion and
661 water positions were then randomly swapped, and the resulting
662 configurations were again equilibrated individually. Finally, for the
663 Argon / Water ``charge void'' systems, the identities of all the SPC/E
664 waters within 6 \AA\ of the center of the equilibrated water
665 configurations were converted to argon.
666
667 These procedures guaranteed us a set of representative configurations
668 from chemically-relevant systems sampled from appropriate
669 ensembles. Force field parameters for the ions and Argon were taken
670 from the force field utilized by {\sc oopse}.\cite{Meineke05}
671
672 \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
673 We compared the following alternative summation methods with results
674 from the reference method ({\sc spme}):
675 \begin{itemize}
676 \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
677 and 0.3 \AA$^{-1}$,
678 \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
679 and 0.3 \AA$^{-1}$,
680 \item reaction field with an infinite dielectric constant, and
681 \item an unmodified cutoff.
682 \end{itemize}
683 Group-based cutoffs with a fifth-order polynomial switching function
684 were utilized for the reaction field simulations. Additionally, we
685 investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
686 cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
687 implementation of {\sc spme},\cite{Ponder87} while all other calculations
688 were performed using the {\sc oopse} molecular mechanics
689 package.\cite{Meineke05} All other portions of the energy calculation
690 (i.e. Lennard-Jones interactions) were handled in exactly the same
691 manner across all systems and configurations.
692
693 The alternative methods were also evaluated with three different
694 cutoff radii (9, 12, and 15 \AA). As noted previously, the
695 convergence parameter ($\alpha$) plays a role in the balance of the
696 real-space and reciprocal-space portions of the Ewald calculation.
697 Typical molecular mechanics packages set this to a value dependent on
698 the cutoff radius and a tolerance (typically less than $1 \times
699 10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
700 increasing accuracy at the expense of computational time spent on the
701 reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
702 The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
703 in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
704 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
705 respectively.
706
707 \section{Results and Discussion}
708
709 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
710 In order to evaluate the performance of the pairwise electrostatic
711 summation methods for Monte Carlo simulations, the energy differences
712 between configurations were compared to the values obtained when using
713 {\sc spme}. The results for the subsequent regression analysis are shown in
714 figure \ref{fig:delE}.
715
716 \begin{figure}
717 \centering
718 \includegraphics[width=5.5in]{./delEplot.pdf}
719 \caption{Statistical analysis of the quality of configurational energy
720 differences for a given electrostatic method compared with the
721 reference Ewald sum. Results with a value equal to 1 (dashed line)
722 indicate $\Delta E$ values indistinguishable from those obtained using
723 {\sc spme}. Different values of the cutoff radius are indicated with
724 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
725 inverted triangles).}
726 \label{fig:delE}
727 \end{figure}
728
729 The most striking feature of this plot is how well the Shifted Force
730 ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
731 differences. For the undamped {\sc sf} method, and the
732 moderately-damped {\sc sp} methods, the results are nearly
733 indistinguishable from the Ewald results. The other common methods do
734 significantly less well.
735
736 The unmodified cutoff method is essentially unusable. This is not
737 surprising since hard cutoffs give large energy fluctuations as atoms
738 or molecules move in and out of the cutoff
739 radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
740 some degree by using group based cutoffs with a switching
741 function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
742 significant improvement using the group-switched cutoff because the
743 salt and salt solution systems contain non-neutral groups. Interested
744 readers can consult the accompanying supporting information for a
745 comparison where all groups are neutral.
746
747 For the {\sc sp} method, inclusion of electrostatic damping improves
748 the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
749 shows an excellent correlation and quality of fit with the {\sc spme}
750 results, particularly with a cutoff radius greater than 12
751 \AA . Use of a larger damping parameter is more helpful for the
752 shortest cutoff shown, but it has a detrimental effect on simulations
753 with larger cutoffs.
754
755 In the {\sc sf} sets, increasing damping results in progressively {\it
756 worse} correlation with Ewald. Overall, the undamped case is the best
757 performing set, as the correlation and quality of fits are
758 consistently superior regardless of the cutoff distance. The undamped
759 case is also less computationally demanding (because no evaluation of
760 the complementary error function is required).
761
762 The reaction field results illustrates some of that method's
763 limitations, primarily that it was developed for use in homogenous
764 systems; although it does provide results that are an improvement over
765 those from an unmodified cutoff.
766
767 \subsection{Magnitudes of the Force and Torque Vectors}
768
769 Evaluation of pairwise methods for use in Molecular Dynamics
770 simulations requires consideration of effects on the forces and
771 torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
772 regression results for the force and torque vector magnitudes,
773 respectively. The data in these figures was generated from an
774 accumulation of the statistics from all of the system types.
775
776 \begin{figure}
777 \centering
778 \includegraphics[width=5.5in]{./frcMagplot.pdf}
779 \caption{Statistical analysis of the quality of the force vector
780 magnitudes for a given electrostatic method compared with the
781 reference Ewald sum. Results with a value equal to 1 (dashed line)
782 indicate force magnitude values indistinguishable from those obtained
783 using {\sc spme}. Different values of the cutoff radius are indicated with
784 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
785 inverted triangles).}
786 \label{fig:frcMag}
787 \end{figure}
788
789 Again, it is striking how well the Shifted Potential and Shifted Force
790 methods are doing at reproducing the {\sc spme} forces. The undamped and
791 weakly-damped {\sc sf} method gives the best agreement with Ewald.
792 This is perhaps expected because this method explicitly incorporates a
793 smooth transition in the forces at the cutoff radius as well as the
794 neutralizing image charges.
795
796 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
797 in the previous $\Delta E$ section. The unmodified cutoff results are
798 poor, but using group based cutoffs and a switching function provides
799 an improvement much more significant than what was seen with $\Delta
800 E$.
801
802 With moderate damping and a large enough cutoff radius, the {\sc sp}
803 method is generating usable forces. Further increases in damping,
804 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
805 detrimental to simulations with larger cutoff radii.
806
807 The reaction field results are surprisingly good, considering the poor
808 quality of the fits for the $\Delta E$ results. There is still a
809 considerable degree of scatter in the data, but the forces correlate
810 well with the Ewald forces in general. We note that the reaction
811 field calculations do not include the pure NaCl systems, so these
812 results are partly biased towards conditions in which the method
813 performs more favorably.
814
815 \begin{figure}
816 \centering
817 \includegraphics[width=5.5in]{./trqMagplot.pdf}
818 \caption{Statistical analysis of the quality of the torque vector
819 magnitudes for a given electrostatic method compared with the
820 reference Ewald sum. Results with a value equal to 1 (dashed line)
821 indicate torque magnitude values indistinguishable from those obtained
822 using {\sc spme}. Different values of the cutoff radius are indicated with
823 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
824 inverted triangles).}
825 \label{fig:trqMag}
826 \end{figure}
827
828 Molecular torques were only available from the systems which contained
829 rigid molecules (i.e. the systems containing water). The data in
830 fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
831
832 Torques appear to be much more sensitive to charges at a longer
833 distance. The striking feature in comparing the new electrostatic
834 methods with {\sc spme} is how much the agreement improves with increasing
835 cutoff radius. Again, the weakly damped and undamped {\sc sf} method
836 appears to be reproducing the {\sc spme} torques most accurately.
837
838 Water molecules are dipolar, and the reaction field method reproduces
839 the effect of the surrounding polarized medium on each of the
840 molecular bodies. Therefore it is not surprising that reaction field
841 performs best of all of the methods on molecular torques.
842
843 \subsection{Directionality of the Force and Torque Vectors}
844
845 It is clearly important that a new electrostatic method can reproduce
846 the magnitudes of the force and torque vectors obtained via the Ewald
847 sum. However, the {\it directionality} of these vectors will also be
848 vital in calculating dynamical quantities accurately. Force and
849 torque directionalities were investigated by measuring the angles
850 formed between these vectors and the same vectors calculated using
851 {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
852 variance ($\sigma^2$) of the Gaussian fits of the angle error
853 distributions of the combined set over all system types.
854
855 \begin{figure}
856 \centering
857 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
858 \caption{Statistical analysis of the width of the angular distribution
859 that the force and torque vectors from a given electrostatic method
860 make with their counterparts obtained using the reference Ewald sum.
861 Results with a variance ($\sigma^2$) equal to zero (dashed line)
862 indicate force and torque directions indistinguishable from those
863 obtained using {\sc spme}. Different values of the cutoff radius are
864 indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
865 and 15\AA\ = inverted triangles).}
866 \label{fig:frcTrqAng}
867 \end{figure}
868
869 Both the force and torque $\sigma^2$ results from the analysis of the
870 total accumulated system data are tabulated in figure
871 \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
872 sp}) method would be essentially unusable for molecular dynamics
873 unless the damping function is added. The Shifted Force ({\sc sf})
874 method, however, is generating force and torque vectors which are
875 within a few degrees of the Ewald results even with weak (or no)
876 damping.
877
878 All of the sets (aside from the over-damped case) show the improvement
879 afforded by choosing a larger cutoff radius. Increasing the cutoff
880 from 9 to 12 \AA\ typically results in a halving of the width of the
881 distribution, with a similar improvement when going from 12 to 15
882 \AA .
883
884 The undamped {\sc sf}, group-based cutoff, and reaction field methods
885 all do equivalently well at capturing the direction of both the force
886 and torque vectors. Using the electrostatic damping improves the
887 angular behavior significantly for the {\sc sp} and moderately for the
888 {\sc sf} methods. Overdamping is detrimental to both methods. Again
889 it is important to recognize that the force vectors cover all
890 particles in all seven systems, while torque vectors are only
891 available for neutral molecular groups. Damping is more beneficial to
892 charged bodies, and this observation is investigated further in the
893 accompanying supporting information.
894
895 Although not discussed previously, group based cutoffs can be applied
896 to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
897 will reintroduce small discontinuities at the cutoff radius, but the
898 effects of these can be minimized by utilizing a switching function.
899 Though there are no significant benefits or drawbacks observed in
900 $\Delta E$ and the force and torque magnitudes when doing this, there
901 is a measurable improvement in the directionality of the forces and
902 torques. Table \ref{tab:groupAngle} shows the angular variances
903 obtained using group based cutoffs along with the results seen in
904 figure \ref{fig:frcTrqAng}. The {\sc sp} (with an $\alpha$ of 0.2
905 \AA$^{-1}$ or smaller) shows much narrower angular distributions when
906 using group-based cutoffs. The {\sc sf} method likewise shows
907 improvement in the undamped and lightly damped cases.
908
909 \begin{table}[htbp]
910 \centering
911 \caption{Statistical analysis of the angular
912 distributions that the force (upper) and torque (lower) vectors
913 from a given electrostatic method make with their counterparts
914 obtained using the reference Ewald sum. Calculations were
915 performed both with (Y) and without (N) group based cutoffs and a
916 switching function. The $\alpha$ values have units of \AA$^{-1}$
917 and the variance values have units of degrees$^2$.}
918
919 \begin{tabular}{@{} ccrrrrrrrr @{}}
920 \\
921 \toprule
922 & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
923 \cmidrule(lr){3-6}
924 \cmidrule(l){7-10}
925 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
926 \midrule
927
928 9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
929 & \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} \\
930 12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
931 & \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} \\
932 15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
933 & \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} \\
934
935 \midrule
936
937 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
938 & \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} \\
939 12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
940 & \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} \\
941 15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
942 & \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} \\
943
944 \bottomrule
945 \end{tabular}
946 \label{tab:groupAngle}
947 \end{table}
948
949 One additional trend in table \ref{tab:groupAngle} is that the
950 $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
951 increases, something that is more obvious with group-based cutoffs.
952 The complimentary error function inserted into the potential weakens
953 the electrostatic interaction as the value of $\alpha$ is increased.
954 However, at larger values of $\alpha$, it is possible to overdamp the
955 electrostatic interaction and to remove it completely. Kast
956 \textit{et al.} developed a method for choosing appropriate $\alpha$
957 values for these types of electrostatic summation methods by fitting
958 to $g(r)$ data, and their methods indicate optimal values of 0.34,
959 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
960 respectively.\cite{Kast03} These appear to be reasonable choices to
961 obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
962 these findings, choices this high would introduce error in the
963 molecular torques, particularly for the shorter cutoffs. Based on our
964 observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
965 but damping may be unnecessary when using the {\sc sf} method.
966
967 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
968
969 Zahn {\it et al.} investigated the structure and dynamics of water
970 using eqs. (\ref{eq:ZahnPot}) and
971 (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
972 that a method similar (but not identical with) the damped {\sc sf}
973 method resulted in properties very similar to those obtained when
974 using the Ewald summation. The properties they studied (pair
975 distribution functions, diffusion constants, and velocity and
976 orientational correlation functions) may not be particularly sensitive
977 to the long-range and collective behavior that governs the
978 low-frequency behavior in crystalline systems. Additionally, the
979 ionic crystals are the worst case scenario for the pairwise methods
980 because they lack the reciprocal space contribution contained in the
981 Ewald summation.
982
983 We are using two separate measures to probe the effects of these
984 alternative electrostatic methods on the dynamics in crystalline
985 materials. For short- and intermediate-time dynamics, we are
986 computing the velocity autocorrelation function, and for long-time
987 and large length-scale collective motions, we are looking at the
988 low-frequency portion of the power spectrum.
989
990 \begin{figure}
991 \centering
992 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
993 \caption{Velocity autocorrelation functions of NaCl crystals at
994 1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
995 sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
996 the first minimum. The times to first collision are nearly identical,
997 but differences can be seen in the peaks and troughs, where the
998 undamped and weakly damped methods are stiffer than the moderately
999 damped and {\sc spme} methods.}
1000 \label{fig:vCorrPlot}
1001 \end{figure}
1002
1003 The short-time decay of the velocity autocorrelation function through
1004 the first collision are nearly identical in figure
1005 \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1006 how the methods differ. The undamped {\sc sf} method has deeper
1007 troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1008 any of the other methods. As the damping parameter ($\alpha$) is
1009 increased, these peaks are smoothed out, and the {\sc sf} method
1010 approaches the {\sc spme} results. With $\alpha$ values of 0.2 \AA$^{-1}$,
1011 the {\sc sf} and {\sc sp} functions are nearly identical and track the
1012 {\sc spme} features quite well. This is not surprising because the {\sc sf}
1013 and {\sc sp} potentials become nearly identical with increased
1014 damping. However, this appears to indicate that once damping is
1015 utilized, the details of the form of the potential (and forces)
1016 constructed out of the damped electrostatic interaction are less
1017 important.
1018
1019 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1020
1021 To evaluate how the differences between the methods affect the
1022 collective long-time motion, we computed power spectra from long-time
1023 traces of the velocity autocorrelation function. The power spectra for
1024 the best-performing alternative methods are shown in
1025 fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1026 a cubic switching function between 40 and 50 ps was used to reduce the
1027 ringing resulting from data truncation. This procedure had no
1028 noticeable effect on peak location or magnitude.
1029
1030 \begin{figure}
1031 \centering
1032 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1033 \caption{Power spectra obtained from the velocity auto-correlation
1034 functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf}
1035 ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset
1036 shows the frequency region below 100 cm$^{-1}$ to highlight where the
1037 spectra differ.}
1038 \label{fig:methodPS}
1039 \end{figure}
1040
1041 While the high frequency regions of the power spectra for the
1042 alternative methods are quantitatively identical with Ewald spectrum,
1043 the low frequency region shows how the summation methods differ.
1044 Considering the low-frequency inset (expanded in the upper frame of
1045 figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1046 correlated motions are blue-shifted when using undamped or weakly
1047 damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1048 \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1049 correlated motion to the Ewald method (which has a convergence
1050 parameter of 0.3119 \AA$^{-1}$). This weakening of the electrostatic
1051 interaction with increased damping explains why the long-ranged
1052 correlated motions are at lower frequencies for the moderately damped
1053 methods than for undamped or weakly damped methods.
1054
1055 To isolate the role of the damping constant, we have computed the
1056 spectra for a single method ({\sc sf}) with a range of damping
1057 constants and compared this with the {\sc spme} spectrum.
1058 Fig. \ref{fig:dampInc} shows more clearly that increasing the
1059 electrostatic damping red-shifts the lowest frequency phonon modes.
1060 However, even without any electrostatic damping, the {\sc sf} method
1061 has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1062 Without the {\sc sf} modifications, an undamped (pure cutoff) method
1063 would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1064 Most} of the collective behavior in the crystal is accurately captured
1065 using the {\sc sf} method. Quantitative agreement with Ewald can be
1066 obtained using moderate damping in addition to the shifting at the
1067 cutoff distance.
1068
1069 \begin{figure}
1070 \centering
1071 \includegraphics[width = \linewidth]{./increasedDamping.pdf}
1072 \caption{Effect of damping on the two lowest-frequency phonon modes in
1073 the NaCl crystal at 1000~K. The undamped shifted force ({\sc sf})
1074 method is off by less than 10 cm$^{-1}$, and increasing the
1075 electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1076 with the power spectrum obtained using the Ewald sum. Overdamping can
1077 result in underestimates of frequencies of the long-wavelength
1078 motions.}
1079 \label{fig:dampInc}
1080 \end{figure}
1081
1082 \section{Conclusions}
1083
1084 This investigation of pairwise electrostatic summation techniques
1085 shows that there are viable and computationally efficient alternatives
1086 to the Ewald summation. These methods are derived from the damped and
1087 cutoff-neutralized Coulombic sum originally proposed by Wolf
1088 \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1089 method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1090 (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1091 energetic and dynamic characteristics exhibited by simulations
1092 employing lattice summation techniques. The cumulative energy
1093 difference results showed the undamped {\sc sf} and moderately damped
1094 {\sc sp} methods produced results nearly identical to {\sc spme}. Similarly
1095 for the dynamic features, the undamped or moderately damped {\sc sf}
1096 and moderately damped {\sc sp} methods produce force and torque vector
1097 magnitude and directions very similar to the expected values. These
1098 results translate into long-time dynamic behavior equivalent to that
1099 produced in simulations using {\sc spme}.
1100
1101 As in all purely-pairwise cutoff methods, these methods are expected
1102 to scale approximately {\it linearly} with system size, and they are
1103 easily parallelizable. This should result in substantial reductions
1104 in the computational cost of performing large simulations.
1105
1106 Aside from the computational cost benefit, these techniques have
1107 applicability in situations where the use of the Ewald sum can prove
1108 problematic. Of greatest interest is their potential use in
1109 interfacial systems, where the unmodified lattice sum techniques
1110 artificially accentuate the periodicity of the system in an
1111 undesirable manner. There have been alterations to the standard Ewald
1112 techniques, via corrections and reformulations, to compensate for
1113 these systems; but the pairwise techniques discussed here require no
1114 modifications, making them natural tools to tackle these problems.
1115 Additionally, this transferability gives them benefits over other
1116 pairwise methods, like reaction field, because estimations of physical
1117 properties (e.g. the dielectric constant) are unnecessary.
1118
1119 If a researcher is using Monte Carlo simulations of large chemical
1120 systems containing point charges, most structural features will be
1121 accurately captured using the undamped {\sc sf} method or the {\sc sp}
1122 method with an electrostatic damping of 0.2 \AA$^{-1}$. These methods
1123 would also be appropriate for molecular dynamics simulations where the
1124 data of interest is either structural or short-time dynamical
1125 quantities. For long-time dynamics and collective motions, the safest
1126 pairwise method we have evaluated is the {\sc sf} method with an
1127 electrostatic damping between 0.2 and 0.25
1128 \AA$^{-1}$.
1129
1130 We are not suggesting that there is any flaw with the Ewald sum; in
1131 fact, it is the standard by which these simple pairwise sums have been
1132 judged. However, these results do suggest that in the typical
1133 simulations performed today, the Ewald summation may no longer be
1134 required to obtain the level of accuracy most researchers have come to
1135 expect.
1136
1137 \section{Acknowledgments}
1138 Support for this project was provided by the National Science
1139 Foundation under grant CHE-0134881. The authors would like to thank
1140 Steve Corcelli and Ed Maginn for helpful discussions and comments.
1141
1142 \newpage
1143
1144 \bibliographystyle{jcp2}
1145 \bibliography{electrostaticMethods}
1146
1147
1148 \end{document}