ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2653
Committed: Tue Mar 21 20:46:55 2006 UTC (18 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 61439 byte(s)
Log Message:
results

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 necessary? Pairwise alternatives to the accepted standard for long-range electrostatics}
29
30 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
31 gezelter@nd.edu} \\
32 Department of Chemistry and Biochemistry\\
33 University of Notre Dame\\
34 Notre Dame, Indiana 46556}
35
36 \date{\today}
37
38 \maketitle
39 \doublespacing
40
41 \nobibliography{}
42 \begin{abstract}
43 A new method for accumulating electrostatic interactions was derived
44 from the previous efforts described in \bibentry{Wolf99} and
45 \bibentry{Zahn02} as a possible replacement for lattice sum methods in
46 molecular simulations. Comparisons were performed with this and other
47 pairwise electrostatic summation techniques against the smooth
48 particle mesh Ewald (SPME) summation to see how well they reproduce
49 the energetics and dynamics of a variety of simulation types. The
50 newly derived Shifted-Force technique shows a remarkable ability to
51 reproduce the behavior exhibited in simulations using SPME with an
52 $\mathscr{O}(N)$ computational cost, equivalent to merely the
53 real-space portion of the lattice summation.
54
55 \end{abstract}
56
57 \newpage
58
59 %\narrowtext
60
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 % BODY OF TEXT
63 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64
65 \section{Introduction}
66
67 In molecular simulations, proper accumulation of the electrostatic
68 interactions is essential and is one of the most
69 computationally-demanding tasks. The common molecular mechanics force
70 fields represent atomic sites with full or partial charges protected
71 by Lennard-Jones (short range) interactions. This means that nearly
72 every pair interaction involves a calculation of charge-charge forces.
73 Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
74 interactions quickly become the most expensive part of molecular
75 simulations. Historically, the electrostatic pair interaction would
76 not have decayed appreciably within the typical box lengths that could
77 be feasibly simulated. In the larger systems that are more typical of
78 modern simulations, large cutoffs should be used to incorporate
79 electrostatics correctly.
80
81 There have been many efforts to address the proper and practical
82 handling of electrostatic interactions, and these have resulted in a
83 variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
84 typically classified as implicit methods (i.e., continuum dielectrics,
85 static dipolar fields),\cite{Born20,Grossfield00} explicit methods
86 (i.e., Ewald summations, interaction shifting or
87 truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
88 reaction field type methods, fast multipole
89 methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
90 often preferred because they physically incorporate solvent molecules
91 in the system of interest, but these methods are sometimes difficult
92 to utilize because of their high computational cost.\cite{Roux99} In
93 addition to the computational cost, there have been some questions
94 regarding possible artifacts caused by the inherent periodicity of the
95 explicit Ewald summation.\cite{Tobias01}
96
97 In this paper, we focus on a new set of shifted methods devised by
98 Wolf {\it et al.},\cite{Wolf99} which we further extend. These
99 methods along with a few other mixed methods (i.e. reaction field) are
100 compared with the smooth particle mesh Ewald
101 sum,\cite{Onsager36,Essmann99} which is our reference method for
102 handling long-range electrostatic interactions. The new methods for
103 handling electrostatics have the potential to scale linearly with
104 increasing system size since they involve only a simple modification
105 to the direct pairwise sum. They also lack the added periodicity of
106 the Ewald sum, so they can be used for systems which are non-periodic
107 or which have one- or two-dimensional periodicity. Below, these
108 methods are evaluated using a variety of model systems to establish
109 their usability in molecular simulations.
110
111 \subsection{The Ewald Sum}
112 The complete accumulation electrostatic interactions in a system with
113 periodic boundary conditions (PBC) requires the consideration of the
114 effect of all charges within a (cubic) simulation box as well as those
115 in the periodic replicas,
116 \begin{equation}
117 V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime \left[ \sum_{i=1}^N\sum_{j=1}^N \phi\left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right) \right],
118 \label{eq:PBCSum}
119 \end{equation}
120 where the sum over $\mathbf{n}$ is a sum over all periodic box
121 replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
122 prime indicates $i = j$ are neglected for $\mathbf{n} =
123 0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
124 particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
125 the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
126 $j$, and $\phi$ is the solution to Poisson's equation
127 ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
128 charge-charge interactions). In the case of monopole electrostatics,
129 eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
130 non-neutral systems.
131
132 The electrostatic summation problem was originally studied by Ewald
133 for the case of an infinite crystal.\cite{Ewald21}. The approach he
134 took was to convert this conditionally convergent sum into two
135 absolutely convergent summations: a short-ranged real-space summation
136 and a long-ranged reciprocal-space summation,
137 \begin{equation}
138 \begin{split}
139 V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}{|\mathbf{r}_{ij}+\mathbf{n}|} \\ &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
140 \end{split}
141 \label{eq:EwaldSum}
142 \end{equation}
143 where $\alpha$ is the damping or convergence parameter with units of
144 \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
145 $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
146 constant of the surrounding medium. The final two terms of
147 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
148 for interacting with a surrounding dielectric.\cite{Allen87} This
149 dipolar term was neglected in early applications in molecular
150 simulations,\cite{Brush66,Woodcock71} until it was introduced by de
151 Leeuw {\it et al.} to address situations where the unit cell has a
152 dipole moment which is magnified through replication of the periodic
153 images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
154 system is said to be using conducting (or ``tin-foil'') boundary
155 conditions, $\epsilon_{\rm S} = \infty$. Figure
156 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
157 time. Initially, due to the small system sizes that could be
158 simulated feasibly, the entire simulation box was replicated to
159 convergence. In more modern simulations, the systems have grown large
160 enough that a real-space cutoff could potentially give convergent
161 behavior. Indeed, it has been observed that with the choice of a
162 small $\alpha$, the reciprocal-space portion of the Ewald sum can be
163 rapidly convergent and small relative to the real-space
164 portion.\cite{Karasawa89,Kolafa92}
165
166 \begin{figure}
167 \centering
168 \includegraphics[width = \linewidth]{./ewaldProgression2.pdf}
169 \caption{The change in the application of the Ewald sum with
170 increasing computational power. Initially, only small systems could
171 be studied, and the Ewald sum replicated the simulation box to
172 convergence. Now, much larger systems of charges are investigated
173 with fixed-distance cutoffs.}
174 \label{fig:ewaldTime}
175 \end{figure}
176
177 The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
178 convergence parameter $(\alpha)$ plays an important role in balancing
179 the computational cost between the direct and reciprocal-space
180 portions of the summation. The choice of this value allows one to
181 select whether the real-space or reciprocal space portion of the
182 summation is an $\mathscr{O}(N^2)$ calculation (with the other being
183 $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
184 $\alpha$ and thoughtful algorithm development, this cost can be
185 reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
186 taken to reduce the cost of the Ewald summation even further is to set
187 $\alpha$ such that the real-space interactions decay rapidly, allowing
188 for a short spherical cutoff. Then the reciprocal space summation is
189 optimized. These optimizations usually involve utilization of the
190 fast Fourier transform (FFT),\cite{Hockney81} leading to the
191 particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
192 methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
193 methods, the cost of the reciprocal-space portion of the Ewald
194 summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
195 \log N)$.
196
197 These developments and optimizations have made the use of the Ewald
198 summation routine in simulations with periodic boundary
199 conditions. However, in certain systems, such as vapor-liquid
200 interfaces and membranes, the intrinsic three-dimensional periodicity
201 can prove problematic. The Ewald sum has been reformulated to handle
202 2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
203 new methods are computationally expensive.\cite{Spohr97,Yeh99}
204 Inclusion of a correction term in the Ewald summation is a possible
205 direction for handling 2D systems while still enabling the use of the
206 modern optimizations.\cite{Yeh99}
207
208 Several studies have recognized that the inherent periodicity in the
209 Ewald sum can also have an effect on three-dimensional
210 systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
211 Solvated proteins are essentially kept at high concentration due to
212 the periodicity of the electrostatic summation method. In these
213 systems, the more compact folded states of a protein can be
214 artificially stabilized by the periodic replicas introduced by the
215 Ewald summation.\cite{Weber00} Thus, care must be taken when
216 considering the use of the Ewald summation where the assumed
217 periodicity would introduce spurious effects in the system dynamics.
218
219 \subsection{The Wolf and Zahn Methods}
220 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
221 for the accurate accumulation of electrostatic interactions in an
222 efficient pairwise fashion. This procedure lacks the inherent
223 periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
224 observed that the electrostatic interaction is effectively
225 short-ranged in condensed phase systems and that neutralization of the
226 charge contained within the cutoff radius is crucial for potential
227 stability. They devised a pairwise summation method that ensures
228 charge neutrality and gives results similar to those obtained with the
229 Ewald summation. The resulting shifted Coulomb potential
230 (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
231 placement on the cutoff sphere and a distance-dependent damping
232 function (identical to that seen in the real-space portion of the
233 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 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 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 (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 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.
583 %
584 %\begin{figure}
585 %\centering
586 %\includegraphics[width = \linewidth]{./gaussFit.pdf}
587 %\caption{Sample fit of the angular distribution of the force vectors
588 %accumulated using all of the studied systems. Gaussian fits were used
589 %to obtain values for the variance in force and torque vectors.}
590 %\label{fig:gaussian}
591 %\end{figure}
592 %
593 %Figure \ref{fig:gaussian} shows an example distribution with applied
594 %non-linear fits. The solid line is a Gaussian profile, while the
595 %dotted line is a Voigt profile, a convolution of a Gaussian and a
596 %Lorentzian.
597 %Since this distribution is a measure of angular error between two
598 %different electrostatic summation methods, there is no {\it a priori}
599 %reason for the profile to adhere to any specific shape.
600 %Gaussian fits was used to compare all the tested methods.
601 The variance ($\sigma^2$) was extracted from each of these fits and
602 was used to compare distribution widths. Values of $\sigma^2$ near
603 zero indicate vector directions indistinguishable from those
604 calculated when using the reference method (SPME).
605
606 \subsection{Short-time Dynamics}
607
608 The effects of the alternative electrostatic summation methods on the
609 short-time dynamics of charged systems were evaluated by considering a
610 NaCl crystal at a temperature of 1000 K. A subset of the best
611 performing pairwise methods was used in this comparison. The NaCl
612 crystal was chosen to avoid possible complications from the treatment
613 of orientational motion in molecular systems. All systems were
614 started with the same initial positions and velocities. Simulations
615 were performed under the microcanonical ensemble, and velocity
616 autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
617 of the trajectories,
618 \begin{equation}
619 C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
620 \label{eq:vCorr}
621 \end{equation}
622 Velocity autocorrelation functions require detailed short time data,
623 thus velocity information was saved every 2 fs over 10 ps
624 trajectories. Because the NaCl crystal is composed of two different
625 atom types, the average of the two resulting velocity autocorrelation
626 functions was used for comparisons.
627
628 \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
629
630 The effects of the same subset of alternative electrostatic methods on
631 the {\it long-time} dynamics of charged systems were evaluated using
632 the same model system (NaCl crystals at 1000K). The power spectrum
633 ($I(\omega)$) was obtained via Fourier transform of the velocity
634 autocorrelation function, \begin{equation} I(\omega) =
635 \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
636 \label{eq:powerSpec}
637 \end{equation}
638 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
639 NaCl crystal is composed of two different atom types, the average of
640 the two resulting power spectra was used for comparisons. Simulations
641 were performed under the microcanonical ensemble, and velocity
642 information was saved every 5 fs over 100 ps trajectories.
643
644 \subsection{Representative Simulations}\label{sec:RepSims}
645 A variety of representative simulations were analyzed to determine the
646 relative effectiveness of the pairwise summation techniques in
647 reproducing the energetics and dynamics exhibited by SPME. We wanted
648 to span the space of modern simulations (i.e. from liquids of neutral
649 molecules to ionic crystals), so the systems studied were:
650 \begin{enumerate}
651 \item liquid water (SPC/E),\cite{Berendsen87}
652 \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
653 \item NaCl crystals,
654 \item NaCl melts,
655 \item a low ionic strength solution of NaCl in water (0.11 M),
656 \item a high ionic strength solution of NaCl in water (1.1 M), and
657 \item a 6 \AA\ radius sphere of Argon in water.
658 \end{enumerate}
659 By utilizing the pairwise techniques (outlined in section
660 \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
661 charged particles, and mixtures of the two, we hope to discern under
662 which conditions it will be possible to use one of the alternative
663 summation methodologies instead of the Ewald sum.
664
665 For the solid and liquid water configurations, configurations were
666 taken at regular intervals from high temperature trajectories of 1000
667 SPC/E water molecules. Each configuration was equilibrated
668 independently at a lower temperature (300~K for the liquid, 200~K for
669 the crystal). The solid and liquid NaCl systems consisted of 500
670 $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
671 these systems were selected and equilibrated in the same manner as the
672 water systems. The equilibrated temperatures were 1000~K for the NaCl
673 crystal and 7000~K for the liquid. The ionic solutions were made by
674 solvating 4 (or 40) ions in a periodic box containing 1000 SPC/E water
675 molecules. Ion and water positions were then randomly swapped, and
676 the resulting configurations were again equilibrated individually.
677 Finally, for the Argon / Water ``charge void'' systems, the identities
678 of all the SPC/E waters within 6 \AA\ of the center of the
679 equilibrated water configurations were converted to argon.
680 %(Fig. \ref{fig:argonSlice}).
681
682 These procedures guaranteed us a set of representative configurations
683 from chemically-relevant systems sampled from appropriate
684 ensembles. Force field parameters for the ions and Argon were taken
685 from the force field utilized by {\sc oopse}.\cite{Meineke05}
686
687 %\begin{figure}
688 %\centering
689 %\includegraphics[width = \linewidth]{./slice.pdf}
690 %\caption{A slice from the center of a water box used in a charge void
691 %simulation. The darkened region represents the boundary sphere within
692 %which the water molecules were converted to argon atoms.}
693 %\label{fig:argonSlice}
694 %\end{figure}
695
696 \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
697 We compared the following alternative summation methods with results
698 from the reference method (SPME):
699 \begin{itemize}
700 \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
701 and 0.3 \AA$^{-1}$,
702 \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
703 and 0.3 \AA$^{-1}$,
704 \item reaction field with an infinite dielectric constant, and
705 \item an unmodified cutoff.
706 \end{itemize}
707 Group-based cutoffs with a fifth-order polynomial switching function
708 were utilized for the reaction field simulations. Additionally, we
709 investigated the use of these cutoffs with the SP, SF, and pure
710 cutoff. The SPME electrostatics were performed using the TINKER
711 implementation of SPME,\cite{Ponder87} while all other calculations
712 were performed using the {\sc oopse} molecular mechanics
713 package.\cite{Meineke05} All other portions of the energy calculation
714 (i.e. Lennard-Jones interactions) were handled in exactly the same
715 manner across all systems and configurations.
716
717 The althernative methods were also evaluated with three different
718 cutoff radii (9, 12, and 15 \AA). As noted previously, the
719 convergence parameter ($\alpha$) plays a role in the balance of the
720 real-space and reciprocal-space portions of the Ewald calculation.
721 Typical molecular mechanics packages set this to a value dependent on
722 the cutoff radius and a tolerance (typically less than $1 \times
723 10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
724 increasing accuracy at the expense of computational time spent on the
725 reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
726 The default TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used
727 in all SPME calculations, resulting in Ewald coefficients of 0.4200,
728 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
729 respectively.
730
731 \section{Results and Discussion}
732
733 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
734 In order to evaluate the performance of the pairwise electrostatic
735 summation methods for Monte Carlo simulations, the energy differences
736 between configurations were compared to the values obtained when using
737 SPME. The results for the subsequent regression analysis are shown in
738 figure \ref{fig:delE}.
739
740 \begin{figure}
741 \centering
742 \includegraphics[width=5.5in]{./delEplot.pdf}
743 \caption{Statistical analysis of the quality of configurational energy
744 differences for a given electrostatic method compared with the
745 reference Ewald sum. Results with a value equal to 1 (dashed line)
746 indicate $\Delta E$ values indistinguishable from those obtained using
747 SPME. Different values of the cutoff radius are indicated with
748 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
749 inverted triangles).}
750 \label{fig:delE}
751 \end{figure}
752
753 The most striking feature of this plot is how well the Shifted Force
754 ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
755 differences. For the undamped {\sc sf} method, and the
756 moderately-damped {\sc sp} methods, the results are nearly
757 indistinguishable from the Ewald results. The other common methods do
758 significantly less well.
759
760 The unmodified cutoff method is essentially unusable. This is not
761 surprising since hard cutoffs give large energy fluctuations as atoms
762 or molecules move in and out of the cutoff
763 radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
764 some degree by using group based cutoffs with a switching
765 function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
766 significant improvement using the group-switched cutoff because the
767 salt and salt solution systems contain non-neutral groups. Interested
768 readers can consult the accompanying supporting information for a
769 comparison where all groups are neutral.
770
771 For the {\sc sp} method, inclusion of electrostatic damping improves
772 the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
773 shows an excellent correlation and quality of fit with the SPME
774 results, particularly with a cutoff radius greater than 12
775 \AA . Use of a larger damping parameter is more helpful for the
776 shortest cutoff shown, but it has a detrimental effect on simulations
777 with larger cutoffs.
778
779 In the {\sc sf} sets, increasing damping results in progressively {\it
780 worse} correlation with Ewald. Overall, the undamped case is the best
781 performing set, as the correlation and quality of fits are
782 consistently superior regardless of the cutoff distance. The undamped
783 case is also less computationally demanding (because no evaluation of
784 the complementary error function is required).
785
786 The reaction field results illustrates some of that method's
787 limitations, primarily that it was developed for use in homogenous
788 systems; although it does provide results that are an improvement over
789 those from an unmodified cutoff.
790
791 \subsection{Magnitudes of the Force and Torque Vectors}
792
793 Evaluation of pairwise methods for use in Molecular Dynamics
794 simulations requires consideration of effects on the forces and
795 torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
796 regression results for the force and torque vector magnitudes,
797 respectively. The data in these figures was generated from an
798 accumulation of the statistics from all of the system types.
799
800 \begin{figure}
801 \centering
802 \includegraphics[width=5.5in]{./frcMagplot.pdf}
803 \caption{Statistical analysis of the quality of the force vector
804 magnitudes for a given electrostatic method compared with the
805 reference Ewald sum. Results with a value equal to 1 (dashed line)
806 indicate force magnitude values indistinguishable from those obtained
807 using SPME. Different values of the cutoff radius are indicated with
808 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
809 inverted triangles).}
810 \label{fig:frcMag}
811 \end{figure}
812
813 Again, it is striking how well the Shifted Potential and Shifted Force
814 methods are doing at reproducing the SPME forces. The undamped and
815 weakly-damped {\sc sf} method gives the best agreement with Ewald.
816 This is perhaps expected because this method explicitly incorporates a
817 smooth transition in the forces at the cutoff radius as well as the
818 neutralizing image charges.
819
820 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
821 in the previous $\Delta E$ section. The unmodified cutoff results are
822 poor, but using group based cutoffs and a switching function provides
823 an improvement much more significant than what was seen with $\Delta
824 E$.
825
826 With moderate damping and a large enough cutoff radius, the {\sc sp}
827 method is generating usable forces. Further increases in damping,
828 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
829 detrimental to simulations with larger cutoff radii.
830
831 The reaction field results are surprisingly good, considering the poor
832 quality of the fits for the $\Delta E$ results. There is still a
833 considerable degree of scatter in the data, but the forces correlate
834 well with the Ewald forces in general. We note that the reaction
835 field calculations do not include the pure NaCl systems, so these
836 results are partly biased towards conditions in which the method
837 performs more favorably.
838
839 \begin{figure}
840 \centering
841 \includegraphics[width=5.5in]{./trqMagplot.pdf}
842 \caption{Statistical analysis of the quality of the torque vector
843 magnitudes for a given electrostatic method compared with the
844 reference Ewald sum. Results with a value equal to 1 (dashed line)
845 indicate torque magnitude values indistinguishable from those obtained
846 using SPME. Different values of the cutoff radius are indicated with
847 different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
848 inverted triangles).}
849 \label{fig:trqMag}
850 \end{figure}
851
852 Molecular torques were only available from the systems which contained
853 rigid molecules (i.e. the systems containing water). The data in
854 fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
855
856 Torques appear to be much more sensitive to charges at a longer
857 distance. The striking feature in comparing the new electrostatic
858 methods with SPME is how much the agreement improves with increasing
859 cutoff radius. Again, the weakly damped and undamped {\sc sf} method
860 appears to be reproducing the SPME torques most accurately.
861
862 Water molecules are dipolar, and the reaction field method reproduces
863 the effect of the surrounding polarized medium on each of the
864 molecular bodies. Therefore it is not surprising that reaction field
865 performs best of all of the methods on molecular torques.
866
867 \subsection{Directionality of the Force and Torque Vectors}
868
869 It is clearly important that a new electrostatic method can reproduce
870 the magnitudes of the force and torque vectors obtained via the Ewald
871 sum. However, the {\it directionality} of these vectors will also be
872 vital in calculating dynamical quantities accurately. Force and
873 torque directionalities were investigated by measuring the angles
874 formed between these vectors and the same vectors calculated using
875 SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
876 variance ($\sigma^2$) of the Gaussian fits of the angle error
877 distributions of the combined set over all system types.
878
879 \begin{figure}
880 \centering
881 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
882 \caption{Statistical analysis of the width of the angular distribution
883 that the force and torque vectors from a given electrostatic method
884 make with their counterparts obtained using the reference Ewald sum.
885 Results with a variance ($\sigma^2$) equal to zero (dashed line)
886 indicate force and torque directions indistinguishable from those
887 obtained using SPME. Different values of the cutoff radius are
888 indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
889 and 15\AA\ = inverted triangles).}
890 \label{fig:frcTrqAng}
891 \end{figure}
892
893 Both the force and torque $\sigma^2$ results from the analysis of the
894 total accumulated system data are tabulated in figure
895 \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
896 sp}) method would be essentially unusable for molecular dynamics until
897 the damping function is added. The Shifted Force ({\sc sf}) method,
898 however, is generating force and torque vectors which are within a few
899 degrees of the Ewald results even with weak (or no) damping.
900
901 All of the sets (aside from the over-damped case) show the improvement
902 afforded by choosing a larger cutoff radius. Increasing the cutoff
903 from 9 to 12 \AA\ typically results in a halving of the width of the
904 distribution, with a similar improvement going from 12 to 15
905 \AA .
906
907 The undamped {\sc sf}, group-based cutoff, and reaction field methods
908 all do equivalently well at capturing the direction of both the force
909 and torque vectors. Using damping improves the angular behavior
910 significantly for the {\sc sp} and moderately for the {\sc sf}
911 methods. Overdamping is detrimental to both methods. Again it is
912 important to recognize that the force vectors cover all particles in
913 the systems, while torque vectors are only available for neutral
914 molecular groups. Damping appears to have a more beneficial effect on
915 charged bodies, and this observation is investigated further in the
916 accompanying supporting information.
917
918 Although not discussed previously, group based cutoffs can be applied
919 to both the {\sc sp} and {\sc sf} methods. Use of a switching
920 function corrects for the discontinuities that arise when atoms of the
921 two groups exit the cutoff radius before the group centers leave each
922 other's cutoff. Though there are no significant benefits or drawbacks
923 observed in $\Delta E$ and vector magnitude results when doing this,
924 there is a measurable improvement in the vector angle results. Table
925 \ref{tab:groupAngle} shows the angular variance values obtained using
926 group based cutoffs and a switching function alongside the results
927 seen in figure \ref{fig:frcTrqAng}. The {\sc sp} shows much narrower
928 angular distributions for both the force and torque vectors when using
929 an $\alpha$ of 0.2 \AA$^{-1}$ or less, while {\sc sf} shows
930 improvements in the undamped and lightly damped cases. Thus, by
931 calculating the electrostatic interactions in terms of molecular pairs
932 rather than atomic pairs, the direction of the force and torque
933 vectors can be determined more accurately.
934
935 \begin{table}[htbp]
936 \centering
937 \caption{Variance ($\sigma^2$) of the force (top set) and torque
938 (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods. Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function. The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}
939 \begin{tabular}{@{} ccrrrrrrrr @{}}
940 \\
941 \toprule
942 & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
943 \cmidrule(lr){3-6}
944 \cmidrule(l){7-10}
945 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
946 \midrule
947
948 9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
949 & \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} \\
950 12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
951 & \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} \\
952 15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
953 & \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} \\
954
955 \midrule
956
957 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
958 & \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} \\
959 12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
960 & \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} \\
961 15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
962 & \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} \\
963
964 \bottomrule
965 \end{tabular}
966 \label{tab:groupAngle}
967 \end{table}
968
969 One additional trend to recognize in table \ref{tab:groupAngle} is
970 that the $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as
971 $\alpha$ increases, something that is easier to see when using group
972 based cutoffs. The reason for this is that the complimentary error
973 function inserted into the potential weakens the electrostatic
974 interaction as $\alpha$ increases. Thus, at larger values of
975 $\alpha$, both summation methods progress toward non-interacting
976 functions, so care is required in choosing large damping functions
977 lest one generate an undesirable loss in the pair interaction. Kast
978 \textit{et al.} developed a method for choosing appropriate $\alpha$
979 values for these types of electrostatic summation methods by fitting
980 to $g(r)$ data, and their methods indicate optimal values of 0.34,
981 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
982 respectively.\cite{Kast03} These appear to be reasonable choices to
983 obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
984 these findings, choices this high would introduce error in the
985 molecular torques, particularly for the shorter cutoffs. Based on the
986 above findings, empirical damping up to 0.2 \AA$^{-1}$ proves to be
987 beneficial, but damping may be unnecessary when using the {\sc sf}
988 method.
989
990 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
991
992 Zahn {\it et al.} investigated the structure and dynamics of water
993 using eqs. (\ref{eq:ZahnPot}) and
994 (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
995 that a method similar (but not identical with) the damped {\sc sf}
996 method resulted in properties very similar to those obtained when
997 using the Ewald summation. The properties they studied (pair
998 distribution functions, diffusion constants, and velocity and
999 orientational correlation functions) may not be particularly sensitive
1000 to the long-range and collective behavior that governs the
1001 low-frequency behavior in crystalline systems.
1002
1003 We are using two separate measures to probe the effects of these
1004 alternative electrostatic methods on the dynamics in crystalline
1005 materials. For short- and intermediate-time dynamics, we are
1006 computing the velocity autocorrelation function, and for long-time
1007 and large length-scale collective motions, we are looking at the
1008 low-frequency portion of the power spectrum.
1009
1010 \begin{figure}
1011 \centering
1012 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
1013 \caption{Velocity auto-correlation functions of NaCl crystals at
1014 1000 K using SPME, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1015 sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
1016 the first minimum. The times to first collision are nearly identical,
1017 but differences can be seen in the peaks and troughs, where the
1018 undamped and weakly damped methods are stiffer than the moderately
1019 damped and SPME methods.}
1020 \label{fig:vCorrPlot}
1021 \end{figure}
1022
1023 The short-time decays through the first collision are nearly identical
1024 in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
1025 functions show how the methods differ. The undamped {\sc sf} method
1026 has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
1027 peaks than any of the other methods. As the damping function is
1028 increased, these peaks are smoothed out, and approach the SPME
1029 curve. The damping acts as a distance dependent Gaussian screening of
1030 the point charges for the pairwise summation methods; thus, the
1031 collisions are more elastic in the undamped {\sc sf} potential, and the
1032 stiffness of the potential is diminished as the electrostatic
1033 interactions are softened by the damping function. With $\alpha$
1034 values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
1035 nearly identical and track the SPME features quite well. This is not
1036 too surprising in that the differences between the {\sc sf} and {\sc
1037 sp} potentials are mitigated with increased damping. However, this
1038 appears to indicate that once damping is utilized, the form of the
1039 potential seems to play a lesser role in the crystal dynamics.
1040
1041 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1042
1043 The short time dynamics were extended to evaluate how the differences
1044 between the methods affect the collective long-time motion. The same
1045 electrostatic summation methods were used as in the short time
1046 velocity autocorrelation function evaluation, but the trajectories
1047 were sampled over a much longer time. The power spectra of the
1048 resulting velocity autocorrelation functions were calculated and are
1049 displayed in figure \ref{fig:methodPS}.
1050
1051 \begin{figure}
1052 \centering
1053 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1054 \caption{Power spectra obtained from the velocity auto-correlation
1055 functions of NaCl crystals at 1000 K while using SPME, {\sc sf}
1056 ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).
1057 Apodization of the correlation functions via a cubic switching
1058 function between 40 and 50 ps was used to clear up the spectral noise
1059 resulting from data truncation, and had no noticeable effect on peak
1060 location or magnitude. The inset shows the frequency region below 100
1061 cm$^{-1}$ to highlight where the spectra begin to differ.}
1062 \label{fig:methodPS}
1063 \end{figure}
1064
1065 While high frequency peaks of the spectra in this figure overlap,
1066 showing the same general features, the low frequency region shows how
1067 the summation methods differ. Considering the low-frequency inset
1068 (expanded in the upper frame of figure \ref{fig:dampInc}), at
1069 frequencies below 100 cm$^{-1}$, the correlated motions are
1070 blue-shifted when using undamped or weakly damped {\sc sf}. When
1071 using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
1072 and {\sc sp} methods give near identical correlated motion behavior as
1073 the Ewald method (which has a damping value of 0.3119). This
1074 weakening of the electrostatic interaction with increased damping
1075 explains why the long-ranged correlated motions are at lower
1076 frequencies for the moderately damped methods than for undamped or
1077 weakly damped methods. To see this effect more clearly, we show how
1078 damping strength alone affects a simple real-space electrostatic
1079 potential,
1080 \begin{equation}
1081 V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
1082 \end{equation}
1083 where $S(r)$ is a switching function that smoothly zeroes the
1084 potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
1085 the low frequency motions are dependent on the damping used in the
1086 direct electrostatic sum. As the damping increases, the peaks drop to
1087 lower frequencies. Incidentally, use of an $\alpha$ of 0.25
1088 \AA$^{-1}$ on a simple electrostatic summation results in low
1089 frequency correlated dynamics equivalent to a simulation using SPME.
1090 When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1091 shift to higher frequency in exponential fashion. Though not shown,
1092 the spectrum for the simple undamped electrostatic potential is
1093 blue-shifted such that the lowest frequency peak resides near 325
1094 cm$^{-1}$. In light of these results, the undamped {\sc sf} method
1095 producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1096 respectable and shows that the shifted force procedure accounts for
1097 most of the effect afforded through use of the Ewald summation.
1098 However, it appears as though moderate damping is required for
1099 accurate reproduction of crystal dynamics.
1100 \begin{figure}
1101 \centering
1102 \includegraphics[width = \linewidth]{./comboSquare.pdf}
1103 \caption{Regions of spectra showing the low-frequency correlated
1104 motions for NaCl crystals at 1000 K using various electrostatic
1105 summation methods. The upper plot is a zoomed inset from figure
1106 \ref{fig:methodPS}. As the damping value for the {\sc sf} potential
1107 increases, the low-frequency peaks red-shift. The lower plot is of
1108 spectra when using SPME and a simple damped Coulombic sum with damping
1109 coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$. As
1110 $\alpha$ increases, the peaks are red-shifted toward and eventually
1111 beyond the values given by SPME. The larger $\alpha$ values weaken
1112 the real-space electrostatics, explaining this shift towards less
1113 strongly correlated motions in the crystal.}
1114 \label{fig:dampInc}
1115 \end{figure}
1116
1117 \section{Conclusions}
1118
1119 This investigation of pairwise electrostatic summation techniques
1120 shows that there are viable and more computationally efficient
1121 electrostatic summation techniques than the Ewald summation, chiefly
1122 methods derived from the damped Coulombic sum originally proposed by
1123 Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1124 {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1125 shows a remarkable ability to reproduce the energetic and dynamic
1126 characteristics exhibited by simulations employing lattice summation
1127 techniques. The cumulative energy difference results showed the
1128 undamped {\sc sf} and moderately damped {\sc sp} methods
1129 produced results nearly identical to SPME. Similarly for the dynamic
1130 features, the undamped or moderately damped {\sc sf} and
1131 moderately damped {\sc sp} methods produce force and torque
1132 vector magnitude and directions very similar to the expected values.
1133 These results translate into long-time dynamic behavior equivalent to
1134 that produced in simulations using SPME.
1135
1136 Aside from the computational cost benefit, these techniques have
1137 applicability in situations where the use of the Ewald sum can prove
1138 problematic. Primary among them is their use in interfacial systems,
1139 where the unmodified lattice sum techniques artificially accentuate
1140 the periodicity of the system in an undesirable manner. There have
1141 been alterations to the standard Ewald techniques, via corrections and
1142 reformulations, to compensate for these systems; but the pairwise
1143 techniques discussed here require no modifications, making them
1144 natural tools to tackle these problems. Additionally, this
1145 transferability gives them benefits over other pairwise methods, like
1146 reaction field, because estimations of physical properties (e.g. the
1147 dielectric constant) are unnecessary.
1148
1149 We are not suggesting any flaw with the Ewald sum; in fact, it is the
1150 standard by which these simple pairwise sums are judged. However,
1151 these results do suggest that in the typical simulations performed
1152 today, the Ewald summation may no longer be required to obtain the
1153 level of accuracy most researchers have come to expect
1154
1155 \section{Acknowledgments}
1156 \newpage
1157
1158 \bibliographystyle{jcp2}
1159 \bibliography{electrostaticMethods}
1160
1161
1162 \end{document}