ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2639
Committed: Mon Mar 20 02:00:26 2006 UTC (18 years, 3 months ago) by chrisfen
Content type: application/x-tex
File size: 60060 byte(s)
Log Message:
All sections filled

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 considered one of the most essential and
69 computationally demanding tasks. The common molecular mechanics force
70 fields are founded on representation of the atomic sites centered on
71 full or partial charges shielded by Lennard-Jones type interactions.
72 This means that nearly every pair interaction involves an
73 charge-charge calculation. Coupled with $r^{-1}$ decay, the monopole
74 interactions quickly become a burden for molecular systems of all
75 sizes. For example, in small systems, the electrostatic pair
76 interaction may not have decayed appreciably within the box length
77 leading to an effect excluded from the pair interactions within a unit
78 box. In large systems, excessively large cutoffs need to be used to
79 accurately incorporate their effect, and since the computational cost
80 increases proportionally with the cutoff sphere, it quickly becomes
81 very time-consuming to perform these calculations.
82
83 There have been many efforts to address this issue of both proper and
84 practical handling of electrostatic interactions, and these have
85 resulted in the availability of a variety of
86 techniques.\cite{Roux99,Sagui99,Tobias01} These are typically
87 classified as implicit methods (i.e., continuum dielectrics, static
88 dipolar fields),\cite{Born20,Grossfield00} explicit methods (i.e.,
89 Ewald summations, interaction shifting or
90 trucation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
91 reaction field type methods, fast multipole
92 methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
93 often preferred because they incorporate dynamic solvent molecules in
94 the system of interest, but these methods are sometimes difficult to
95 utilize because of their high computational cost.\cite{Roux99} In
96 addition to this cost, there has been some question of the inherent
97 periodicity of the explicit Ewald summation artificially influencing
98 systems dynamics.\cite{Tobias01}
99
100 In this paper, we focus on the common mixed and explicit methods of
101 reaction filed and smooth particle mesh
102 Ewald\cite{Onsager36,Essmann99} and a new set of shifted methods
103 devised by Wolf {\it et al.} which we further extend.\cite{Wolf99}
104 These new methods for handling electrostatics are quite
105 computationally efficient, since they involve only a simple
106 modification to the direct pairwise sum, and they lack the added
107 periodicity of the Ewald sum. Below, these methods are evaluated using
108 a variety of model systems and comparison methodologies to establish
109 their useability 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 simulation box, as well as those in the
115 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 Poisson's equation ($\phi(\mathbf{r}_{ij}) = q_i
127 q_j |\mathbf{r}_{ij}|^{-1}$ for charge-charge interactions). In the
128 case of monopole electrostatics, eq. (\ref{eq:PBCSum}) is
129 conditionally convergent and is discontiuous for non-neutral systems.
130
131 This electrostatic summation problem was originally studied by Ewald
132 for the case of an infinite crystal.\cite{Ewald21}. The approach he
133 took was to convert this conditionally convergent sum into two
134 absolutely convergent summations: a short-ranged real-space summation
135 and a long-ranged reciprocal-space summation,
136 \begin{equation}
137 \begin{split}
138 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,
139 \end{split}
140 \label{eq:EwaldSum}
141 \end{equation}
142 where $\alpha$ is a damping parameter, or separation constant, with
143 units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and equal
144 $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
145 constant of the encompassing medium. The final two terms of
146 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
147 for interacting with a surrounding dielectric.\cite{Allen87} This
148 dipolar term was neglected in early applications in molecular
149 simulations,\cite{Brush66,Woodcock71} until it was introduced by de
150 Leeuw {\it et al.} to address situations where the unit cell has a
151 dipole moment and this dipole moment gets magnified through
152 replication of the periodic images.\cite{deLeeuw80,Smith81} If this
153 term is taken to be zero, the system is using conducting boundary
154 conditions, $\epsilon_{\rm S} = \infty$. Figure
155 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
156 time. Initially, due to the small size of systems, the entire
157 simulation box was replicated to convergence. Currently, we balance a
158 spherical real-space cutoff with the reciprocal sum and consider the
159 surrounding dielectric.
160 \begin{figure}
161 \centering
162 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
163 \caption{How the application of the Ewald summation has changed with
164 the increase in computer power. Initially, only small numbers of
165 particles could be studied, and the Ewald sum acted to replicate the
166 unit cell charge distribution out to convergence. Now, much larger
167 systems of charges are investigated with fixed distance cutoffs. The
168 calculated structure factor is used to sum out to great distance, and
169 a surrounding dielectric term is included.}
170 \label{fig:ewaldTime}
171 \end{figure}
172
173 The Ewald summation in the straight-forward form is an
174 $\mathscr{O}(N^2)$ algorithm. The separation constant $(\alpha)$
175 plays an important role in the computational cost balance between the
176 direct and reciprocal-space portions of the summation. The choice of
177 the magnitude of this value allows one to select whether the
178 real-space or reciprocal space portion of the summation is an
179 $\mathscr{O}(N^2)$ calcualtion (with the other being
180 $\mathscr{O}(N)$).\cite{Sagui99} With appropriate choice of $\alpha$
181 and thoughtful algorithm development, this cost can be brought down to
182 $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
183 reduce the cost of the Ewald summation further is to set $\alpha$ such
184 that the real-space interactions decay rapidly, allowing for a short
185 spherical cutoff, and then optimize the reciprocal space summation.
186 These optimizations usually involve the utilization of the fast
187 Fourier transform (FFT),\cite{Hockney81} leading to the
188 particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
189 methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
190 methods, the cost of the reciprocal-space portion of the Ewald
191 summation is from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N \log N)$.
192
193 These developments and optimizations have led the use of the Ewald
194 summation to become routine in simulations with periodic boundary
195 conditions. However, in certain systems the intrinsic three
196 dimensional periodicity can prove to be problematic, such as two
197 dimensional surfaces and membranes. The Ewald sum has been
198 reformulated to handle 2D
199 systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the new
200 methods have been found to be computationally
201 expensive.\cite{Spohr97,Yeh99} Inclusion of a correction term in the
202 full Ewald summation is a possible direction for enabling the handling
203 of 2D systems and the inclusion of the optimizations described
204 previously.\cite{Yeh99}
205
206 Several studies have recognized that the inherent periodicity in the
207 Ewald sum can also have an effect on systems that have the same
208 dimensionality.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
209 Good examples are solvated proteins kept at high relative
210 concentration due to the periodicity of the electrostatics. In these
211 systems, the more compact folded states of a protein can be
212 artificially stabilized by the periodic replicas introduced by the
213 Ewald summation.\cite{Weber00} Thus, care ought to be taken when
214 considering the use of the Ewald summation where the intrinsic
215 perodicity may negatively affect the system dynamics.
216
217
218 \subsection{The Wolf and Zahn Methods}
219 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
220 for the accurate accumulation of electrostatic interactions in an
221 efficient pairwise fashion and lacks the inherent periodicity of the
222 Ewald summation.\cite{Wolf99} Wolf \textit{et al.} observed that the
223 electrostatic interaction is effectively short-ranged in condensed
224 phase systems and that neutralization of the charge contained within
225 the cutoff radius is crucial for potential stability. They devised a
226 pairwise summation method that ensures charge neutrality and gives
227 results similar to those obtained with the Ewald summation. The
228 resulting shifted Coulomb potential (Eq. \ref{eq:WolfPot}) includes
229 image-charges subtracted out through placement on the cutoff sphere
230 and a distance-dependent damping function (identical to that seen in
231 the real-space portion of the Ewald sum) to aid convergence
232 \begin{equation}
233 V_{\textrm{Wolf}}(r_{ij})= \frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
234 \label{eq:WolfPot}
235 \end{equation}
236 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
237 potential. However, neutralizing the charge contained within each
238 cutoff sphere requires the placement of a self-image charge on the
239 surface of the cutoff sphere. This additional self-term in the total
240 potential enabled Wolf {\it et al.} to obtain excellent estimates of
241 Madelung energies for many crystals.
242
243 In order to use their charge-neutralized potential in molecular
244 dynamics simulations, Wolf \textit{et al.} suggested taking the
245 derivative of this potential prior to evaluation of the limit. This
246 procedure gives an expression for the forces,
247 \begin{equation}
248 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\},
249 \label{eq:WolfForces}
250 \end{equation}
251 that incorporates both image charges and damping of the electrostatic
252 interaction.
253
254 More recently, Zahn \textit{et al.} investigated these potential and
255 force expressions for use in simulations involving water.\cite{Zahn02}
256 In their work, they pointed out that the forces and derivative of
257 the potential are not commensurate. Attempts to use both
258 Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
259 to poor energy conservation. They correctly observed that taking the
260 limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
261 derivatives gives forces for a different potential energy function
262 than the one shown in Eq. (\ref{eq:WolfPot}).
263
264 Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
265 method'' as a way to use this technique in Molecular Dynamics
266 simulations. Taking the integral of the forces shown in equation
267 \ref{eq:WolfForces}, they proposed a new damped Coulomb
268 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 They 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 distance, while the neutralizing image charges are present
377 in both the energy and force expressions. It would be simple to add
378 the 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 they found that
387 it was still insufficient for accurate determination of the energy
388 with reasonable cutoff distances. The calculated Madelung energies
389 fluctuate around the expected value with increasing cutoff radius, but
390 the oscillations converge toward the correct value.\cite{Wolf99} A
391 damping function was incorporated to accelerate the convergence; and
392 though alternative functional forms 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})) can be reacquired using
402 eq. (\ref{eq:shiftingForm}),
403 \begin{equation}
404 v_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}-\frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right) \quad r\leqslant R_\textrm{c},
405 \label{eq:DSPPot}
406 \end{equation}
407 with associated forces,
408 \begin{equation}
409 f_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \quad r\leqslant R_\textrm{c}.
410 \label{eq:DSPForces}
411 \end{equation}
412 Again, this damped shifted potential suffers from a discontinuity and
413 a lack of the image charges in the forces. To remedy these concerns,
414 one may derive a {\sc sf} variant by including the derivative
415 term in 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 chosen to be zero, the undamped
430 case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
431 from eqs. (\ref{eq:DSPPot}-\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 would be a potential source of errors,
444 however. In fact, it introduces a discontinuity in the forces at the
445 cutoff, because the force function is shifted in the wrong direction
446 and 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 that is continuous in both the
450 potential and forces and which incorporates the damping function
451 proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
452 paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
453 sf}, damping) are at reproducing the correct electrostatic summation
454 performed by the Ewald sum.
455
456 \subsection{Other alternatives}
457 In addition to the methods described above, we will consider some
458 other techniques that commonly get used in molecular simulations. The
459 simplest of these is group-based cutoffs. Though of little use for
460 non-neutral 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 groups, an orientational aspect is
464 introduced to the interactions. Consequently, as these molecular
465 particles move through $R_\textrm{c}$, the energy will drift upward
466 due to the anisotropy of the net molecular dipole
467 interactions.\cite{Rahman71} To maintain good energy conservation,
468 both the potential and derivative need to be smoothly switched to zero
469 at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
470 switching function,
471 \begin{equation}
472 S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
473 \frac{(R_\textrm{c}+2r-3r_\textrm{sw})(R_\textrm{c}-r)^2}{(R_\textrm{c}-r_\textrm{sw})^3} &\quad r_\textrm{sw}<r\leqslant R_\textrm{c} \\
474 0 &\quad r>R_\textrm{c}
475 \end{cases},
476 \end{equation}
477 where the above form is for a cubic function. If a smooth second
478 derivative is desired, a fifth (or higher) order polynomial can be
479 used.\cite{Andrea83}
480
481 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
482 and to incorporate their effect, a method like Reaction Field ({\sc
483 rf}) can be used. The original theory for {\sc rf} was originally
484 developed by Onsager,\cite{Onsager36} and it was applied in
485 simulations for the study of water by Barker and Watts.\cite{Barker73}
486 In application, it is simply an extension of the group-based cutoff
487 method where the net dipole within the cutoff sphere polarizes an
488 external dielectric, which reacts back on the central dipole. The
489 same switching function considerations for group-based cutoffs need to
490 made for {\sc rf}, with the additional pre-specification of a
491 dielectric constant.
492
493 \section{Methods}
494
495 In classical molecular mechanics simulations, there are two primary
496 techniques utilized to obtain information about the system of
497 interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
498 techniques utilize pairwise summations of interactions between
499 particle sites, but they use these summations in different ways.
500
501 In MC, the potential energy difference between two subsequent
502 configurations dictates the progression of MC sampling. Going back to
503 the origins of this method, the acceptance criterion for the canonical
504 ensemble laid out by Metropolis \textit{et al.} states that a
505 subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
506 \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
507 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
508 alternate method for handling the long-range electrostatics will
509 ensure proper sampling from the ensemble.
510
511 In MD, the derivative of the potential governs how the system will
512 progress in time. Consequently, the force and torque vectors on each
513 body in the system dictate how the system evolves. If the magnitude
514 and direction of these vectors are similar when using alternate
515 electrostatic summation techniques, the dynamics in the short term
516 will be indistinguishable. Because error in MD calculations is
517 cumulative, one should expect greater deviation at longer times,
518 although methods which have large differences in the force and torque
519 vectors will diverge from each other more rapidly.
520
521 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
522 The pairwise summation techniques (outlined in section
523 \ref{sec:ESMethods}) were evaluated for use in MC simulations by
524 studying the energy differences between conformations. We took the
525 SPME-computed energy difference between two conformations to be the
526 correct behavior. An ideal performance by an alternative method would
527 reproduce these energy differences exactly. Since none of the methods
528 provide exact energy differences, we used linear least squares
529 regressions of the $\Delta E$ values between configurations using SPME
530 against $\Delta E$ values using tested methods provides a quantitative
531 comparison of this agreement. Unitary results for both the
532 correlation and correlation coefficient for these regressions indicate
533 equivalent energetic results between the method under consideration
534 and electrostatics handled using SPME. Sample correlation plots for
535 two alternate methods are shown in Fig. \ref{fig:linearFit}.
536
537 \begin{figure}
538 \centering
539 \includegraphics[width = \linewidth]{./dualLinear.pdf}
540 \caption{Example least squares regressions of the configuration energy differences for SPC/E water systems. The upper plot shows a data set with a poor correlation coefficient ($R^2$), while the lower plot shows a data set with a good correlation coefficient.}
541 \label{fig:linearFit}
542 \end{figure}
543
544 Each system type (detailed in section \ref{sec:RepSims}) was
545 represented using 500 independent configurations. Additionally, we
546 used seven different system types, so each of the alternate
547 (non-Ewald) electrostatic summation methods was evaluated using
548 873,250 configurational energy differences.
549
550 Results and discussion for the individual analysis of each of the
551 system types appear in the supporting information, while the
552 cumulative results over all the investigated systems appears below in
553 section \ref{sec:EnergyResults}.
554
555 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
556 We evaluated the pairwise methods (outlined in section
557 \ref{sec:ESMethods}) for use in MD simulations by
558 comparing the force and torque vectors with those obtained using the
559 reference Ewald summation (SPME). Both the magnitude and the
560 direction of these vectors on each of the bodies in the system were
561 analyzed. For the magnitude of these vectors, linear least squares
562 regression analyses were performed as described previously for
563 comparing $\Delta E$ values. Instead of a single energy difference
564 between two system configurations, we compared the magnitudes of the
565 forces (and torques) on each molecule in each configuration. For a
566 system of 1000 water molecules and 40 ions, there are 1040 force
567 vectors and 1000 torque vectors. With 500 configurations, this
568 results in 520,000 force and 500,000 torque vector comparisons.
569 Additionally, data from seven different system types was aggregated
570 before the comparison was made.
571
572 The {\it directionality} of the force and torque vectors was
573 investigated through measurement of the angle ($\theta$) formed
574 between those computed from the particular method and those from SPME,
575 \begin{equation}
576 \theta_f = \cos^{-1} \left(\hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method}\right),
577 \end{equation}
578 where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
579 force vector computed using method $M$.
580
581 Each of these $\theta$ values was accumulated in a distribution
582 function, weighted by the area on the unit sphere. Non-linear
583 Gaussian fits were used to measure the width of the resulting
584 distributions.
585
586 \begin{figure}
587 \centering
588 \includegraphics[width = \linewidth]{./gaussFit.pdf}
589 \caption{Sample fit of the angular distribution of the force vectors over all of the studied systems. Gaussian fits were used to obtain values for the variance in force and torque vectors used in the following figure.}
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. Since this distribution is a measure of angular error
597 between two different electrostatic summation methods, there is no
598 {\it a priori} reason for the profile to adhere to any specific shape.
599 Gaussian fits was used to compare all the tested methods. The
600 variance ($\sigma^2$) was extracted from each of these fits and was
601 used to compare distribution widths. Values of $\sigma^2$ near zero
602 indicate vector directions indistinguishable from those calculated
603 when using the reference method (SPME).
604
605 \subsection{Short-time Dynamics}
606 Evaluation of the short-time dynamics of charged systems was performed
607 by considering the 1000 K NaCl crystal system while using a subset of the
608 best performing pairwise methods. The NaCl crystal was chosen to
609 avoid possible complications involving the propagation techniques of
610 orientational motion in molecular systems. All systems were started
611 with the same initial positions and velocities. Simulations were
612 performed under the microcanonical ensemble, and velocity
613 autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
614 of the trajectories,
615 \begin{equation}
616 C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
617 \label{eq:vCorr}
618 \end{equation}
619 Velocity autocorrelation functions require detailed short time data,
620 thus velocity information was saved every 2 fs over 10 ps
621 trajectories. Because the NaCl crystal is composed of two different
622 atom types, the average of the two resulting velocity autocorrelation
623 functions was used for comparisons.
624
625 \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
626 Evaluation of the long-time dynamics of charged systems was performed
627 by considering the NaCl crystal system, again while using a subset of
628 the best performing pairwise methods. To enhance the atomic motion,
629 these crystals were equilibrated at 1000 K, near the experimental
630 $T_m$ for NaCl. Simulations were performed under the microcanonical
631 ensemble, and velocity information was saved every 5 fs over 100 ps
632 trajectories. The power spectrum ($I(\omega)$) was obtained via
633 Fourier transform of the velocity autocorrelation function
634 \begin{equation}
635 I(\omega) = \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.
641
642 \subsection{Representative Simulations}\label{sec:RepSims}
643 A variety of common and representative simulations were analyzed to
644 determine the relative effectiveness of the pairwise summation
645 techniques in reproducing the energetics and dynamics exhibited by
646 SPME. The studied systems were as follows:
647 \begin{enumerate}
648 \item Liquid Water
649 \item Crystalline Water (Ice I$_\textrm{c}$)
650 \item NaCl Crystal
651 \item NaCl Melt
652 \item Low Ionic Strength Solution of NaCl in Water
653 \item High Ionic Strength Solution of NaCl in Water
654 \item 6 \AA\ Radius Sphere of Argon in Water
655 \end{enumerate}
656 By utilizing the pairwise techniques (outlined in section
657 \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
658 charged particles, and mixtures of the two, we can comment on possible
659 system dependence and/or universal applicability of the techniques.
660
661 Generation of the system configurations was dependent on the system
662 type. For the solid and liquid water configurations, configuration
663 snapshots were taken at regular intervals from higher temperature 1000
664 SPC/E water molecule trajectories and each equilibrated individually.
665 The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
666 ions and were selected and equilibrated in the same fashion as the
667 water systems. For the low and high ionic strength NaCl solutions, 4
668 and 40 ions were first solvated in a 1000 water molecule boxes
669 respectively. Ion and water positions were then randomly swapped, and
670 the resulting configurations were again equilibrated individually.
671 Finally, for the Argon/Water "charge void" systems, the identities of
672 all the SPC/E waters within 6 \AA\ of the center of the equilibrated
673 water configurations were converted to argon
674 (Fig. \ref{fig:argonSlice}).
675
676 \begin{figure}
677 \centering
678 \includegraphics[width = \linewidth]{./slice.pdf}
679 \caption{A slice from the center of a water box used in a charge void simulation. The darkened region represents the boundary sphere within which the water molecules were converted to argon atoms.}
680 \label{fig:argonSlice}
681 \end{figure}
682
683 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
684 Electrostatic summation method comparisons were performed using SPME,
685 the {\sc sp} and {\sc sf} methods - both with damping
686 parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
687 moderate, and strong damping respectively), reaction field with an
688 infinite dielectric constant, and an unmodified cutoff. Group-based
689 cutoffs with a fifth-order polynomial switching function were
690 necessary for the reaction field simulations and were utilized in the
691 SP, SF, and pure cutoff methods for comparison to the standard lack of
692 group-based cutoffs with a hard truncation. The SPME calculations
693 were performed using the TINKER implementation of SPME,\cite{Ponder87}
694 while all other method calculations were performed using the OOPSE
695 molecular mechanics package.\cite{Meineke05}
696
697 These methods were additionally evaluated with three different cutoff
698 radii (9, 12, and 15 \AA) to investigate possible cutoff radius
699 dependence. It should be noted that the damping parameter chosen in
700 SPME, or so called ``Ewald Coefficient", has a significant effect on
701 the energies and forces calculated. Typical molecular mechanics
702 packages default this to a value dependent on the cutoff radius and a
703 tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller
704 tolerances are typically associated with increased accuracy, but this
705 usually means more time spent calculating the reciprocal-space portion
706 of the summation.\cite{Perram88,Essmann95} The default TINKER
707 tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
708 calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
709 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
710
711 \section{Results and Discussion}
712
713 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
714 In order to evaluate the performance of the pairwise electrostatic
715 summation methods for Monte Carlo simulations, the energy differences
716 between configurations were compared to the values obtained when using
717 SPME. The results for the subsequent regression analysis are shown in
718 figure \ref{fig:delE}.
719
720 \begin{figure}
721 \centering
722 \includegraphics[width=5.5in]{./delEplot.pdf}
723 \caption{Statistical analysis of the quality of configurational energy differences for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate $\Delta E$ values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
724 \label{fig:delE}
725 \end{figure}
726
727 In this figure, it is apparent that it is unreasonable to expect
728 realistic results using an unmodified cutoff. This is not all that
729 surprising since this results in large energy fluctuations as atoms
730 move in and out of the cutoff radius. These fluctuations can be
731 alleviated to some degree by using group based cutoffs with a
732 switching function.\cite{Steinbach94} The Group Switch Cutoff row
733 doesn't show a significant improvement in this plot because the salt
734 and salt solution systems contain non-neutral groups, see the
735 accompanying supporting information for a comparison where all groups
736 are neutral.
737
738 Correcting the resulting charged cutoff sphere is one of the purposes
739 of the damped Coulomb summation proposed by Wolf \textit{et
740 al.},\cite{Wolf99} and this correction indeed improves the results as
741 seen in the Shifted-Potental rows. While the undamped case of this
742 method is a significant improvement over the pure cutoff, it still
743 doesn't correlate that well with SPME. Inclusion of potential damping
744 improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
745 an excellent correlation and quality of fit with the SPME results,
746 particularly with a cutoff radius greater than 12 \AA . Use of a
747 larger damping parameter is more helpful for the shortest cutoff
748 shown, but it has a detrimental effect on simulations with larger
749 cutoffs. In the {\sc sf} sets, increasing damping results in
750 progressively poorer correlation. Overall, the undamped case is the
751 best performing set, as the correlation and quality of fits are
752 consistently superior regardless of the cutoff distance. This result
753 is beneficial in that the undamped case is less computationally
754 prohibitive do to the lack of complimentary error function calculation
755 when performing the electrostatic pair interaction. The reaction
756 field results illustrates some of that method's limitations, primarily
757 that it was developed for use in homogenous systems; although it does
758 provide results that are an improvement over those from an unmodified
759 cutoff.
760
761 \subsection{Magnitudes of the Force and Torque Vectors}
762
763 Evaluation of pairwise methods for use in Molecular Dynamics
764 simulations requires consideration of effects on the forces and
765 torques. Investigation of the force and torque vector magnitudes
766 provides a measure of the strength of these values relative to SPME.
767 Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
768 force and torque vector magnitude regression results for the
769 accumulated analysis over all the system types.
770
771 \begin{figure}
772 \centering
773 \includegraphics[width=5.5in]{./frcMagplot.pdf}
774 \caption{Statistical analysis of the quality of the force vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate force magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
775 \label{fig:frcMag}
776 \end{figure}
777
778 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
779 in the previous $\Delta E$ section. The unmodified cutoff results are
780 poor, but using group based cutoffs and a switching function provides
781 a improvement much more significant than what was seen with $\Delta
782 E$. Looking at the {\sc sp} sets, the slope and $R^2$
783 improve with the use of damping to an optimal result of 0.2 \AA
784 $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping,
785 while beneficial for simulations with a cutoff radius of 9 \AA\ , is
786 detrimental to simulations with larger cutoff radii. The undamped
787 {\sc sf} method gives forces in line with those obtained using
788 SPME, and use of a damping function results in minor improvement. The
789 reaction field results are surprisingly good, considering the poor
790 quality of the fits for the $\Delta E$ results. There is still a
791 considerable degree of scatter in the data, but it correlates well in
792 general. To be fair, we again note that the reaction field
793 calculations do not encompass NaCl crystal and melt systems, so these
794 results are partly biased towards conditions in which the method
795 performs more favorably.
796
797 \begin{figure}
798 \centering
799 \includegraphics[width=5.5in]{./trqMagplot.pdf}
800 \caption{Statistical analysis of the quality of the torque vector magnitudes for a given electrostatic method compared with the reference Ewald sum. Results with a value equal to 1 (dashed line) indicate torque magnitude values indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
801 \label{fig:trqMag}
802 \end{figure}
803
804 To evaluate the torque vector magnitudes, the data set from which
805 values are drawn is limited to rigid molecules in the systems
806 (i.e. water molecules). In spite of this smaller sampling pool, the
807 torque vector magnitude results in figure \ref{fig:trqMag} are still
808 similar to those seen for the forces; however, they more clearly show
809 the improved behavior that comes with increasing the cutoff radius.
810 Moderate damping is beneficial to the {\sc sp} and helpful
811 yet possibly unnecessary with the {\sc sf} method, and they also
812 show that over-damping adversely effects all cutoff radii rather than
813 showing an improvement for systems with short cutoffs. The reaction
814 field method performs well when calculating the torques, better than
815 the Shifted Force method over this limited data set.
816
817 \subsection{Directionality of the Force and Torque Vectors}
818
819 Having force and torque vectors with magnitudes that are well
820 correlated to SPME is good, but if they are not pointing in the proper
821 direction the results will be incorrect. These vector directions were
822 investigated through measurement of the angle formed between them and
823 those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared
824 through the variance ($\sigma^2$) of the Gaussian fits of the angle
825 error distributions of the combined set over all system types.
826
827 \begin{figure}
828 \centering
829 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
830 \caption{Statistical analysis of the quality of the Gaussian fit of the force and torque vector angular distributions for a given electrostatic method compared with the reference Ewald sum. Results with a variance ($\sigma^2$) equal to zero (dashed line) indicate force and torque directions indistinguishable from those obtained using SPME. Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
831 \label{fig:frcTrqAng}
832 \end{figure}
833
834 Both the force and torque $\sigma^2$ results from the analysis of the
835 total accumulated system data are tabulated in figure
836 \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case
837 show the improvement afforded by choosing a longer simulation cutoff.
838 Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
839 of the distribution widths, with a similar improvement going from 12
840 to 15 \AA . The undamped {\sc sf}, Group Based Cutoff, and
841 Reaction Field methods all do equivalently well at capturing the
842 direction of both the force and torque vectors. Using damping
843 improves the angular behavior significantly for the {\sc sp}
844 and moderately for the {\sc sf} methods. Increasing the damping
845 too far is destructive for both methods, particularly to the torque
846 vectors. Again it is important to recognize that the force vectors
847 cover all particles in the systems, while torque vectors are only
848 available for neutral molecular groups. Damping appears to have a
849 more beneficial effect on non-neutral bodies, and this observation is
850 investigated further in the accompanying supporting information.
851
852 \begin{table}[htbp]
853 \centering
854 \caption{Variance ($\sigma^2$) of the force (top set) and torque (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods. Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function. The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}
855 \begin{tabular}{@{} ccrrrrrrrr @{}}
856 \\
857 \toprule
858 & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
859 \cmidrule(lr){3-6}
860 \cmidrule(l){7-10}
861 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
862 \midrule
863
864 9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
865 & \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} \\
866 12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
867 & \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} \\
868 15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
869 & \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} \\
870
871 \midrule
872
873 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
874 & \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} \\
875 12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
876 & \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} \\
877 15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
878 & \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} \\
879
880 \bottomrule
881 \end{tabular}
882 \label{tab:groupAngle}
883 \end{table}
884
885 Although not discussed previously, group based cutoffs can be applied
886 to both the {\sc sp} and {\sc sf} methods. Use off a
887 switching function corrects for the discontinuities that arise when
888 atoms of a group exit the cutoff before the group's center of mass.
889 Though there are no significant benefit or drawbacks observed in
890 $\Delta E$ and vector magnitude results when doing this, there is a
891 measurable improvement in the vector angle results. Table
892 \ref{tab:groupAngle} shows the angular variance values obtained using
893 group based cutoffs and a switching function alongside the standard
894 results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
895 The {\sc sp} shows much narrower angular distributions for
896 both the force and torque vectors when using an $\alpha$ of 0.2
897 \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
898 undamped and lightly damped cases. Thus, by calculating the
899 electrostatic interactions in terms of molecular pairs rather than
900 atomic pairs, the direction of the force and torque vectors are
901 determined more accurately.
902
903 One additional trend to recognize in table \ref{tab:groupAngle} is
904 that the $\sigma^2$ values for both {\sc sp} and
905 {\sc sf} converge as $\alpha$ increases, something that is easier
906 to see when using group based cutoffs. Looking back on figures
907 \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
908 behavior clearly at large $\alpha$ and cutoff values. The reason for
909 this is that the complimentary error function inserted into the
910 potential weakens the electrostatic interaction as $\alpha$ increases.
911 Thus, at larger values of $\alpha$, both the summation method types
912 progress toward non-interacting functions, so care is required in
913 choosing large damping functions lest one generate an undesirable loss
914 in the pair interaction. Kast \textit{et al.} developed a method for
915 choosing appropriate $\alpha$ values for these types of electrostatic
916 summation methods by fitting to $g(r)$ data, and their methods
917 indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
918 values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
919 to be reasonable choices to obtain proper MC behavior
920 (Fig. \ref{fig:delE}); however, based on these findings, choices this
921 high would introduce error in the molecular torques, particularly for
922 the shorter cutoffs. Based on the above findings, empirical damping
923 up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
924 unnecessary when using the {\sc sf} method.
925
926 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
927
928 In the previous studies using a {\sc sf} variant of the damped
929 Wolf coulomb potential, the structure and dynamics of water were
930 investigated rather extensively.\cite{Zahn02,Kast03} Their results
931 indicated that the damped {\sc sf} method results in properties
932 very similar to those obtained when using the Ewald summation.
933 Considering the statistical results shown above, the good performance
934 of this method is not that surprising. Rather than consider the same
935 systems and simply recapitulate their results, we decided to look at
936 the solid state dynamical behavior obtained using the best performing
937 summation methods from the above results.
938
939 \begin{figure}
940 \centering
941 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
942 \caption{Velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset is a magnification of the first trough. The times to first collision are nearly identical, but the differences can be seen in the peaks and troughs, where the undamped to weakly damped methods are stiffer than the moderately damped and SPME methods.}
943 \label{fig:vCorrPlot}
944 \end{figure}
945
946 The short-time decays through the first collision are nearly identical
947 in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
948 functions show how the methods differ. The undamped {\sc sf} method
949 has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
950 peaks than any of the other methods. As the damping function is
951 increased, these peaks are smoothed out, and approach the SPME
952 curve. The damping acts as a distance dependent Gaussian screening of
953 the point charges for the pairwise summation methods; thus, the
954 collisions are more elastic in the undamped {\sc sf} potental, and the
955 stiffness of the potential is diminished as the electrostatic
956 interactions are softened by the damping function. With $\alpha$
957 values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
958 nearly identical and track the SPME features quite well. This is not
959 too surprising in that the differences between the {\sc sf} and {\sc
960 sp} potentials are mitigated with increased damping. However, this
961 appears to indicate that once damping is utilized, the form of the
962 potential seems to play a lesser role in the crystal dynamics.
963
964 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
965
966 The short time dynamics were extended to evaluate how the differences
967 between the methods affect the collective long-time motion. The same
968 electrostatic summation methods were used as in the short time
969 velocity autocorrelation function evaluation, but the trajectories
970 were sampled over a much longer time. The power spectra of the
971 resulting velocity autocorrelation functions were calculated and are
972 displayed in figure \ref{fig:methodPS}.
973
974 \begin{figure}
975 \centering
976 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
977 \caption{Power spectra obtained from the velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, {\sc sf} ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). Apodization of the correlation functions via a cubic switching function between 40 and 50 ps was used to clear up the spectral noise resulting from data truncation, and had no noticeable effect on peak location or magnitude. The inset shows the frequency region below 100 cm$^{-1}$ to highlight where the spectra begin to differ.}
978 \label{fig:methodPS}
979 \end{figure}
980
981 While high frequency peaks of the spectra in this figure overlap,
982 showing the same general features, the low frequency region shows how
983 the summation methods differ. Considering the low-frequency inset
984 (expanded in the upper frame of figure \ref{fig:dampInc}), at
985 frequencies below 100 cm$^{-1}$, the correlated motions are
986 blue-shifted when using undamped or weakly damped {\sc sf}. When
987 using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
988 and {\sc sp} methods give near identical correlated motion behavior as
989 the Ewald method (which has a damping value of 0.3119). This
990 weakening of the electrostatic interaction with increased damping
991 explains why the long-ranged correlated motions are at lower
992 frequencies for the moderately damped methods than for undamped or
993 weakly damped methods. To see this effect more clearly, we show how
994 damping strength alone affects a simple real-space electrostatic
995 potential,
996 \begin{equation}
997 V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
998 \end{equation}
999 where $S(r)$ is a switching function that smoothly zeroes the
1000 potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
1001 the low frequency motions are dependent on the damping used in the
1002 direct electrostatic sum. As the damping increases, the peaks drop to
1003 lower frequencies. Incidentally, use of an $\alpha$ of 0.25
1004 \AA$^{-1}$ on a simple electrostatic summation results in low
1005 frequency correlated dynamics equivalent to a simulation using SPME.
1006 When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1007 shift to higher frequency in exponential fashion. Though not shown,
1008 the spectrum for the simple undamped electrostatic potential is
1009 blue-shifted such that the lowest frequency peak resides near 325
1010 cm$^{-1}$. In light of these results, the undamped {\sc sf} method
1011 producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1012 respectable and shows that the shifted force procedure accounts for
1013 most of the effect afforded through use of the Ewald summation.
1014 However, it appears as though moderate damping is required for
1015 accurate reproduction of crystal dynamics.
1016 \begin{figure}
1017 \centering
1018 \includegraphics[width = \linewidth]{./comboSquare.pdf}
1019 \caption{Regions of spectra showing the low-frequency correlated motions for NaCl crystals at 1000 K using various electrostatic summation methods. The upper plot is a zoomed inset from figure \ref{fig:methodPS}. As the damping value for the {\sc sf} potential increases, the low-frequency peaks red-shift. The lower plot is of spectra when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$. As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME. The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
1020 \label{fig:dampInc}
1021 \end{figure}
1022
1023 \section{Conclusions}
1024
1025 This investigation of pairwise electrostatic summation techniques
1026 shows that there are viable and more computationally efficient
1027 electrostatic summation techniques than the Ewald summation, chiefly
1028 methods derived from the damped Coulombic sum originally proposed by
1029 Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1030 {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1031 shows a remarkable ability to reproduce the energetic and dynamic
1032 characteristics exhibited by simulations employing lattice summation
1033 techniques. The cumulative energy difference results showed the
1034 undamped {\sc sf} and moderately damped {\sc sp} methods
1035 produced results nearly identical to SPME. Similarly for the dynamic
1036 features, the undamped or moderately damped {\sc sf} and
1037 moderately damped {\sc sp} methods produce force and torque
1038 vector magnitude and directions very similar to the expected values.
1039 These results translate into long-time dynamic behavior equivalent to
1040 that produced in simulations using SPME.
1041
1042 Aside from the computational cost benefit, these techniques have
1043 applicability in situations where the use of the Ewald sum can prove
1044 problematic. Primary among them is their use in interfacial systems,
1045 where the unmodified lattice sum techniques artificially accentuate
1046 the periodicity of the system in an undesirable manner. There have
1047 been alterations to the standard Ewald techniques, via corrections and
1048 reformulations, to compensate for these systems; but the pairwise
1049 techniques discussed here require no modifications, making them
1050 natural tools to tackle these problems. Additionally, this
1051 transferability gives them benefits over other pairwise methods, like
1052 reaction field, because estimations of physical properties (e.g. the
1053 dielectric constant) are unnecessary.
1054
1055 We are not suggesting any flaw with the Ewald sum; in fact, it is the
1056 standard by which these simple pairwise sums are judged. However,
1057 these results do suggest that in the typical simulations performed
1058 today, the Ewald summation may no longer be required to obtain the
1059 level of accuracy most researchers have come to expect
1060
1061 \section{Acknowledgments}
1062 \newpage
1063
1064 \bibliographystyle{jcp2}
1065 \bibliography{electrostaticMethods}
1066
1067
1068 \end{document}