ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2636
Committed: Sat Mar 18 05:31:17 2006 UTC (18 years, 6 months ago) by chrisfen
Content type: application/x-tex
File size: 53582 byte(s)
Log Message:
ewald is a changin'

File Contents

# User Rev Content
1 chrisfen 2575 %\documentclass[prb,aps,twocolumn,tabularx]{revtex4}
2 gezelter 2617 %\documentclass[aps,prb,preprint]{revtex4}
3     \documentclass[11pt]{article}
4 chrisfen 2575 \usepackage{endfloat}
5 chrisfen 2636 \usepackage{amsmath,bm}
6 chrisfen 2594 \usepackage{amssymb}
7 chrisfen 2575 \usepackage{epsf}
8     \usepackage{times}
9 gezelter 2617 \usepackage{mathptmx}
10 chrisfen 2575 \usepackage{setspace}
11     \usepackage{tabularx}
12     \usepackage{graphicx}
13 chrisfen 2595 \usepackage{booktabs}
14 chrisfen 2605 \usepackage{bibentry}
15     \usepackage{mathrsfs}
16 chrisfen 2575 \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 gezelter 2617 \title{Is the Ewald Summation necessary? Pairwise alternatives to the accepted standard for long-range electrostatics}
29 chrisfen 2575
30 gezelter 2617 \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
31     gezelter@nd.edu} \\
32 chrisfen 2575 Department of Chemistry and Biochemistry\\
33     University of Notre Dame\\
34     Notre Dame, Indiana 46556}
35    
36     \date{\today}
37    
38     \maketitle
39 gezelter 2617 \doublespacing
40    
41 chrisfen 2605 \nobibliography{}
42 chrisfen 2575 \begin{abstract}
43 gezelter 2617 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 chrisfen 2619
55 chrisfen 2575 \end{abstract}
56    
57 gezelter 2617 \newpage
58    
59 chrisfen 2575 %\narrowtext
60    
61 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62 chrisfen 2575 % BODY OF TEXT
63 chrisfen 2620 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64 chrisfen 2575
65     \section{Introduction}
66    
67 gezelter 2617 In molecular simulations, proper accumulation of the electrostatic
68     interactions is considered one of the most essential and
69 chrisfen 2620 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 an
81     impractical task to perform these calculations.
82 chrisfen 2604
83 chrisfen 2608 \subsection{The Ewald Sum}
84 chrisfen 2636 The complete accumulation electrostatic interactions in a system with periodic boundary conditions (PBC) requires the consideration of the effect of all charges within a simulation box, as well as those in the periodic replicas,
85     \begin{equation}
86     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],
87     \label{eq:PBCSum}
88     \end{equation}
89     where the sum over $\mathbf{n}$ is a sum over all periodic box replicas
90     with integer coordinates $\mathbf{n} = (l,m,n)$, and the prime indicates
91     $i = j$ are neglected for $\mathbf{n} = 0$.\cite{deLeeuw80} Within the
92     sum, $N$ is the number of electrostatic particles, $\mathbf{r}_{ij}$ is
93     $\mathbf{r}_j - \mathbf{r}_i$, $L$ is the cell length, $\bm{\Omega}_{i,j}$ are
94     the Euler angles for $i$ and $j$, and $\phi$ is Poisson's equation
95     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for charge-charge
96     interactions). In the case of monopole electrostatics,
97     eq. (\ref{eq:PBCSum}) is conditionally convergent and is discontiuous
98     for non-neutral systems.
99 chrisfen 2604
100 chrisfen 2636 This electrostatic summation problem was originally studied by Ewald
101     for the case of an infinite crystal.\cite{Ewald21}. The approach he
102     took was to convert this conditionally convergent sum into two
103     absolutely convergent summations: a short-ranged real-space summation
104     and a long-ranged reciprocal-space summation,
105     \begin{equation}
106     \begin{split}
107     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}{3L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
108     \end{split}
109     \label{eq:EwaldSum}
110     \end{equation}
111     where $\alpha$ is a damping parameter, or separation constant, with
112     units of \AA$^{-1}$, and $\mathbf{k}$ are the reciprocal vectors and
113     equal $2\pi\mathbf{n}/L^2$. The final two terms of
114     eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
115     for interacting with a surrounding dielectric.\cite{Allen87} This
116     dipolar term was neglected in early applications in molecular
117     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
118     Leeuw {\it et al.} to address situations where the unit cell has a
119     dipole moment and this dipole moment gets magnified through
120     replication of the periodic images.\cite{deLeeuw80} This term is zero
121     for systems where $\epsilon_{\rm S} = \infty$. Figure
122     \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
123     time. Initially, due to the small size of systems, the entire
124     simulation box was replicated to convergence. Currently, we balance a
125     spherical real-space cutoff with the reciprocal sum and consider the
126     surrounding dielectric.
127 chrisfen 2610 \begin{figure}
128     \centering
129 gezelter 2617 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
130     \caption{How the application of the Ewald summation has changed with
131     the increase in computer power. Initially, only small numbers of
132     particles could be studied, and the Ewald sum acted to replicate the
133     unit cell charge distribution out to convergence. Now, much larger
134     systems of charges are investigated with fixed distance cutoffs. The
135     calculated structure factor is used to sum out to great distance, and
136     a surrounding dielectric term is included.}
137 chrisfen 2610 \label{fig:ewaldTime}
138     \end{figure}
139    
140 chrisfen 2636 The Ewald summation in the straight-forward form is an
141     $\mathscr{O}(N^2)$ algorithm. The separation constant $(\alpha)$
142     plays an important role in the computational cost balance between the
143     direct and reciprocal-space portions of the summation. The choice of
144     the magnitude of this value allows one to whether the real-space or
145     reciprocal space portion of the summation is an $\mathscr{O}(N^2)$
146     calcualtion, with the other being $\mathscr{O}(N)$.\cite{Sagui99} With
147     appropriate choice of $\alpha$ and thoughtful algorithm development,
148     this cost can be brought down to
149     $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
150     accelerate the Ewald summation is to se
151    
152 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
153 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
154 gezelter 2624 for the accurate accumulation of electrostatic interactions in an
155 gezelter 2617 efficient pairwise fashion.\cite{Wolf99} Wolf \textit{et al.} observed
156     that the electrostatic interaction is effectively short-ranged in
157     condensed phase systems and that neutralization of the charge
158     contained within the cutoff radius is crucial for potential
159     stability. They devised a pairwise summation method that ensures
160     charge neutrality and gives results similar to those obtained with
161     the Ewald summation. The resulting shifted Coulomb potential
162     (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
163     placement on the cutoff sphere and a distance-dependent damping
164     function (identical to that seen in the real-space portion of the
165     Ewald sum) to aid convergence
166 chrisfen 2601 \begin{equation}
167 gezelter 2624 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\}.
168 chrisfen 2601 \label{eq:WolfPot}
169     \end{equation}
170 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
171     potential. However, neutralizing the charge contained within each
172     cutoff sphere requires the placement of a self-image charge on the
173     surface of the cutoff sphere. This additional self-term in the total
174 gezelter 2624 potential enabled Wolf {\it et al.} to obtain excellent estimates of
175 gezelter 2617 Madelung energies for many crystals.
176    
177     In order to use their charge-neutralized potential in molecular
178     dynamics simulations, Wolf \textit{et al.} suggested taking the
179     derivative of this potential prior to evaluation of the limit. This
180     procedure gives an expression for the forces,
181 chrisfen 2601 \begin{equation}
182 chrisfen 2636 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\},
183 chrisfen 2601 \label{eq:WolfForces}
184     \end{equation}
185 gezelter 2617 that incorporates both image charges and damping of the electrostatic
186     interaction.
187    
188     More recently, Zahn \textit{et al.} investigated these potential and
189     force expressions for use in simulations involving water.\cite{Zahn02}
190 gezelter 2624 In their work, they pointed out that the forces and derivative of
191     the potential are not commensurate. Attempts to use both
192     Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
193     to poor energy conservation. They correctly observed that taking the
194     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
195     derivatives gives forces for a different potential energy function
196     than the one shown in Eq. (\ref{eq:WolfPot}).
197 gezelter 2617
198     Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
199     method'' as a way to use this technique in Molecular Dynamics
200     simulations. Taking the integral of the forces shown in equation
201     \ref{eq:WolfForces}, they proposed a new damped Coulomb
202     potential,
203 chrisfen 2601 \begin{equation}
204 gezelter 2624 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\}.
205 chrisfen 2601 \label{eq:ZahnPot}
206     \end{equation}
207 gezelter 2617 They showed that this potential does fairly well at capturing the
208     structural and dynamic properties of water compared the same
209     properties obtained using the Ewald sum.
210 chrisfen 2601
211 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
212    
213 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
214     al.} are constructed using two different (and separable) computational
215 gezelter 2624 tricks: \begin{enumerate}
216 gezelter 2617 \item shifting through the use of image charges, and
217     \item damping the electrostatic interaction.
218 gezelter 2624 \end{enumerate} Wolf \textit{et al.} treated the
219 gezelter 2617 development of their summation method as a progressive application of
220     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
221     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
222     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
223     both techniques. It is possible, however, to separate these
224     tricks and study their effects independently.
225    
226     Starting with the original observation that the effective range of the
227     electrostatic interaction in condensed phases is considerably less
228     than $r^{-1}$, either the cutoff sphere neutralization or the
229     distance-dependent damping technique could be used as a foundation for
230     a new pairwise summation method. Wolf \textit{et al.} made the
231     observation that charge neutralization within the cutoff sphere plays
232     a significant role in energy convergence; therefore we will begin our
233     analysis with the various shifted forms that maintain this charge
234     neutralization. We can evaluate the methods of Wolf
235     \textit{et al.} and Zahn \textit{et al.} by considering the standard
236     shifted potential,
237 chrisfen 2601 \begin{equation}
238 gezelter 2624 v_\textrm{SP}(r) = \begin{cases}
239 gezelter 2617 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
240     R_\textrm{c}
241     \end{cases},
242     \label{eq:shiftingPotForm}
243     \end{equation}
244     and shifted force,
245     \begin{equation}
246 gezelter 2624 v_\textrm{SF}(r) = \begin{cases}
247     v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
248 gezelter 2617 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
249 chrisfen 2601 \end{cases},
250 chrisfen 2612 \label{eq:shiftingForm}
251 chrisfen 2601 \end{equation}
252 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
253     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
254     that both the potential and the forces goes to zero at the cutoff
255     radius, while the Shifted Potential ({\sc sp}) form only ensures the
256     potential is smooth at the cutoff radius
257     ($R_\textrm{c}$).\cite{Allen87}
258    
259 gezelter 2624 The forces associated with the shifted potential are simply the forces
260     of the unshifted potential itself (when inside the cutoff sphere),
261 chrisfen 2601 \begin{equation}
262 chrisfen 2636 f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
263 chrisfen 2612 \end{equation}
264 gezelter 2624 and are zero outside. Inside the cutoff sphere, the forces associated
265     with the shifted force form can be written,
266 chrisfen 2612 \begin{equation}
267 chrisfen 2636 f_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
268 gezelter 2624 v(r)}{dr} \right)_{r=R_\textrm{c}}.
269     \end{equation}
270    
271     If the potential ($v(r)$) is taken to be the normal Coulomb potential,
272     \begin{equation}
273     v(r) = \frac{q_i q_j}{r},
274     \label{eq:Coulomb}
275     \end{equation}
276     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
277     al.}'s undamped prescription:
278     \begin{equation}
279 chrisfen 2636 v_\textrm{SP}(r) =
280 gezelter 2624 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
281     r\leqslant R_\textrm{c},
282 chrisfen 2636 \label{eq:SPPot}
283 gezelter 2624 \end{equation}
284     with associated forces,
285     \begin{equation}
286 chrisfen 2636 f_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
287     \label{eq:SPForces}
288 chrisfen 2612 \end{equation}
289 gezelter 2624 These forces are identical to the forces of the standard Coulomb
290     interaction, and cutting these off at $R_c$ was addressed by Wolf
291     \textit{et al.} as undesirable. They pointed out that the effect of
292     the image charges is neglected in the forces when this form is
293     used,\cite{Wolf99} thereby eliminating any benefit from the method in
294     molecular dynamics. Additionally, there is a discontinuity in the
295     forces at the cutoff radius which results in energy drift during MD
296     simulations.
297 chrisfen 2612
298 gezelter 2624 The shifted force ({\sc sf}) form using the normal Coulomb potential
299     will give,
300 chrisfen 2612 \begin{equation}
301 chrisfen 2636 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}.
302 chrisfen 2612 \label{eq:SFPot}
303     \end{equation}
304 gezelter 2624 with associated forces,
305 chrisfen 2612 \begin{equation}
306 chrisfen 2636 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}.
307 chrisfen 2612 \label{eq:SFForces}
308     \end{equation}
309 gezelter 2624 This formulation has the benefits that there are no discontinuities at
310     the cutoff distance, while the neutralizing image charges are present
311     in both the energy and force expressions. It would be simple to add
312     the self-neutralizing term back when computing the total energy of the
313     system, thereby maintaining the agreement with the Madelung energies.
314     A side effect of this treatment is the alteration in the shape of the
315     potential that comes from the derivative term. Thus, a degree of
316     clarity about agreement with the empirical potential is lost in order
317     to gain functionality in dynamics simulations.
318 chrisfen 2612
319 chrisfen 2620 Wolf \textit{et al.} originally discussed the energetics of the
320 chrisfen 2636 shifted Coulomb potential (Eq. \ref{eq:SPPot}), and they found that
321 gezelter 2624 it was still insufficient for accurate determination of the energy
322     with reasonable cutoff distances. The calculated Madelung energies
323     fluctuate around the expected value with increasing cutoff radius, but
324     the oscillations converge toward the correct value.\cite{Wolf99} A
325     damping function was incorporated to accelerate the convergence; and
326     though alternative functional forms could be
327     used,\cite{Jones56,Heyes81} the complimentary error function was
328     chosen to mirror the effective screening used in the Ewald summation.
329     Incorporating this error function damping into the simple Coulomb
330     potential,
331 chrisfen 2612 \begin{equation}
332 gezelter 2624 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
333 chrisfen 2601 \label{eq:dampCoulomb}
334     \end{equation}
335 chrisfen 2636 the shifted potential (Eq. (\ref{eq:SPPot})) can be reacquired using
336     eq. (\ref{eq:shiftingForm}),
337 chrisfen 2601 \begin{equation}
338 chrisfen 2636 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},
339 chrisfen 2612 \label{eq:DSPPot}
340 chrisfen 2629 \end{equation}
341 gezelter 2624 with associated forces,
342 chrisfen 2612 \begin{equation}
343 chrisfen 2636 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}.
344 chrisfen 2612 \label{eq:DSPForces}
345     \end{equation}
346 gezelter 2624 Again, this damped shifted potential suffers from a discontinuity and
347     a lack of the image charges in the forces. To remedy these concerns,
348 chrisfen 2629 one may derive a {\sc sf} variant by including the derivative
349     term in eq. (\ref{eq:shiftingForm}),
350 chrisfen 2612 \begin{equation}
351 chrisfen 2620 \begin{split}
352 gezelter 2624 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}.
353 chrisfen 2612 \label{eq:DSFPot}
354 chrisfen 2620 \end{split}
355 chrisfen 2612 \end{equation}
356 chrisfen 2636 The derivative of the above potential will lead to the following forces,
357 chrisfen 2612 \begin{equation}
358 chrisfen 2620 \begin{split}
359 chrisfen 2636 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}.
360 chrisfen 2612 \label{eq:DSFForces}
361 chrisfen 2620 \end{split}
362 chrisfen 2612 \end{equation}
363 chrisfen 2636 If the damping parameter $(\alpha)$ is chosen to be zero, the undamped
364     case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
365     from eqs. (\ref{eq:DSPPot}-\ref{eq:DSFForces}).
366 chrisfen 2601
367 chrisfen 2636 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
368     derived by Zahn \textit{et al.}; however, there are two important
369     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
370     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
371     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
372     in the Zahn potential, resulting in a potential discontinuity as
373     particles cross $R_\textrm{c}$. Second, the sign of the derivative
374     portion is different. The missing $v_\textrm{c}$ term would not
375     affect molecular dynamics simulations (although the computed energy
376     would be expected to have sudden jumps as particle distances crossed
377     $R_c$). The sign problem would be a potential source of errors,
378     however. In fact, it introduces a discontinuity in the forces at the
379     cutoff, because the force function is shifted in the wrong direction
380     and doesn't cross zero at $R_\textrm{c}$.
381 chrisfen 2602
382 gezelter 2624 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
383     electrostatic summation method that is continuous in both the
384     potential and forces and which incorporates the damping function
385     proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
386     paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
387     sf}, damping) are at reproducing the correct electrostatic summation
388     performed by the Ewald sum.
389    
390     \subsection{Other alternatives}
391 chrisfen 2629 In addition to the methods described above, we will consider some
392     other techniques that commonly get used in molecular simulations. The
393     simplest of these is group-based cutoffs. Though of little use for
394     non-neutral molecules, collecting atoms into neutral groups takes
395     advantage of the observation that the electrostatic interactions decay
396     faster than those for monopolar pairs.\cite{Steinbach94} When
397     considering these molecules as groups, an orientational aspect is
398     introduced to the interactions. Consequently, as these molecular
399     particles move through $R_\textrm{c}$, the energy will drift upward
400     due to the anisotropy of the net molecular dipole
401     interactions.\cite{Rahman71} To maintain good energy conservation,
402     both the potential and derivative need to be smoothly switched to zero
403     at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
404     switching function,
405     \begin{equation}
406     S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
407     \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} \\
408     0 &\quad r>R_\textrm{c}
409     \end{cases},
410     \end{equation}
411     where the above form is for a cubic function. If a smooth second
412     derivative is desired, a fifth (or higher) order polynomial can be
413     used.\cite{Andrea83}
414 gezelter 2624
415 chrisfen 2629 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
416     and to incorporate their effect, a method like Reaction Field ({\sc
417 chrisfen 2630 rf}) can be used. The original theory for {\sc rf} was originally
418 chrisfen 2629 developed by Onsager,\cite{Onsager36} and it was applied in
419     simulations for the study of water by Barker and Watts.\cite{Barker73}
420     In application, it is simply an extension of the group-based cutoff
421     method where the net dipole within the cutoff sphere polarizes an
422     external dielectric, which reacts back on the central dipole. The
423     same switching function considerations for group-based cutoffs need to
424 chrisfen 2630 made for {\sc rf}, with the additional pre-specification of a
425 chrisfen 2629 dielectric constant.
426 gezelter 2624
427 chrisfen 2608 \section{Methods}
428    
429 chrisfen 2620 In classical molecular mechanics simulations, there are two primary
430     techniques utilized to obtain information about the system of
431     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
432     techniques utilize pairwise summations of interactions between
433     particle sites, but they use these summations in different ways.
434 chrisfen 2608
435 chrisfen 2620 In MC, the potential energy difference between two subsequent
436     configurations dictates the progression of MC sampling. Going back to
437 gezelter 2624 the origins of this method, the acceptance criterion for the canonical
438     ensemble laid out by Metropolis \textit{et al.} states that a
439     subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
440     \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
441     1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
442     alternate method for handling the long-range electrostatics will
443     ensure proper sampling from the ensemble.
444 chrisfen 2608
445 gezelter 2624 In MD, the derivative of the potential governs how the system will
446 chrisfen 2620 progress in time. Consequently, the force and torque vectors on each
447 gezelter 2624 body in the system dictate how the system evolves. If the magnitude
448     and direction of these vectors are similar when using alternate
449     electrostatic summation techniques, the dynamics in the short term
450     will be indistinguishable. Because error in MD calculations is
451     cumulative, one should expect greater deviation at longer times,
452     although methods which have large differences in the force and torque
453     vectors will diverge from each other more rapidly.
454 chrisfen 2608
455 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
456 gezelter 2624 The pairwise summation techniques (outlined in section
457     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
458     studying the energy differences between conformations. We took the
459     SPME-computed energy difference between two conformations to be the
460     correct behavior. An ideal performance by an alternative method would
461     reproduce these energy differences exactly. Since none of the methods
462     provide exact energy differences, we used linear least squares
463     regressions of the $\Delta E$ values between configurations using SPME
464     against $\Delta E$ values using tested methods provides a quantitative
465     comparison of this agreement. Unitary results for both the
466     correlation and correlation coefficient for these regressions indicate
467     equivalent energetic results between the method under consideration
468     and electrostatics handled using SPME. Sample correlation plots for
469     two alternate methods are shown in Fig. \ref{fig:linearFit}.
470 chrisfen 2608
471 chrisfen 2609 \begin{figure}
472     \centering
473 chrisfen 2619 \includegraphics[width = \linewidth]{./dualLinear.pdf}
474     \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.}
475 chrisfen 2609 \label{fig:linearFit}
476     \end{figure}
477    
478 gezelter 2624 Each system type (detailed in section \ref{sec:RepSims}) was
479     represented using 500 independent configurations. Additionally, we
480     used seven different system types, so each of the alternate
481     (non-Ewald) electrostatic summation methods was evaluated using
482     873,250 configurational energy differences.
483 chrisfen 2609
484 gezelter 2624 Results and discussion for the individual analysis of each of the
485     system types appear in the supporting information, while the
486     cumulative results over all the investigated systems appears below in
487     section \ref{sec:EnergyResults}.
488    
489 chrisfen 2609 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
490 gezelter 2624 We evaluated the pairwise methods (outlined in section
491     \ref{sec:ESMethods}) for use in MD simulations by
492     comparing the force and torque vectors with those obtained using the
493     reference Ewald summation (SPME). Both the magnitude and the
494     direction of these vectors on each of the bodies in the system were
495     analyzed. For the magnitude of these vectors, linear least squares
496     regression analyses were performed as described previously for
497     comparing $\Delta E$ values. Instead of a single energy difference
498     between two system configurations, we compared the magnitudes of the
499     forces (and torques) on each molecule in each configuration. For a
500     system of 1000 water molecules and 40 ions, there are 1040 force
501     vectors and 1000 torque vectors. With 500 configurations, this
502     results in 520,000 force and 500,000 torque vector comparisons.
503     Additionally, data from seven different system types was aggregated
504     before the comparison was made.
505 chrisfen 2609
506 gezelter 2624 The {\it directionality} of the force and torque vectors was
507     investigated through measurement of the angle ($\theta$) formed
508     between those computed from the particular method and those from SPME,
509 chrisfen 2610 \begin{equation}
510 gezelter 2624 \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
511 chrisfen 2610 \end{equation}
512 gezelter 2624 where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
513     force vector computed using method $M$.
514    
515 chrisfen 2620 Each of these $\theta$ values was accumulated in a distribution
516 gezelter 2624 function, weighted by the area on the unit sphere. Non-linear
517     Gaussian fits were used to measure the width of the resulting
518     distributions.
519 chrisfen 2609
520     \begin{figure}
521     \centering
522 gezelter 2617 \includegraphics[width = \linewidth]{./gaussFit.pdf}
523 chrisfen 2609 \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.}
524     \label{fig:gaussian}
525     \end{figure}
526    
527 chrisfen 2620 Figure \ref{fig:gaussian} shows an example distribution with applied
528     non-linear fits. The solid line is a Gaussian profile, while the
529     dotted line is a Voigt profile, a convolution of a Gaussian and a
530     Lorentzian. Since this distribution is a measure of angular error
531 gezelter 2624 between two different electrostatic summation methods, there is no
532     {\it a priori} reason for the profile to adhere to any specific shape.
533     Gaussian fits was used to compare all the tested methods. The
534     variance ($\sigma^2$) was extracted from each of these fits and was
535     used to compare distribution widths. Values of $\sigma^2$ near zero
536     indicate vector directions indistinguishable from those calculated
537     when using the reference method (SPME).
538 chrisfen 2609
539 gezelter 2624 \subsection{Short-time Dynamics}
540    
541 chrisfen 2609 \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
542 chrisfen 2620 Evaluation of the long-time dynamics of charged systems was performed
543     by considering the NaCl crystal system while using a subset of the
544     best performing pairwise methods. The NaCl crystal was chosen to
545     avoid possible complications involving the propagation techniques of
546     orientational motion in molecular systems. To enhance the atomic
547     motion, these crystals were equilibrated at 1000 K, near the
548     experimental $T_m$ for NaCl. Simulations were performed under the
549     microcanonical ensemble, and velocity autocorrelation functions
550     (Eq. \ref{eq:vCorr}) were computed for each of the trajectories,
551 chrisfen 2609 \begin{equation}
552     C_v(t) = \langle v_i(0)\cdot v_i(t)\rangle.
553     \label{eq:vCorr}
554     \end{equation}
555 chrisfen 2620 Velocity autocorrelation functions require detailed short time data
556     and long trajectories for good statistics, thus velocity information
557     was saved every 5 fs over 100 ps trajectories. The power spectrum
558     ($I(\omega)$) is obtained via Fourier transform of the autocorrelation
559     function
560 chrisfen 2609 \begin{equation}
561     I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
562     \label{eq:powerSpec}
563     \end{equation}
564     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$.
565    
566     \subsection{Representative Simulations}\label{sec:RepSims}
567 chrisfen 2620 A variety of common and representative simulations were analyzed to
568     determine the relative effectiveness of the pairwise summation
569     techniques in reproducing the energetics and dynamics exhibited by
570     SPME. The studied systems were as follows:
571 chrisfen 2599 \begin{enumerate}
572 chrisfen 2586 \item Liquid Water
573     \item Crystalline Water (Ice I$_\textrm{c}$)
574 chrisfen 2595 \item NaCl Crystal
575     \item NaCl Melt
576 chrisfen 2599 \item Low Ionic Strength Solution of NaCl in Water
577     \item High Ionic Strength Solution of NaCl in Water
578 chrisfen 2586 \item 6 \AA\ Radius Sphere of Argon in Water
579 chrisfen 2599 \end{enumerate}
580 chrisfen 2620 By utilizing the pairwise techniques (outlined in section
581     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
582     charged particles, and mixtures of the two, we can comment on possible
583     system dependence and/or universal applicability of the techniques.
584 chrisfen 2586
585 chrisfen 2620 Generation of the system configurations was dependent on the system
586     type. For the solid and liquid water configurations, configuration
587     snapshots were taken at regular intervals from higher temperature 1000
588     SPC/E water molecule trajectories and each equilibrated individually.
589     The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
590     ions and were selected and equilibrated in the same fashion as the
591     water systems. For the low and high ionic strength NaCl solutions, 4
592     and 40 ions were first solvated in a 1000 water molecule boxes
593     respectively. Ion and water positions were then randomly swapped, and
594     the resulting configurations were again equilibrated individually.
595     Finally, for the Argon/Water "charge void" systems, the identities of
596     all the SPC/E waters within 6 \AA\ of the center of the equilibrated
597     water configurations were converted to argon
598     (Fig. \ref{fig:argonSlice}).
599 chrisfen 2586
600     \begin{figure}
601     \centering
602 gezelter 2617 \includegraphics[width = \linewidth]{./slice.pdf}
603 chrisfen 2586 \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.}
604 chrisfen 2601 \label{fig:argonSlice}
605 chrisfen 2586 \end{figure}
606    
607 chrisfen 2609 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
608 chrisfen 2620 Electrostatic summation method comparisons were performed using SPME,
609 chrisfen 2629 the {\sc sp} and {\sc sf} methods - both with damping
610 chrisfen 2620 parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
611     moderate, and strong damping respectively), reaction field with an
612     infinite dielectric constant, and an unmodified cutoff. Group-based
613     cutoffs with a fifth-order polynomial switching function were
614     necessary for the reaction field simulations and were utilized in the
615     SP, SF, and pure cutoff methods for comparison to the standard lack of
616     group-based cutoffs with a hard truncation. The SPME calculations
617     were performed using the TINKER implementation of SPME,\cite{Ponder87}
618     while all other method calculations were performed using the OOPSE
619     molecular mechanics package.\cite{Meineke05}
620 chrisfen 2586
621 chrisfen 2620 These methods were additionally evaluated with three different cutoff
622     radii (9, 12, and 15 \AA) to investigate possible cutoff radius
623     dependence. It should be noted that the damping parameter chosen in
624     SPME, or so called ``Ewald Coefficient", has a significant effect on
625     the energies and forces calculated. Typical molecular mechanics
626     packages default this to a value dependent on the cutoff radius and a
627     tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller
628 chrisfen 2636 tolerances are typically associated with increased accuracy, but this
629     usually means more time spent calculating the reciprocal-space portion
630     of the summation.\cite{Perram88,Essmann95} The default TINKER
631     tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
632 chrisfen 2620 calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
633     0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
634 chrisfen 2609
635 chrisfen 2575 \section{Results and Discussion}
636    
637 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
638 chrisfen 2620 In order to evaluate the performance of the pairwise electrostatic
639     summation methods for Monte Carlo simulations, the energy differences
640     between configurations were compared to the values obtained when using
641     SPME. The results for the subsequent regression analysis are shown in
642     figure \ref{fig:delE}.
643 chrisfen 2590
644     \begin{figure}
645     \centering
646 gezelter 2617 \includegraphics[width=5.5in]{./delEplot.pdf}
647 chrisfen 2608 \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).}
648 chrisfen 2601 \label{fig:delE}
649 chrisfen 2594 \end{figure}
650    
651 chrisfen 2620 In this figure, it is apparent that it is unreasonable to expect
652     realistic results using an unmodified cutoff. This is not all that
653     surprising since this results in large energy fluctuations as atoms
654     move in and out of the cutoff radius. These fluctuations can be
655     alleviated to some degree by using group based cutoffs with a
656     switching function.\cite{Steinbach94} The Group Switch Cutoff row
657     doesn't show a significant improvement in this plot because the salt
658     and salt solution systems contain non-neutral groups, see the
659     accompanying supporting information for a comparison where all groups
660     are neutral.
661 chrisfen 2594
662 chrisfen 2620 Correcting the resulting charged cutoff sphere is one of the purposes
663     of the damped Coulomb summation proposed by Wolf \textit{et
664     al.},\cite{Wolf99} and this correction indeed improves the results as
665     seen in the Shifted-Potental rows. While the undamped case of this
666     method is a significant improvement over the pure cutoff, it still
667     doesn't correlate that well with SPME. Inclusion of potential damping
668     improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
669     an excellent correlation and quality of fit with the SPME results,
670     particularly with a cutoff radius greater than 12 \AA . Use of a
671     larger damping parameter is more helpful for the shortest cutoff
672     shown, but it has a detrimental effect on simulations with larger
673 chrisfen 2629 cutoffs. In the {\sc sf} sets, increasing damping results in
674 chrisfen 2620 progressively poorer correlation. Overall, the undamped case is the
675     best performing set, as the correlation and quality of fits are
676     consistently superior regardless of the cutoff distance. This result
677     is beneficial in that the undamped case is less computationally
678     prohibitive do to the lack of complimentary error function calculation
679     when performing the electrostatic pair interaction. The reaction
680     field results illustrates some of that method's limitations, primarily
681     that it was developed for use in homogenous systems; although it does
682     provide results that are an improvement over those from an unmodified
683     cutoff.
684 chrisfen 2609
685 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
686 chrisfen 2599
687 chrisfen 2620 Evaluation of pairwise methods for use in Molecular Dynamics
688     simulations requires consideration of effects on the forces and
689     torques. Investigation of the force and torque vector magnitudes
690     provides a measure of the strength of these values relative to SPME.
691     Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
692     force and torque vector magnitude regression results for the
693     accumulated analysis over all the system types.
694 chrisfen 2594
695     \begin{figure}
696     \centering
697 gezelter 2617 \includegraphics[width=5.5in]{./frcMagplot.pdf}
698 chrisfen 2608 \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).}
699 chrisfen 2601 \label{fig:frcMag}
700 chrisfen 2594 \end{figure}
701    
702 chrisfen 2620 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
703     in the previous $\Delta E$ section. The unmodified cutoff results are
704     poor, but using group based cutoffs and a switching function provides
705     a improvement much more significant than what was seen with $\Delta
706 chrisfen 2629 E$. Looking at the {\sc sp} sets, the slope and $R^2$
707 chrisfen 2620 improve with the use of damping to an optimal result of 0.2 \AA
708     $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping,
709     while beneficial for simulations with a cutoff radius of 9 \AA\ , is
710     detrimental to simulations with larger cutoff radii. The undamped
711 chrisfen 2629 {\sc sf} method gives forces in line with those obtained using
712 chrisfen 2620 SPME, and use of a damping function results in minor improvement. The
713     reaction field results are surprisingly good, considering the poor
714     quality of the fits for the $\Delta E$ results. There is still a
715     considerable degree of scatter in the data, but it correlates well in
716     general. To be fair, we again note that the reaction field
717     calculations do not encompass NaCl crystal and melt systems, so these
718     results are partly biased towards conditions in which the method
719     performs more favorably.
720 chrisfen 2594
721     \begin{figure}
722     \centering
723 gezelter 2617 \includegraphics[width=5.5in]{./trqMagplot.pdf}
724 chrisfen 2608 \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).}
725 chrisfen 2601 \label{fig:trqMag}
726 chrisfen 2594 \end{figure}
727    
728 chrisfen 2620 To evaluate the torque vector magnitudes, the data set from which
729     values are drawn is limited to rigid molecules in the systems
730     (i.e. water molecules). In spite of this smaller sampling pool, the
731     torque vector magnitude results in figure \ref{fig:trqMag} are still
732     similar to those seen for the forces; however, they more clearly show
733     the improved behavior that comes with increasing the cutoff radius.
734 chrisfen 2629 Moderate damping is beneficial to the {\sc sp} and helpful
735     yet possibly unnecessary with the {\sc sf} method, and they also
736 chrisfen 2620 show that over-damping adversely effects all cutoff radii rather than
737     showing an improvement for systems with short cutoffs. The reaction
738     field method performs well when calculating the torques, better than
739     the Shifted Force method over this limited data set.
740 chrisfen 2594
741 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
742 chrisfen 2599
743 chrisfen 2620 Having force and torque vectors with magnitudes that are well
744     correlated to SPME is good, but if they are not pointing in the proper
745     direction the results will be incorrect. These vector directions were
746     investigated through measurement of the angle formed between them and
747     those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared
748     through the variance ($\sigma^2$) of the Gaussian fits of the angle
749     error distributions of the combined set over all system types.
750 chrisfen 2594
751     \begin{figure}
752     \centering
753 gezelter 2617 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
754 chrisfen 2608 \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).}
755 chrisfen 2601 \label{fig:frcTrqAng}
756 chrisfen 2594 \end{figure}
757    
758 chrisfen 2620 Both the force and torque $\sigma^2$ results from the analysis of the
759     total accumulated system data are tabulated in figure
760     \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case
761     show the improvement afforded by choosing a longer simulation cutoff.
762     Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
763     of the distribution widths, with a similar improvement going from 12
764 chrisfen 2629 to 15 \AA . The undamped {\sc sf}, Group Based Cutoff, and
765 chrisfen 2620 Reaction Field methods all do equivalently well at capturing the
766     direction of both the force and torque vectors. Using damping
767 chrisfen 2629 improves the angular behavior significantly for the {\sc sp}
768     and moderately for the {\sc sf} methods. Increasing the damping
769 chrisfen 2620 too far is destructive for both methods, particularly to the torque
770     vectors. Again it is important to recognize that the force vectors
771     cover all particles in the systems, while torque vectors are only
772     available for neutral molecular groups. Damping appears to have a
773     more beneficial effect on non-neutral bodies, and this observation is
774     investigated further in the accompanying supporting information.
775 chrisfen 2594
776 chrisfen 2595 \begin{table}[htbp]
777     \centering
778     \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$^2$.}
779 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
780 chrisfen 2595 \\
781     \toprule
782     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
783     \cmidrule(lr){3-6}
784     \cmidrule(l){7-10}
785 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
786 chrisfen 2595 \midrule
787 chrisfen 2599
788     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
789     & \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} \\
790     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
791     & \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} \\
792     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
793     & \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} \\
794 chrisfen 2594
795 chrisfen 2595 \midrule
796    
797 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
798     & \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} \\
799     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
800     & \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} \\
801     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
802     & \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} \\
803 chrisfen 2595
804     \bottomrule
805     \end{tabular}
806 chrisfen 2601 \label{tab:groupAngle}
807 chrisfen 2595 \end{table}
808    
809 chrisfen 2620 Although not discussed previously, group based cutoffs can be applied
810 chrisfen 2629 to both the {\sc sp} and {\sc sf} methods. Use off a
811 chrisfen 2620 switching function corrects for the discontinuities that arise when
812     atoms of a group exit the cutoff before the group's center of mass.
813     Though there are no significant benefit or drawbacks observed in
814     $\Delta E$ and vector magnitude results when doing this, there is a
815     measurable improvement in the vector angle results. Table
816     \ref{tab:groupAngle} shows the angular variance values obtained using
817     group based cutoffs and a switching function alongside the standard
818     results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
819 chrisfen 2629 The {\sc sp} shows much narrower angular distributions for
820 chrisfen 2620 both the force and torque vectors when using an $\alpha$ of 0.2
821 chrisfen 2629 \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
822 chrisfen 2620 undamped and lightly damped cases. Thus, by calculating the
823     electrostatic interactions in terms of molecular pairs rather than
824     atomic pairs, the direction of the force and torque vectors are
825     determined more accurately.
826 chrisfen 2595
827 chrisfen 2620 One additional trend to recognize in table \ref{tab:groupAngle} is
828 chrisfen 2629 that the $\sigma^2$ values for both {\sc sp} and
829     {\sc sf} converge as $\alpha$ increases, something that is easier
830 chrisfen 2620 to see when using group based cutoffs. Looking back on figures
831     \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
832     behavior clearly at large $\alpha$ and cutoff values. The reason for
833     this is that the complimentary error function inserted into the
834     potential weakens the electrostatic interaction as $\alpha$ increases.
835     Thus, at larger values of $\alpha$, both the summation method types
836     progress toward non-interacting functions, so care is required in
837     choosing large damping functions lest one generate an undesirable loss
838     in the pair interaction. Kast \textit{et al.} developed a method for
839     choosing appropriate $\alpha$ values for these types of electrostatic
840     summation methods by fitting to $g(r)$ data, and their methods
841     indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
842     values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
843     to be reasonable choices to obtain proper MC behavior
844     (Fig. \ref{fig:delE}); however, based on these findings, choices this
845     high would introduce error in the molecular torques, particularly for
846     the shorter cutoffs. Based on the above findings, empirical damping
847     up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
848 chrisfen 2629 unnecessary when using the {\sc sf} method.
849 chrisfen 2595
850 chrisfen 2608 \subsection{Collective Motion: Power Spectra of NaCl Crystals}
851 chrisfen 2601
852 chrisfen 2629 In the previous studies using a {\sc sf} variant of the damped
853 chrisfen 2620 Wolf coulomb potential, the structure and dynamics of water were
854     investigated rather extensively.\cite{Zahn02,Kast03} Their results
855 chrisfen 2629 indicated that the damped {\sc sf} method results in properties
856 chrisfen 2620 very similar to those obtained when using the Ewald summation.
857     Considering the statistical results shown above, the good performance
858     of this method is not that surprising. Rather than consider the same
859     systems and simply recapitulate their results, we decided to look at
860     the solid state dynamical behavior obtained using the best performing
861     summation methods from the above results.
862 chrisfen 2601
863     \begin{figure}
864     \centering
865 gezelter 2617 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
866 chrisfen 2629 \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.}
867 chrisfen 2610 \label{fig:methodPS}
868 chrisfen 2601 \end{figure}
869    
870 chrisfen 2620 Figure \ref{fig:methodPS} shows the power spectra for the NaCl
871     crystals (from averaged Na and Cl ion velocity autocorrelation
872     functions) using the stated electrostatic summation methods. While
873     high frequency peaks of all the spectra overlap, showing the same
874     general features, the low frequency region shows how the summation
875     methods differ. Considering the low-frequency inset (expanded in the
876     upper frame of figure \ref{fig:dampInc}), at frequencies below 100
877     cm$^{-1}$, the correlated motions are blue-shifted when using undamped
878 chrisfen 2629 or weakly damped {\sc sf}. When using moderate damping ($\alpha
879     = 0.2$ \AA$^{-1}$) both the {\sc sf} and {\sc sp}
880 chrisfen 2620 methods give near identical correlated motion behavior as the Ewald
881     method (which has a damping value of 0.3119). The damping acts as a
882     distance dependent Gaussian screening of the point charges for the
883     pairwise summation methods. This weakening of the electrostatic
884     interaction with distance explains why the long-ranged correlated
885     motions are at lower frequencies for the moderately damped methods
886     than for undamped or weakly damped methods. To see this effect more
887     clearly, we show how damping strength affects a simple real-space
888     electrostatic potential,
889 chrisfen 2601 \begin{equation}
890 gezelter 2624 V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
891 chrisfen 2601 \end{equation}
892 chrisfen 2620 where $S(r)$ is a switching function that smoothly zeroes the
893     potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
894     the low frequency motions are dependent on the damping used in the
895     direct electrostatic sum. As the damping increases, the peaks drop to
896     lower frequencies. Incidentally, use of an $\alpha$ of 0.25
897     \AA$^{-1}$ on a simple electrostatic summation results in low
898     frequency correlated dynamics equivalent to a simulation using SPME.
899     When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
900     shift to higher frequency in exponential fashion. Though not shown,
901     the spectrum for the simple undamped electrostatic potential is
902     blue-shifted such that the lowest frequency peak resides near 325
903 chrisfen 2629 cm$^{-1}$. In light of these results, the undamped {\sc sf}
904 chrisfen 2620 method producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is
905     quite respectable; however, it appears as though moderate damping is
906     required for accurate reproduction of crystal dynamics.
907 chrisfen 2601 \begin{figure}
908     \centering
909 gezelter 2617 \includegraphics[width = \linewidth]{./comboSquare.pdf}
910 chrisfen 2636 \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.}
911 chrisfen 2601 \label{fig:dampInc}
912     \end{figure}
913    
914 chrisfen 2575 \section{Conclusions}
915    
916 chrisfen 2620 This investigation of pairwise electrostatic summation techniques
917     shows that there are viable and more computationally efficient
918     electrostatic summation techniques than the Ewald summation, chiefly
919     methods derived from the damped Coulombic sum originally proposed by
920     Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
921 chrisfen 2629 {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
922 chrisfen 2620 shows a remarkable ability to reproduce the energetic and dynamic
923     characteristics exhibited by simulations employing lattice summation
924     techniques. The cumulative energy difference results showed the
925 chrisfen 2629 undamped {\sc sf} and moderately damped {\sc sp} methods
926 chrisfen 2620 produced results nearly identical to SPME. Similarly for the dynamic
927 chrisfen 2629 features, the undamped or moderately damped {\sc sf} and
928     moderately damped {\sc sp} methods produce force and torque
929 chrisfen 2620 vector magnitude and directions very similar to the expected values.
930     These results translate into long-time dynamic behavior equivalent to
931     that produced in simulations using SPME.
932 chrisfen 2604
933 chrisfen 2620 Aside from the computational cost benefit, these techniques have
934     applicability in situations where the use of the Ewald sum can prove
935     problematic. Primary among them is their use in interfacial systems,
936     where the unmodified lattice sum techniques artificially accentuate
937     the periodicity of the system in an undesirable manner. There have
938     been alterations to the standard Ewald techniques, via corrections and
939     reformulations, to compensate for these systems; but the pairwise
940     techniques discussed here require no modifications, making them
941     natural tools to tackle these problems. Additionally, this
942     transferability gives them benefits over other pairwise methods, like
943     reaction field, because estimations of physical properties (e.g. the
944     dielectric constant) are unnecessary.
945 chrisfen 2605
946 chrisfen 2620 We are not suggesting any flaw with the Ewald sum; in fact, it is the
947     standard by which these simple pairwise sums are judged. However,
948     these results do suggest that in the typical simulations performed
949     today, the Ewald summation may no longer be required to obtain the
950     level of accuracy most researcher have come to expect
951 chrisfen 2605
952 chrisfen 2575 \section{Acknowledgments}
953 chrisfen 2594 \newpage
954    
955 gezelter 2617 \bibliographystyle{jcp2}
956 chrisfen 2575 \bibliography{electrostaticMethods}
957    
958    
959     \end{document}