ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2643
Committed: Mon Mar 20 17:32:33 2006 UTC (18 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 60066 byte(s)
Log Message:
intro

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 gezelter 2643 interactions is essential and is one of the most
69     computationally-demanding tasks. The common molecular mechanics force
70     fields represent atomic sites with full or partial charges protected
71     by Lennard-Jones (short range) interactions. This means that nearly
72     every pair interaction involves a calculation of charge-charge forces.
73     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
74     interactions quickly become the most expensive part of molecular
75     simulations. Historically, the electrostatic pair interaction would
76     not have decayed appreciably within the typical box lengths that could
77     be feasibly simulated. In the larger systems that are more typical of
78     modern simulations, large cutoffs should be used to incorporate
79     electrostatics correctly.
80 chrisfen 2604
81 gezelter 2643 There have been many efforts to address the proper and practical
82     handling of electrostatic interactions, and these have resulted in a
83     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
84     typically classified as implicit methods (i.e., continuum dielectrics,
85     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
86     (i.e., Ewald summations, interaction shifting or
87 chrisfen 2640 truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
88 chrisfen 2639 reaction field type methods, fast multipole
89     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
90 gezelter 2643 often preferred because they physically incorporate solvent molecules
91     in the system of interest, but these methods are sometimes difficult
92     to utilize because of their high computational cost.\cite{Roux99} In
93     addition to the computational cost, there have been some questions
94     regarding possible artifacts caused by the inherent periodicity of the
95     explicit Ewald summation.\cite{Tobias01}
96 chrisfen 2639
97 gezelter 2643 In this paper, we focus on a new set of shifted methods devised by
98     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
99     methods along with a few other mixed methods (i.e. reaction field) are
100     compared with the smooth particle mesh Ewald
101     sum,\cite{Onsager36,Essmann99} which is our reference method for
102     handling long-range electrostatic interactions. The new methods for
103     handling electrostatics have the potential to scale linearly with
104     increasing system size since they involve only a simple modification
105     to the direct pairwise sum. They also lack the added periodicity of
106     the Ewald sum, so they can be used for systems which are non-periodic
107     or which have one- or two-dimensional periodicity. Below, these
108     methods are evaluated using a variety of model systems to establish
109 chrisfen 2640 their usability in molecular simulations.
110 chrisfen 2639
111 chrisfen 2608 \subsection{The Ewald Sum}
112 chrisfen 2639 The complete accumulation electrostatic interactions in a system with
113     periodic boundary conditions (PBC) requires the consideration of the
114 gezelter 2643 effect of all charges within a (cubic) simulation box as well as those
115     in the periodic replicas,
116 chrisfen 2636 \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 chrisfen 2639 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 gezelter 2643 $j$, and $\phi$ is the solution to Poisson's equation
127     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
128     charge-charge interactions). In the case of monopole electrostatics,
129     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
130     non-neutral systems.
131 chrisfen 2604
132 gezelter 2643 The electrostatic summation problem was originally studied by Ewald
133 chrisfen 2636 for the case of an infinite crystal.\cite{Ewald21}. The approach he
134     took was to convert this conditionally convergent sum into two
135     absolutely convergent summations: a short-ranged real-space summation
136     and a long-ranged reciprocal-space summation,
137     \begin{equation}
138     \begin{split}
139 chrisfen 2637 V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}{|\mathbf{r}_{ij}+\mathbf{n}|} \\ &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2} \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right) \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\ &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2 + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}\left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
140 chrisfen 2636 \end{split}
141     \label{eq:EwaldSum}
142     \end{equation}
143     where $\alpha$ is a damping parameter, or separation constant, with
144 gezelter 2643 units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are
145     equal to $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the
146     dielectric constant of the surrounding medium. The final two terms of
147 chrisfen 2636 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
148     for interacting with a surrounding dielectric.\cite{Allen87} This
149     dipolar term was neglected in early applications in molecular
150     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
151     Leeuw {\it et al.} to address situations where the unit cell has a
152 gezelter 2643 dipole moment which is magnified through replication of the periodic
153     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
154     system is said to be using conducting (or ``tin-foil'') boundary
155 chrisfen 2637 conditions, $\epsilon_{\rm S} = \infty$. Figure
156 chrisfen 2636 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
157 gezelter 2643 time. Initially, due to the small sizes of the systems that could be
158     feasibly simulated, the entire simulation box was replicated to
159     convergence. In more modern simulations, the simulation boxes have
160     grown large enough that a real-space cutoff could potentially give
161     convergent behavior. Indeed, it has often been observed that the
162     reciprocal-space portion of the Ewald sum can be vanishingly
163     small compared to the real-space portion.\cite{XXX}
164    
165 chrisfen 2610 \begin{figure}
166     \centering
167 gezelter 2617 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
168     \caption{How the application of the Ewald summation has changed with
169     the increase in computer power. Initially, only small numbers of
170     particles could be studied, and the Ewald sum acted to replicate the
171     unit cell charge distribution out to convergence. Now, much larger
172     systems of charges are investigated with fixed distance cutoffs. The
173     calculated structure factor is used to sum out to great distance, and
174     a surrounding dielectric term is included.}
175 chrisfen 2610 \label{fig:ewaldTime}
176     \end{figure}
177    
178 gezelter 2643 The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
179     separation constant $(\alpha)$ plays an important role in balancing
180     the computational cost between the direct and reciprocal-space
181     portions of the summation. The choice of this value allows one to
182     select whether the real-space or reciprocal space portion of the
183     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
184     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
185     $\alpha$ and thoughtful algorithm development, this cost can be
186     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
187     taken to reduce the cost of the Ewald summation even further is to set
188     $\alpha$ such that the real-space interactions decay rapidly, allowing
189     for a short spherical cutoff. Then the reciprocal space summation is
190     optimized. These optimizations usually involve utilization of the
191     fast Fourier transform (FFT),\cite{Hockney81} leading to the
192 chrisfen 2637 particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
193     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
194     methods, the cost of the reciprocal-space portion of the Ewald
195 gezelter 2643 summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
196     \log N)$.
197 chrisfen 2636
198 gezelter 2643 These developments and optimizations have made the use of the Ewald
199     summation routine in simulations with periodic boundary
200     conditions. However, in certain systems, such as vapor-liquid
201     interfaces and membranes, the intrinsic three-dimensional periodicity
202     can prove problematic. The Ewald sum has been reformulated to handle
203     2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
204     new methods are computationally expensive.\cite{Spohr97,Yeh99}
205     Inclusion of a correction term in the Ewald summation is a possible
206     direction for handling 2D systems while still enabling the use of the
207     modern optimizations.\cite{Yeh99}
208 chrisfen 2637
209     Several studies have recognized that the inherent periodicity in the
210 gezelter 2643 Ewald sum can also have an effect on three-dimensional
211     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
212     Solvated proteins are essentially kept at high concentration due to
213     the periodicity of the electrostatic summation method. In these
214 chrisfen 2637 systems, the more compact folded states of a protein can be
215     artificially stabilized by the periodic replicas introduced by the
216 gezelter 2643 Ewald summation.\cite{Weber00} Thus, care must be taken when
217     considering the use of the Ewald summation where the assumed
218     periodicity would introduce spurious effects in the system dynamics.
219 chrisfen 2637
220 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
221 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
222 gezelter 2624 for the accurate accumulation of electrostatic interactions in an
223 gezelter 2643 efficient pairwise fashion. This procedure lacks the inherent
224     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
225     observed that the electrostatic interaction is effectively
226     short-ranged in condensed phase systems and that neutralization of the
227     charge contained within the cutoff radius is crucial for potential
228     stability. They devised a pairwise summation method that ensures
229     charge neutrality and gives results similar to those obtained with the
230     Ewald summation. The resulting shifted Coulomb potential
231     (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
232     placement on the cutoff sphere and a distance-dependent damping
233     function (identical to that seen in the real-space portion of the
234     Ewald sum) to aid convergence
235 chrisfen 2601 \begin{equation}
236 chrisfen 2640 V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
237 chrisfen 2601 \label{eq:WolfPot}
238     \end{equation}
239 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
240     potential. However, neutralizing the charge contained within each
241     cutoff sphere requires the placement of a self-image charge on the
242     surface of the cutoff sphere. This additional self-term in the total
243 gezelter 2624 potential enabled Wolf {\it et al.} to obtain excellent estimates of
244 gezelter 2617 Madelung energies for many crystals.
245    
246     In order to use their charge-neutralized potential in molecular
247     dynamics simulations, Wolf \textit{et al.} suggested taking the
248     derivative of this potential prior to evaluation of the limit. This
249     procedure gives an expression for the forces,
250 chrisfen 2601 \begin{equation}
251 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\},
252 chrisfen 2601 \label{eq:WolfForces}
253     \end{equation}
254 gezelter 2617 that incorporates both image charges and damping of the electrostatic
255     interaction.
256    
257     More recently, Zahn \textit{et al.} investigated these potential and
258     force expressions for use in simulations involving water.\cite{Zahn02}
259 gezelter 2624 In their work, they pointed out that the forces and derivative of
260     the potential are not commensurate. Attempts to use both
261 gezelter 2643 eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
262 gezelter 2624 to poor energy conservation. They correctly observed that taking the
263     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
264     derivatives gives forces for a different potential energy function
265 gezelter 2643 than the one shown in eq. (\ref{eq:WolfPot}).
266 gezelter 2617
267 gezelter 2643 Zahn \textit{et al.} introduced a modified form of this summation
268     method as a way to use the technique in Molecular Dynamics
269     simulations. They proposed a new damped Coulomb potential,
270 chrisfen 2601 \begin{equation}
271 gezelter 2643 V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\left\{\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}-\left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right]\left(r_{ij}-R_\mathrm{c}\right)\right\},
272 chrisfen 2601 \label{eq:ZahnPot}
273     \end{equation}
274 gezelter 2643 and showed that this potential does fairly well at capturing the
275 gezelter 2617 structural and dynamic properties of water compared the same
276     properties obtained using the Ewald sum.
277 chrisfen 2601
278 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
279    
280 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
281     al.} are constructed using two different (and separable) computational
282 gezelter 2624 tricks: \begin{enumerate}
283 gezelter 2617 \item shifting through the use of image charges, and
284     \item damping the electrostatic interaction.
285 gezelter 2624 \end{enumerate} Wolf \textit{et al.} treated the
286 gezelter 2617 development of their summation method as a progressive application of
287     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
288     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
289     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
290     both techniques. It is possible, however, to separate these
291     tricks and study their effects independently.
292    
293     Starting with the original observation that the effective range of the
294     electrostatic interaction in condensed phases is considerably less
295     than $r^{-1}$, either the cutoff sphere neutralization or the
296     distance-dependent damping technique could be used as a foundation for
297     a new pairwise summation method. Wolf \textit{et al.} made the
298     observation that charge neutralization within the cutoff sphere plays
299     a significant role in energy convergence; therefore we will begin our
300     analysis with the various shifted forms that maintain this charge
301     neutralization. We can evaluate the methods of Wolf
302     \textit{et al.} and Zahn \textit{et al.} by considering the standard
303     shifted potential,
304 chrisfen 2601 \begin{equation}
305 gezelter 2643 V_\textrm{SP}(r) = \begin{cases}
306 gezelter 2617 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
307     R_\textrm{c}
308     \end{cases},
309     \label{eq:shiftingPotForm}
310     \end{equation}
311     and shifted force,
312     \begin{equation}
313 gezelter 2643 V_\textrm{SF}(r) = \begin{cases}
314 gezelter 2624 v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
315 gezelter 2617 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
316 chrisfen 2601 \end{cases},
317 chrisfen 2612 \label{eq:shiftingForm}
318 chrisfen 2601 \end{equation}
319 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
320     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
321     that both the potential and the forces goes to zero at the cutoff
322     radius, while the Shifted Potential ({\sc sp}) form only ensures the
323     potential is smooth at the cutoff radius
324     ($R_\textrm{c}$).\cite{Allen87}
325    
326 gezelter 2624 The forces associated with the shifted potential are simply the forces
327     of the unshifted potential itself (when inside the cutoff sphere),
328 chrisfen 2601 \begin{equation}
329 gezelter 2643 F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
330 chrisfen 2612 \end{equation}
331 gezelter 2624 and are zero outside. Inside the cutoff sphere, the forces associated
332     with the shifted force form can be written,
333 chrisfen 2612 \begin{equation}
334 gezelter 2643 F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
335 gezelter 2624 v(r)}{dr} \right)_{r=R_\textrm{c}}.
336     \end{equation}
337    
338 gezelter 2643 If the potential, $v(r)$, is taken to be the normal Coulomb potential,
339 gezelter 2624 \begin{equation}
340     v(r) = \frac{q_i q_j}{r},
341     \label{eq:Coulomb}
342     \end{equation}
343     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
344     al.}'s undamped prescription:
345     \begin{equation}
346 gezelter 2643 V_\textrm{SP}(r) =
347 gezelter 2624 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
348     r\leqslant R_\textrm{c},
349 chrisfen 2636 \label{eq:SPPot}
350 gezelter 2624 \end{equation}
351     with associated forces,
352     \begin{equation}
353 gezelter 2643 F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
354 chrisfen 2636 \label{eq:SPForces}
355 chrisfen 2612 \end{equation}
356 gezelter 2624 These forces are identical to the forces of the standard Coulomb
357     interaction, and cutting these off at $R_c$ was addressed by Wolf
358     \textit{et al.} as undesirable. They pointed out that the effect of
359     the image charges is neglected in the forces when this form is
360     used,\cite{Wolf99} thereby eliminating any benefit from the method in
361     molecular dynamics. Additionally, there is a discontinuity in the
362     forces at the cutoff radius which results in energy drift during MD
363     simulations.
364 chrisfen 2612
365 gezelter 2624 The shifted force ({\sc sf}) form using the normal Coulomb potential
366     will give,
367 chrisfen 2612 \begin{equation}
368 gezelter 2643 V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}+\left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right] \quad r\leqslant R_\textrm{c}.
369 chrisfen 2612 \label{eq:SFPot}
370     \end{equation}
371 gezelter 2624 with associated forces,
372 chrisfen 2612 \begin{equation}
373 gezelter 2643 F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right) \quad r\leqslant R_\textrm{c}.
374 chrisfen 2612 \label{eq:SFForces}
375     \end{equation}
376 gezelter 2624 This formulation has the benefits that there are no discontinuities at
377 gezelter 2643 the cutoff radius, while the neutralizing image charges are present in
378     both the energy and force expressions. It would be simple to add the
379     self-neutralizing term back when computing the total energy of the
380 gezelter 2624 system, thereby maintaining the agreement with the Madelung energies.
381     A side effect of this treatment is the alteration in the shape of the
382     potential that comes from the derivative term. Thus, a degree of
383     clarity about agreement with the empirical potential is lost in order
384     to gain functionality in dynamics simulations.
385 chrisfen 2612
386 chrisfen 2620 Wolf \textit{et al.} originally discussed the energetics of the
387 gezelter 2643 shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
388     insufficient for accurate determination of the energy with reasonable
389     cutoff distances. The calculated Madelung energies fluctuated around
390     the expected value as the cutoff radius was increased, but the
391     oscillations converged toward the correct value.\cite{Wolf99} A
392 gezelter 2624 damping function was incorporated to accelerate the convergence; and
393 gezelter 2643 though alternative forms for the damping function could be
394 gezelter 2624 used,\cite{Jones56,Heyes81} the complimentary error function was
395     chosen to mirror the effective screening used in the Ewald summation.
396     Incorporating this error function damping into the simple Coulomb
397     potential,
398 chrisfen 2612 \begin{equation}
399 gezelter 2624 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
400 chrisfen 2601 \label{eq:dampCoulomb}
401     \end{equation}
402 gezelter 2643 the shifted potential (eq. (\ref{eq:SPPot})) becomes
403 chrisfen 2601 \begin{equation}
404 gezelter 2643 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 chrisfen 2612 \label{eq:DSPPot}
406 chrisfen 2629 \end{equation}
407 gezelter 2624 with associated forces,
408 chrisfen 2612 \begin{equation}
409 gezelter 2643 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 chrisfen 2612 \label{eq:DSPForces}
411     \end{equation}
412 gezelter 2643 Again, this damped shifted potential suffers from a
413     force-discontinuity at the cutoff radius, and the image charges play
414     no role in the forces. To remedy these concerns, one may derive a
415     {\sc sf} variant by including the derivative term in
416     eq. (\ref{eq:shiftingForm}),
417 chrisfen 2612 \begin{equation}
418 chrisfen 2620 \begin{split}
419 gezelter 2643 V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\frac{\mathrm{erfc}\left(\alpha r\right)}{r}-\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\ &\left.+\left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}\right)\left(r-R_\mathrm{c}\right)\ \right] \quad r\leqslant R_\textrm{c}.
420 chrisfen 2612 \label{eq:DSFPot}
421 chrisfen 2620 \end{split}
422 chrisfen 2612 \end{equation}
423 chrisfen 2636 The derivative of the above potential will lead to the following forces,
424 chrisfen 2612 \begin{equation}
425 chrisfen 2620 \begin{split}
426 gezelter 2643 F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\ &\left.-\left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}+\frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}\right)\right] \quad r\leqslant R_\textrm{c}.
427 chrisfen 2612 \label{eq:DSFForces}
428 chrisfen 2620 \end{split}
429 chrisfen 2612 \end{equation}
430 gezelter 2643 If the damping parameter $(\alpha)$ is set to zero, the undamped case,
431     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
432     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
433 chrisfen 2601
434 chrisfen 2636 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
435     derived by Zahn \textit{et al.}; however, there are two important
436     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
437     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
438     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
439     in the Zahn potential, resulting in a potential discontinuity as
440     particles cross $R_\textrm{c}$. Second, the sign of the derivative
441     portion is different. The missing $v_\textrm{c}$ term would not
442     affect molecular dynamics simulations (although the computed energy
443     would be expected to have sudden jumps as particle distances crossed
444 gezelter 2643 $R_c$). The sign problem is a potential source of errors, however.
445     In fact, it introduces a discontinuity in the forces at the cutoff,
446     because the force function is shifted in the wrong direction and
447     doesn't cross zero at $R_\textrm{c}$.
448 chrisfen 2602
449 gezelter 2624 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
450 gezelter 2643 electrostatic summation method in which the potential and forces are
451     continuous at the cutoff radius and which incorporates the damping
452     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
453     this paper, we will evaluate exactly how good these methods ({\sc sp},
454     {\sc sf}, damping) are at reproducing the correct electrostatic
455     summation performed by the Ewald sum.
456 gezelter 2624
457     \subsection{Other alternatives}
458 gezelter 2643 In addition to the methods described above, we considered some other
459     techniques that are commonly used in molecular simulations. The
460 chrisfen 2629 simplest of these is group-based cutoffs. Though of little use for
461 gezelter 2643 charged molecules, collecting atoms into neutral groups takes
462 chrisfen 2629 advantage of the observation that the electrostatic interactions decay
463     faster than those for monopolar pairs.\cite{Steinbach94} When
464 gezelter 2643 considering these molecules as neutral groups, the relative
465     orientations of the molecules control the strength of the interactions
466     at the cutoff radius. Consequently, as these molecular particles move
467     through $R_\textrm{c}$, the energy will drift upward due to the
468     anisotropy of the net molecular dipole interactions.\cite{Rahman71} To
469     maintain good energy conservation, both the potential and derivative
470     need to be smoothly switched to zero at $R_\textrm{c}$.\cite{Adams79}
471     This is accomplished using a standard switching function. If a smooth
472     second derivative is desired, a fifth (or higher) order polynomial can
473     be used.\cite{Andrea83}
474 gezelter 2624
475 chrisfen 2629 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
476 gezelter 2643 and to incorporate the effects of the surroundings, a method like
477     Reaction Field ({\sc rf}) can be used. The original theory for {\sc
478     rf} was originally developed by Onsager,\cite{Onsager36} and it was
479     applied in simulations for the study of water by Barker and
480     Watts.\cite{Barker73} In modern simulation codes, {\sc rf} is simply
481     an extension of the group-based cutoff method where the net dipole
482     within the cutoff sphere polarizes an external dielectric, which
483     reacts back on the central dipole. The same switching function
484     considerations for group-based cutoffs need to made for {\sc rf}, with
485     the additional pre-specification of a dielectric constant.
486 gezelter 2624
487 chrisfen 2608 \section{Methods}
488    
489 chrisfen 2620 In classical molecular mechanics simulations, there are two primary
490     techniques utilized to obtain information about the system of
491     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
492     techniques utilize pairwise summations of interactions between
493     particle sites, but they use these summations in different ways.
494 chrisfen 2608
495 chrisfen 2620 In MC, the potential energy difference between two subsequent
496     configurations dictates the progression of MC sampling. Going back to
497 gezelter 2624 the origins of this method, the acceptance criterion for the canonical
498     ensemble laid out by Metropolis \textit{et al.} states that a
499     subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
500     \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
501     1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
502     alternate method for handling the long-range electrostatics will
503     ensure proper sampling from the ensemble.
504 chrisfen 2608
505 gezelter 2624 In MD, the derivative of the potential governs how the system will
506 chrisfen 2620 progress in time. Consequently, the force and torque vectors on each
507 gezelter 2624 body in the system dictate how the system evolves. If the magnitude
508     and direction of these vectors are similar when using alternate
509     electrostatic summation techniques, the dynamics in the short term
510     will be indistinguishable. Because error in MD calculations is
511     cumulative, one should expect greater deviation at longer times,
512     although methods which have large differences in the force and torque
513     vectors will diverge from each other more rapidly.
514 chrisfen 2608
515 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
516 gezelter 2624 The pairwise summation techniques (outlined in section
517     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
518     studying the energy differences between conformations. We took the
519     SPME-computed energy difference between two conformations to be the
520     correct behavior. An ideal performance by an alternative method would
521     reproduce these energy differences exactly. Since none of the methods
522     provide exact energy differences, we used linear least squares
523     regressions of the $\Delta E$ values between configurations using SPME
524     against $\Delta E$ values using tested methods provides a quantitative
525     comparison of this agreement. Unitary results for both the
526     correlation and correlation coefficient for these regressions indicate
527     equivalent energetic results between the method under consideration
528     and electrostatics handled using SPME. Sample correlation plots for
529     two alternate methods are shown in Fig. \ref{fig:linearFit}.
530 chrisfen 2608
531 chrisfen 2609 \begin{figure}
532     \centering
533 chrisfen 2619 \includegraphics[width = \linewidth]{./dualLinear.pdf}
534     \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.}
535 chrisfen 2609 \label{fig:linearFit}
536     \end{figure}
537    
538 gezelter 2624 Each system type (detailed in section \ref{sec:RepSims}) was
539     represented using 500 independent configurations. Additionally, we
540     used seven different system types, so each of the alternate
541     (non-Ewald) electrostatic summation methods was evaluated using
542     873,250 configurational energy differences.
543 chrisfen 2609
544 gezelter 2624 Results and discussion for the individual analysis of each of the
545     system types appear in the supporting information, while the
546     cumulative results over all the investigated systems appears below in
547     section \ref{sec:EnergyResults}.
548    
549 chrisfen 2609 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
550 gezelter 2624 We evaluated the pairwise methods (outlined in section
551     \ref{sec:ESMethods}) for use in MD simulations by
552     comparing the force and torque vectors with those obtained using the
553     reference Ewald summation (SPME). Both the magnitude and the
554     direction of these vectors on each of the bodies in the system were
555     analyzed. For the magnitude of these vectors, linear least squares
556     regression analyses were performed as described previously for
557     comparing $\Delta E$ values. Instead of a single energy difference
558     between two system configurations, we compared the magnitudes of the
559     forces (and torques) on each molecule in each configuration. For a
560     system of 1000 water molecules and 40 ions, there are 1040 force
561     vectors and 1000 torque vectors. With 500 configurations, this
562     results in 520,000 force and 500,000 torque vector comparisons.
563     Additionally, data from seven different system types was aggregated
564     before the comparison was made.
565 chrisfen 2609
566 gezelter 2624 The {\it directionality} of the force and torque vectors was
567     investigated through measurement of the angle ($\theta$) formed
568     between those computed from the particular method and those from SPME,
569 chrisfen 2610 \begin{equation}
570 chrisfen 2639 \theta_f = \cos^{-1} \left(\hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method}\right),
571 chrisfen 2610 \end{equation}
572 gezelter 2624 where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
573     force vector computed using method $M$.
574    
575 chrisfen 2620 Each of these $\theta$ values was accumulated in a distribution
576 gezelter 2624 function, weighted by the area on the unit sphere. Non-linear
577     Gaussian fits were used to measure the width of the resulting
578     distributions.
579 chrisfen 2609
580     \begin{figure}
581     \centering
582 gezelter 2617 \includegraphics[width = \linewidth]{./gaussFit.pdf}
583 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.}
584     \label{fig:gaussian}
585     \end{figure}
586    
587 chrisfen 2620 Figure \ref{fig:gaussian} shows an example distribution with applied
588     non-linear fits. The solid line is a Gaussian profile, while the
589     dotted line is a Voigt profile, a convolution of a Gaussian and a
590     Lorentzian. Since this distribution is a measure of angular error
591 gezelter 2624 between two different electrostatic summation methods, there is no
592     {\it a priori} reason for the profile to adhere to any specific shape.
593     Gaussian fits was used to compare all the tested methods. The
594     variance ($\sigma^2$) was extracted from each of these fits and was
595     used to compare distribution widths. Values of $\sigma^2$ near zero
596     indicate vector directions indistinguishable from those calculated
597     when using the reference method (SPME).
598 chrisfen 2609
599 gezelter 2624 \subsection{Short-time Dynamics}
600 chrisfen 2638 Evaluation of the short-time dynamics of charged systems was performed
601     by considering the 1000 K NaCl crystal system while using a subset of the
602 chrisfen 2620 best performing pairwise methods. The NaCl crystal was chosen to
603     avoid possible complications involving the propagation techniques of
604 chrisfen 2638 orientational motion in molecular systems. All systems were started
605     with the same initial positions and velocities. Simulations were
606     performed under the microcanonical ensemble, and velocity
607     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
608     of the trajectories,
609 chrisfen 2609 \begin{equation}
610 chrisfen 2638 C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
611 chrisfen 2609 \label{eq:vCorr}
612     \end{equation}
613 chrisfen 2638 Velocity autocorrelation functions require detailed short time data,
614     thus velocity information was saved every 2 fs over 10 ps
615     trajectories. Because the NaCl crystal is composed of two different
616     atom types, the average of the two resulting velocity autocorrelation
617     functions was used for comparisons.
618    
619     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
620     Evaluation of the long-time dynamics of charged systems was performed
621     by considering the NaCl crystal system, again while using a subset of
622     the best performing pairwise methods. To enhance the atomic motion,
623     these crystals were equilibrated at 1000 K, near the experimental
624     $T_m$ for NaCl. Simulations were performed under the microcanonical
625     ensemble, and velocity information was saved every 5 fs over 100 ps
626     trajectories. The power spectrum ($I(\omega)$) was obtained via
627     Fourier transform of the velocity autocorrelation function
628 chrisfen 2609 \begin{equation}
629     I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
630     \label{eq:powerSpec}
631     \end{equation}
632 chrisfen 2638 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
633     NaCl crystal is composed of two different atom types, the average of
634     the two resulting power spectra was used for comparisons.
635 chrisfen 2609
636     \subsection{Representative Simulations}\label{sec:RepSims}
637 chrisfen 2620 A variety of common and representative simulations were analyzed to
638     determine the relative effectiveness of the pairwise summation
639     techniques in reproducing the energetics and dynamics exhibited by
640     SPME. The studied systems were as follows:
641 chrisfen 2599 \begin{enumerate}
642 chrisfen 2586 \item Liquid Water
643     \item Crystalline Water (Ice I$_\textrm{c}$)
644 chrisfen 2595 \item NaCl Crystal
645     \item NaCl Melt
646 chrisfen 2599 \item Low Ionic Strength Solution of NaCl in Water
647     \item High Ionic Strength Solution of NaCl in Water
648 chrisfen 2586 \item 6 \AA\ Radius Sphere of Argon in Water
649 chrisfen 2599 \end{enumerate}
650 chrisfen 2620 By utilizing the pairwise techniques (outlined in section
651     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
652     charged particles, and mixtures of the two, we can comment on possible
653     system dependence and/or universal applicability of the techniques.
654 chrisfen 2586
655 chrisfen 2620 Generation of the system configurations was dependent on the system
656     type. For the solid and liquid water configurations, configuration
657     snapshots were taken at regular intervals from higher temperature 1000
658 chrisfen 2641 SPC/E water molecule trajectories and each equilibrated
659     individually.\cite{Berendsen87} The solid and liquid NaCl systems
660     consisted of 500 Na+ and 500 Cl- ions and were selected and
661     equilibrated in the same fashion as the water systems. For the low
662     and high ionic strength NaCl solutions, 4 and 40 ions were first
663     solvated in a 1000 water molecule boxes respectively. Ion and water
664     positions were then randomly swapped, and the resulting configurations
665     were again equilibrated individually. Finally, for the Argon/Water
666     "charge void" systems, the identities of all the SPC/E waters within 6
667     \AA\ of the center of the equilibrated water configurations were
668     converted to argon (Fig. \ref{fig:argonSlice}).
669 chrisfen 2586
670     \begin{figure}
671     \centering
672 gezelter 2617 \includegraphics[width = \linewidth]{./slice.pdf}
673 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.}
674 chrisfen 2601 \label{fig:argonSlice}
675 chrisfen 2586 \end{figure}
676    
677 chrisfen 2609 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
678 chrisfen 2620 Electrostatic summation method comparisons were performed using SPME,
679 chrisfen 2629 the {\sc sp} and {\sc sf} methods - both with damping
680 chrisfen 2620 parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
681     moderate, and strong damping respectively), reaction field with an
682     infinite dielectric constant, and an unmodified cutoff. Group-based
683     cutoffs with a fifth-order polynomial switching function were
684     necessary for the reaction field simulations and were utilized in the
685     SP, SF, and pure cutoff methods for comparison to the standard lack of
686     group-based cutoffs with a hard truncation. The SPME calculations
687     were performed using the TINKER implementation of SPME,\cite{Ponder87}
688     while all other method calculations were performed using the OOPSE
689     molecular mechanics package.\cite{Meineke05}
690 chrisfen 2586
691 chrisfen 2620 These methods were additionally evaluated with three different cutoff
692     radii (9, 12, and 15 \AA) to investigate possible cutoff radius
693     dependence. It should be noted that the damping parameter chosen in
694     SPME, or so called ``Ewald Coefficient", has a significant effect on
695     the energies and forces calculated. Typical molecular mechanics
696     packages default this to a value dependent on the cutoff radius and a
697     tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller
698 chrisfen 2636 tolerances are typically associated with increased accuracy, but this
699     usually means more time spent calculating the reciprocal-space portion
700     of the summation.\cite{Perram88,Essmann95} The default TINKER
701     tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
702 chrisfen 2620 calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
703     0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
704 chrisfen 2609
705 chrisfen 2575 \section{Results and Discussion}
706    
707 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
708 chrisfen 2620 In order to evaluate the performance of the pairwise electrostatic
709     summation methods for Monte Carlo simulations, the energy differences
710     between configurations were compared to the values obtained when using
711     SPME. The results for the subsequent regression analysis are shown in
712     figure \ref{fig:delE}.
713 chrisfen 2590
714     \begin{figure}
715     \centering
716 gezelter 2617 \includegraphics[width=5.5in]{./delEplot.pdf}
717 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).}
718 chrisfen 2601 \label{fig:delE}
719 chrisfen 2594 \end{figure}
720    
721 chrisfen 2620 In this figure, it is apparent that it is unreasonable to expect
722     realistic results using an unmodified cutoff. This is not all that
723 chrisfen 2641 surprising since this results in large energy fluctuations as atoms or
724     molecules move in and out of the cutoff radius.\cite{Rahman71,Adams79}
725     These fluctuations can be alleviated to some degree by using group
726     based cutoffs with a switching
727     function.\cite{Adams79,Steinbach94,Leach01} The Group Switch Cutoff
728     row doesn't show a significant improvement in this plot because the
729     salt and salt solution systems contain non-neutral groups, see the
730 chrisfen 2620 accompanying supporting information for a comparison where all groups
731     are neutral.
732 chrisfen 2594
733 chrisfen 2620 Correcting the resulting charged cutoff sphere is one of the purposes
734     of the damped Coulomb summation proposed by Wolf \textit{et
735     al.},\cite{Wolf99} and this correction indeed improves the results as
736 chrisfen 2640 seen in the {\sc sp} rows. While the undamped case of this
737 chrisfen 2620 method is a significant improvement over the pure cutoff, it still
738     doesn't correlate that well with SPME. Inclusion of potential damping
739     improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
740     an excellent correlation and quality of fit with the SPME results,
741     particularly with a cutoff radius greater than 12 \AA . Use of a
742     larger damping parameter is more helpful for the shortest cutoff
743     shown, but it has a detrimental effect on simulations with larger
744 chrisfen 2629 cutoffs. In the {\sc sf} sets, increasing damping results in
745 chrisfen 2620 progressively poorer correlation. Overall, the undamped case is the
746     best performing set, as the correlation and quality of fits are
747     consistently superior regardless of the cutoff distance. This result
748     is beneficial in that the undamped case is less computationally
749     prohibitive do to the lack of complimentary error function calculation
750     when performing the electrostatic pair interaction. The reaction
751     field results illustrates some of that method's limitations, primarily
752     that it was developed for use in homogenous systems; although it does
753     provide results that are an improvement over those from an unmodified
754     cutoff.
755 chrisfen 2609
756 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
757 chrisfen 2599
758 chrisfen 2620 Evaluation of pairwise methods for use in Molecular Dynamics
759     simulations requires consideration of effects on the forces and
760     torques. Investigation of the force and torque vector magnitudes
761     provides a measure of the strength of these values relative to SPME.
762     Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
763     force and torque vector magnitude regression results for the
764     accumulated analysis over all the system types.
765 chrisfen 2594
766     \begin{figure}
767     \centering
768 gezelter 2617 \includegraphics[width=5.5in]{./frcMagplot.pdf}
769 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).}
770 chrisfen 2601 \label{fig:frcMag}
771 chrisfen 2594 \end{figure}
772    
773 chrisfen 2620 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
774     in the previous $\Delta E$ section. The unmodified cutoff results are
775     poor, but using group based cutoffs and a switching function provides
776     a improvement much more significant than what was seen with $\Delta
777 chrisfen 2629 E$. Looking at the {\sc sp} sets, the slope and $R^2$
778 chrisfen 2620 improve with the use of damping to an optimal result of 0.2 \AA
779     $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping,
780     while beneficial for simulations with a cutoff radius of 9 \AA\ , is
781     detrimental to simulations with larger cutoff radii. The undamped
782 chrisfen 2629 {\sc sf} method gives forces in line with those obtained using
783 chrisfen 2620 SPME, and use of a damping function results in minor improvement. The
784     reaction field results are surprisingly good, considering the poor
785     quality of the fits for the $\Delta E$ results. There is still a
786     considerable degree of scatter in the data, but it correlates well in
787     general. To be fair, we again note that the reaction field
788     calculations do not encompass NaCl crystal and melt systems, so these
789     results are partly biased towards conditions in which the method
790     performs more favorably.
791 chrisfen 2594
792     \begin{figure}
793     \centering
794 gezelter 2617 \includegraphics[width=5.5in]{./trqMagplot.pdf}
795 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).}
796 chrisfen 2601 \label{fig:trqMag}
797 chrisfen 2594 \end{figure}
798    
799 chrisfen 2620 To evaluate the torque vector magnitudes, the data set from which
800     values are drawn is limited to rigid molecules in the systems
801     (i.e. water molecules). In spite of this smaller sampling pool, the
802     torque vector magnitude results in figure \ref{fig:trqMag} are still
803     similar to those seen for the forces; however, they more clearly show
804     the improved behavior that comes with increasing the cutoff radius.
805 chrisfen 2629 Moderate damping is beneficial to the {\sc sp} and helpful
806     yet possibly unnecessary with the {\sc sf} method, and they also
807 chrisfen 2620 show that over-damping adversely effects all cutoff radii rather than
808     showing an improvement for systems with short cutoffs. The reaction
809     field method performs well when calculating the torques, better than
810     the Shifted Force method over this limited data set.
811 chrisfen 2594
812 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
813 chrisfen 2599
814 chrisfen 2620 Having force and torque vectors with magnitudes that are well
815     correlated to SPME is good, but if they are not pointing in the proper
816     direction the results will be incorrect. These vector directions were
817     investigated through measurement of the angle formed between them and
818     those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared
819     through the variance ($\sigma^2$) of the Gaussian fits of the angle
820     error distributions of the combined set over all system types.
821 chrisfen 2594
822     \begin{figure}
823     \centering
824 gezelter 2617 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
825 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).}
826 chrisfen 2601 \label{fig:frcTrqAng}
827 chrisfen 2594 \end{figure}
828    
829 chrisfen 2620 Both the force and torque $\sigma^2$ results from the analysis of the
830     total accumulated system data are tabulated in figure
831     \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case
832     show the improvement afforded by choosing a longer simulation cutoff.
833     Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
834     of the distribution widths, with a similar improvement going from 12
835 chrisfen 2629 to 15 \AA . The undamped {\sc sf}, Group Based Cutoff, and
836 chrisfen 2620 Reaction Field methods all do equivalently well at capturing the
837     direction of both the force and torque vectors. Using damping
838 chrisfen 2629 improves the angular behavior significantly for the {\sc sp}
839     and moderately for the {\sc sf} methods. Increasing the damping
840 chrisfen 2620 too far is destructive for both methods, particularly to the torque
841     vectors. Again it is important to recognize that the force vectors
842     cover all particles in the systems, while torque vectors are only
843     available for neutral molecular groups. Damping appears to have a
844     more beneficial effect on non-neutral bodies, and this observation is
845     investigated further in the accompanying supporting information.
846 chrisfen 2594
847 chrisfen 2595 \begin{table}[htbp]
848     \centering
849     \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$^2$.}
850 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
851 chrisfen 2595 \\
852     \toprule
853     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
854     \cmidrule(lr){3-6}
855     \cmidrule(l){7-10}
856 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
857 chrisfen 2595 \midrule
858 chrisfen 2599
859     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
860     & \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} \\
861     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
862     & \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} \\
863     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
864     & \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} \\
865 chrisfen 2594
866 chrisfen 2595 \midrule
867    
868 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
869     & \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} \\
870     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
871     & \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} \\
872     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
873     & \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} \\
874 chrisfen 2595
875     \bottomrule
876     \end{tabular}
877 chrisfen 2601 \label{tab:groupAngle}
878 chrisfen 2595 \end{table}
879    
880 chrisfen 2620 Although not discussed previously, group based cutoffs can be applied
881 chrisfen 2629 to both the {\sc sp} and {\sc sf} methods. Use off a
882 chrisfen 2620 switching function corrects for the discontinuities that arise when
883     atoms of a group exit the cutoff before the group's center of mass.
884     Though there are no significant benefit or drawbacks observed in
885     $\Delta E$ and vector magnitude results when doing this, there is a
886     measurable improvement in the vector angle results. Table
887     \ref{tab:groupAngle} shows the angular variance values obtained using
888     group based cutoffs and a switching function alongside the standard
889     results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
890 chrisfen 2629 The {\sc sp} shows much narrower angular distributions for
891 chrisfen 2620 both the force and torque vectors when using an $\alpha$ of 0.2
892 chrisfen 2629 \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
893 chrisfen 2620 undamped and lightly damped cases. Thus, by calculating the
894     electrostatic interactions in terms of molecular pairs rather than
895     atomic pairs, the direction of the force and torque vectors are
896     determined more accurately.
897 chrisfen 2595
898 chrisfen 2620 One additional trend to recognize in table \ref{tab:groupAngle} is
899 chrisfen 2629 that the $\sigma^2$ values for both {\sc sp} and
900     {\sc sf} converge as $\alpha$ increases, something that is easier
901 chrisfen 2620 to see when using group based cutoffs. Looking back on figures
902     \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
903     behavior clearly at large $\alpha$ and cutoff values. The reason for
904     this is that the complimentary error function inserted into the
905     potential weakens the electrostatic interaction as $\alpha$ increases.
906     Thus, at larger values of $\alpha$, both the summation method types
907     progress toward non-interacting functions, so care is required in
908     choosing large damping functions lest one generate an undesirable loss
909     in the pair interaction. Kast \textit{et al.} developed a method for
910     choosing appropriate $\alpha$ values for these types of electrostatic
911     summation methods by fitting to $g(r)$ data, and their methods
912     indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
913     values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
914     to be reasonable choices to obtain proper MC behavior
915     (Fig. \ref{fig:delE}); however, based on these findings, choices this
916     high would introduce error in the molecular torques, particularly for
917     the shorter cutoffs. Based on the above findings, empirical damping
918     up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
919 chrisfen 2629 unnecessary when using the {\sc sf} method.
920 chrisfen 2595
921 chrisfen 2638 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
922 chrisfen 2601
923 chrisfen 2629 In the previous studies using a {\sc sf} variant of the damped
924 chrisfen 2620 Wolf coulomb potential, the structure and dynamics of water were
925     investigated rather extensively.\cite{Zahn02,Kast03} Their results
926 chrisfen 2629 indicated that the damped {\sc sf} method results in properties
927 chrisfen 2620 very similar to those obtained when using the Ewald summation.
928     Considering the statistical results shown above, the good performance
929     of this method is not that surprising. Rather than consider the same
930     systems and simply recapitulate their results, we decided to look at
931     the solid state dynamical behavior obtained using the best performing
932     summation methods from the above results.
933 chrisfen 2601
934     \begin{figure}
935     \centering
936 chrisfen 2638 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
937     \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.}
938     \label{fig:vCorrPlot}
939     \end{figure}
940    
941     The short-time decays through the first collision are nearly identical
942     in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
943     functions show how the methods differ. The undamped {\sc sf} method
944     has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
945     peaks than any of the other methods. As the damping function is
946     increased, these peaks are smoothed out, and approach the SPME
947     curve. The damping acts as a distance dependent Gaussian screening of
948     the point charges for the pairwise summation methods; thus, the
949 chrisfen 2640 collisions are more elastic in the undamped {\sc sf} potential, and the
950 chrisfen 2638 stiffness of the potential is diminished as the electrostatic
951     interactions are softened by the damping function. With $\alpha$
952     values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
953     nearly identical and track the SPME features quite well. This is not
954     too surprising in that the differences between the {\sc sf} and {\sc
955     sp} potentials are mitigated with increased damping. However, this
956     appears to indicate that once damping is utilized, the form of the
957     potential seems to play a lesser role in the crystal dynamics.
958    
959     \subsection{Collective Motion: Power Spectra of NaCl Crystals}
960    
961     The short time dynamics were extended to evaluate how the differences
962     between the methods affect the collective long-time motion. The same
963     electrostatic summation methods were used as in the short time
964     velocity autocorrelation function evaluation, but the trajectories
965     were sampled over a much longer time. The power spectra of the
966     resulting velocity autocorrelation functions were calculated and are
967     displayed in figure \ref{fig:methodPS}.
968    
969     \begin{figure}
970     \centering
971 gezelter 2617 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
972 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.}
973 chrisfen 2610 \label{fig:methodPS}
974 chrisfen 2601 \end{figure}
975    
976 chrisfen 2638 While high frequency peaks of the spectra in this figure overlap,
977     showing the same general features, the low frequency region shows how
978     the summation methods differ. Considering the low-frequency inset
979     (expanded in the upper frame of figure \ref{fig:dampInc}), at
980     frequencies below 100 cm$^{-1}$, the correlated motions are
981     blue-shifted when using undamped or weakly damped {\sc sf}. When
982     using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
983     and {\sc sp} methods give near identical correlated motion behavior as
984     the Ewald method (which has a damping value of 0.3119). This
985     weakening of the electrostatic interaction with increased damping
986     explains why the long-ranged correlated motions are at lower
987     frequencies for the moderately damped methods than for undamped or
988     weakly damped methods. To see this effect more clearly, we show how
989     damping strength alone affects a simple real-space electrostatic
990     potential,
991 chrisfen 2601 \begin{equation}
992 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),
993 chrisfen 2601 \end{equation}
994 chrisfen 2620 where $S(r)$ is a switching function that smoothly zeroes the
995     potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
996     the low frequency motions are dependent on the damping used in the
997     direct electrostatic sum. As the damping increases, the peaks drop to
998     lower frequencies. Incidentally, use of an $\alpha$ of 0.25
999     \AA$^{-1}$ on a simple electrostatic summation results in low
1000     frequency correlated dynamics equivalent to a simulation using SPME.
1001     When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1002     shift to higher frequency in exponential fashion. Though not shown,
1003     the spectrum for the simple undamped electrostatic potential is
1004     blue-shifted such that the lowest frequency peak resides near 325
1005 chrisfen 2638 cm$^{-1}$. In light of these results, the undamped {\sc sf} method
1006     producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1007     respectable and shows that the shifted force procedure accounts for
1008     most of the effect afforded through use of the Ewald summation.
1009     However, it appears as though moderate damping is required for
1010     accurate reproduction of crystal dynamics.
1011 chrisfen 2601 \begin{figure}
1012     \centering
1013 gezelter 2617 \includegraphics[width = \linewidth]{./comboSquare.pdf}
1014 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.}
1015 chrisfen 2601 \label{fig:dampInc}
1016     \end{figure}
1017    
1018 chrisfen 2575 \section{Conclusions}
1019    
1020 chrisfen 2620 This investigation of pairwise electrostatic summation techniques
1021     shows that there are viable and more computationally efficient
1022     electrostatic summation techniques than the Ewald summation, chiefly
1023     methods derived from the damped Coulombic sum originally proposed by
1024     Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1025 chrisfen 2629 {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1026 chrisfen 2620 shows a remarkable ability to reproduce the energetic and dynamic
1027     characteristics exhibited by simulations employing lattice summation
1028     techniques. The cumulative energy difference results showed the
1029 chrisfen 2629 undamped {\sc sf} and moderately damped {\sc sp} methods
1030 chrisfen 2620 produced results nearly identical to SPME. Similarly for the dynamic
1031 chrisfen 2629 features, the undamped or moderately damped {\sc sf} and
1032     moderately damped {\sc sp} methods produce force and torque
1033 chrisfen 2620 vector magnitude and directions very similar to the expected values.
1034     These results translate into long-time dynamic behavior equivalent to
1035     that produced in simulations using SPME.
1036 chrisfen 2604
1037 chrisfen 2620 Aside from the computational cost benefit, these techniques have
1038     applicability in situations where the use of the Ewald sum can prove
1039     problematic. Primary among them is their use in interfacial systems,
1040     where the unmodified lattice sum techniques artificially accentuate
1041     the periodicity of the system in an undesirable manner. There have
1042     been alterations to the standard Ewald techniques, via corrections and
1043     reformulations, to compensate for these systems; but the pairwise
1044     techniques discussed here require no modifications, making them
1045     natural tools to tackle these problems. Additionally, this
1046     transferability gives them benefits over other pairwise methods, like
1047     reaction field, because estimations of physical properties (e.g. the
1048     dielectric constant) are unnecessary.
1049 chrisfen 2605
1050 chrisfen 2620 We are not suggesting any flaw with the Ewald sum; in fact, it is the
1051     standard by which these simple pairwise sums are judged. However,
1052     these results do suggest that in the typical simulations performed
1053     today, the Ewald summation may no longer be required to obtain the
1054 chrisfen 2638 level of accuracy most researchers have come to expect
1055 chrisfen 2605
1056 chrisfen 2575 \section{Acknowledgments}
1057 chrisfen 2594 \newpage
1058    
1059 gezelter 2617 \bibliographystyle{jcp2}
1060 chrisfen 2575 \bibliography{electrostaticMethods}
1061    
1062    
1063     \end{document}