ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
Revision: 2638
Committed: Sun Mar 19 19:34:53 2006 UTC (18 years, 3 months ago) by chrisfen
Content type: application/x-tex
File size: 58464 byte(s)
Log Message:
Added a figure and some more text

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 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,
108 chrisfen 2636 \end{split}
109     \label{eq:EwaldSum}
110     \end{equation}
111     where $\alpha$ is a damping parameter, or separation constant, with
112 chrisfen 2637 units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and equal
113     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
114     constant of the encompassing medium. The final two terms of
115 chrisfen 2636 eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
116     for interacting with a surrounding dielectric.\cite{Allen87} This
117     dipolar term was neglected in early applications in molecular
118     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
119     Leeuw {\it et al.} to address situations where the unit cell has a
120     dipole moment and this dipole moment gets magnified through
121 chrisfen 2637 replication of the periodic images.\cite{deLeeuw80,Smith81} If this
122     term is taken to be zero, the system is using conducting boundary
123     conditions, $\epsilon_{\rm S} = \infty$. Figure
124 chrisfen 2636 \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
125     time. Initially, due to the small size of systems, the entire
126     simulation box was replicated to convergence. Currently, we balance a
127     spherical real-space cutoff with the reciprocal sum and consider the
128     surrounding dielectric.
129 chrisfen 2610 \begin{figure}
130     \centering
131 gezelter 2617 \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
132     \caption{How the application of the Ewald summation has changed with
133     the increase in computer power. Initially, only small numbers of
134     particles could be studied, and the Ewald sum acted to replicate the
135     unit cell charge distribution out to convergence. Now, much larger
136     systems of charges are investigated with fixed distance cutoffs. The
137     calculated structure factor is used to sum out to great distance, and
138     a surrounding dielectric term is included.}
139 chrisfen 2610 \label{fig:ewaldTime}
140     \end{figure}
141    
142 chrisfen 2636 The Ewald summation in the straight-forward form is an
143     $\mathscr{O}(N^2)$ algorithm. The separation constant $(\alpha)$
144     plays an important role in the computational cost balance between the
145     direct and reciprocal-space portions of the summation. The choice of
146 chrisfen 2637 the magnitude of this value allows one to select whether the
147     real-space or reciprocal space portion of the summation is an
148     $\mathscr{O}(N^2)$ calcualtion (with the other being
149     $\mathscr{O}(N)$).\cite{Sagui99} With appropriate choice of $\alpha$
150     and thoughtful algorithm development, this cost can be brought down to
151 chrisfen 2636 $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route taken to
152 chrisfen 2637 reduce the cost of the Ewald summation further is to set $\alpha$ such
153     that the real-space interactions decay rapidly, allowing for a short
154     spherical cutoff, and then optimize the reciprocal space summation.
155     These optimizations usually involve the utilization of the fast
156     Fourier transform (FFT),\cite{Hockney81} leading to the
157     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
158     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
159     methods, the cost of the reciprocal-space portion of the Ewald
160     summation is from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N \log N)$.
161 chrisfen 2636
162 chrisfen 2637 These developments and optimizations have led the use of the Ewald
163     summation to become routine in simulations with periodic boundary
164     conditions. However, in certain systems the intrinsic three
165     dimensional periodicity can prove to be problematic, such as two
166     dimensional surfaces and membranes. The Ewald sum has been
167     reformulated to handle 2D
168     systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the new
169     methods have been found to be computationally
170     expensive.\cite{Spohr97,Yeh99} Inclusion of a correction term in the
171     full Ewald summation is a possible direction for enabling the handling
172     of 2D systems and the inclusion of the optimizations described
173     previously.\cite{Yeh99}
174    
175     Several studies have recognized that the inherent periodicity in the
176     Ewald sum can also have an effect on systems that have the same
177     dimensionality.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
178     Good examples are solvated proteins kept at high relative
179     concentration due to the periodicity of the electrostatics. In these
180     systems, the more compact folded states of a protein can be
181     artificially stabilized by the periodic replicas introduced by the
182     Ewald summation.\cite{Weber00} Thus, care ought to be taken when
183     considering the use of the Ewald summation where the intrinsic
184     perodicity may negatively affect the system dynamics.
185    
186    
187 chrisfen 2608 \subsection{The Wolf and Zahn Methods}
188 gezelter 2617 In a recent paper by Wolf \textit{et al.}, a procedure was outlined
189 gezelter 2624 for the accurate accumulation of electrostatic interactions in an
190 chrisfen 2637 efficient pairwise fashion and lacks the inherent periodicity of the
191     Ewald summation.\cite{Wolf99} Wolf \textit{et al.} observed that the
192     electrostatic interaction is effectively short-ranged in condensed
193     phase systems and that neutralization of the charge contained within
194     the cutoff radius is crucial for potential stability. They devised a
195     pairwise summation method that ensures charge neutrality and gives
196     results similar to those obtained with the Ewald summation. The
197     resulting shifted Coulomb potential (Eq. \ref{eq:WolfPot}) includes
198     image-charges subtracted out through placement on the cutoff sphere
199     and a distance-dependent damping function (identical to that seen in
200     the real-space portion of the Ewald sum) to aid convergence
201 chrisfen 2601 \begin{equation}
202 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\}.
203 chrisfen 2601 \label{eq:WolfPot}
204     \end{equation}
205 gezelter 2617 Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
206     potential. However, neutralizing the charge contained within each
207     cutoff sphere requires the placement of a self-image charge on the
208     surface of the cutoff sphere. This additional self-term in the total
209 gezelter 2624 potential enabled Wolf {\it et al.} to obtain excellent estimates of
210 gezelter 2617 Madelung energies for many crystals.
211    
212     In order to use their charge-neutralized potential in molecular
213     dynamics simulations, Wolf \textit{et al.} suggested taking the
214     derivative of this potential prior to evaluation of the limit. This
215     procedure gives an expression for the forces,
216 chrisfen 2601 \begin{equation}
217 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\},
218 chrisfen 2601 \label{eq:WolfForces}
219     \end{equation}
220 gezelter 2617 that incorporates both image charges and damping of the electrostatic
221     interaction.
222    
223     More recently, Zahn \textit{et al.} investigated these potential and
224     force expressions for use in simulations involving water.\cite{Zahn02}
225 gezelter 2624 In their work, they pointed out that the forces and derivative of
226     the potential are not commensurate. Attempts to use both
227     Eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
228     to poor energy conservation. They correctly observed that taking the
229     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
230     derivatives gives forces for a different potential energy function
231     than the one shown in Eq. (\ref{eq:WolfPot}).
232 gezelter 2617
233     Zahn \textit{et al.} proposed a modified form of this ``Wolf summation
234     method'' as a way to use this technique in Molecular Dynamics
235     simulations. Taking the integral of the forces shown in equation
236     \ref{eq:WolfForces}, they proposed a new damped Coulomb
237     potential,
238 chrisfen 2601 \begin{equation}
239 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\}.
240 chrisfen 2601 \label{eq:ZahnPot}
241     \end{equation}
242 gezelter 2617 They showed that this potential does fairly well at capturing the
243     structural and dynamic properties of water compared the same
244     properties obtained using the Ewald sum.
245 chrisfen 2601
246 chrisfen 2608 \subsection{Simple Forms for Pairwise Electrostatics}
247    
248 gezelter 2617 The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
249     al.} are constructed using two different (and separable) computational
250 gezelter 2624 tricks: \begin{enumerate}
251 gezelter 2617 \item shifting through the use of image charges, and
252     \item damping the electrostatic interaction.
253 gezelter 2624 \end{enumerate} Wolf \textit{et al.} treated the
254 gezelter 2617 development of their summation method as a progressive application of
255     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
256     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
257     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
258     both techniques. It is possible, however, to separate these
259     tricks and study their effects independently.
260    
261     Starting with the original observation that the effective range of the
262     electrostatic interaction in condensed phases is considerably less
263     than $r^{-1}$, either the cutoff sphere neutralization or the
264     distance-dependent damping technique could be used as a foundation for
265     a new pairwise summation method. Wolf \textit{et al.} made the
266     observation that charge neutralization within the cutoff sphere plays
267     a significant role in energy convergence; therefore we will begin our
268     analysis with the various shifted forms that maintain this charge
269     neutralization. We can evaluate the methods of Wolf
270     \textit{et al.} and Zahn \textit{et al.} by considering the standard
271     shifted potential,
272 chrisfen 2601 \begin{equation}
273 gezelter 2624 v_\textrm{SP}(r) = \begin{cases}
274 gezelter 2617 v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
275     R_\textrm{c}
276     \end{cases},
277     \label{eq:shiftingPotForm}
278     \end{equation}
279     and shifted force,
280     \begin{equation}
281 gezelter 2624 v_\textrm{SF}(r) = \begin{cases}
282     v(r)-v_\textrm{c}-\left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
283 gezelter 2617 &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
284 chrisfen 2601 \end{cases},
285 chrisfen 2612 \label{eq:shiftingForm}
286 chrisfen 2601 \end{equation}
287 gezelter 2617 functions where $v(r)$ is the unshifted form of the potential, and
288     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
289     that both the potential and the forces goes to zero at the cutoff
290     radius, while the Shifted Potential ({\sc sp}) form only ensures the
291     potential is smooth at the cutoff radius
292     ($R_\textrm{c}$).\cite{Allen87}
293    
294 gezelter 2624 The forces associated with the shifted potential are simply the forces
295     of the unshifted potential itself (when inside the cutoff sphere),
296 chrisfen 2601 \begin{equation}
297 chrisfen 2636 f_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
298 chrisfen 2612 \end{equation}
299 gezelter 2624 and are zero outside. Inside the cutoff sphere, the forces associated
300     with the shifted force form can be written,
301 chrisfen 2612 \begin{equation}
302 chrisfen 2636 f_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
303 gezelter 2624 v(r)}{dr} \right)_{r=R_\textrm{c}}.
304     \end{equation}
305    
306     If the potential ($v(r)$) is taken to be the normal Coulomb potential,
307     \begin{equation}
308     v(r) = \frac{q_i q_j}{r},
309     \label{eq:Coulomb}
310     \end{equation}
311     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
312     al.}'s undamped prescription:
313     \begin{equation}
314 chrisfen 2636 v_\textrm{SP}(r) =
315 gezelter 2624 q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
316     r\leqslant R_\textrm{c},
317 chrisfen 2636 \label{eq:SPPot}
318 gezelter 2624 \end{equation}
319     with associated forces,
320     \begin{equation}
321 chrisfen 2636 f_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right) \quad r\leqslant R_\textrm{c}.
322     \label{eq:SPForces}
323 chrisfen 2612 \end{equation}
324 gezelter 2624 These forces are identical to the forces of the standard Coulomb
325     interaction, and cutting these off at $R_c$ was addressed by Wolf
326     \textit{et al.} as undesirable. They pointed out that the effect of
327     the image charges is neglected in the forces when this form is
328     used,\cite{Wolf99} thereby eliminating any benefit from the method in
329     molecular dynamics. Additionally, there is a discontinuity in the
330     forces at the cutoff radius which results in energy drift during MD
331     simulations.
332 chrisfen 2612
333 gezelter 2624 The shifted force ({\sc sf}) form using the normal Coulomb potential
334     will give,
335 chrisfen 2612 \begin{equation}
336 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}.
337 chrisfen 2612 \label{eq:SFPot}
338     \end{equation}
339 gezelter 2624 with associated forces,
340 chrisfen 2612 \begin{equation}
341 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}.
342 chrisfen 2612 \label{eq:SFForces}
343     \end{equation}
344 gezelter 2624 This formulation has the benefits that there are no discontinuities at
345     the cutoff distance, while the neutralizing image charges are present
346     in both the energy and force expressions. It would be simple to add
347     the self-neutralizing term back when computing the total energy of the
348     system, thereby maintaining the agreement with the Madelung energies.
349     A side effect of this treatment is the alteration in the shape of the
350     potential that comes from the derivative term. Thus, a degree of
351     clarity about agreement with the empirical potential is lost in order
352     to gain functionality in dynamics simulations.
353 chrisfen 2612
354 chrisfen 2620 Wolf \textit{et al.} originally discussed the energetics of the
355 chrisfen 2636 shifted Coulomb potential (Eq. \ref{eq:SPPot}), and they found that
356 gezelter 2624 it was still insufficient for accurate determination of the energy
357     with reasonable cutoff distances. The calculated Madelung energies
358     fluctuate around the expected value with increasing cutoff radius, but
359     the oscillations converge toward the correct value.\cite{Wolf99} A
360     damping function was incorporated to accelerate the convergence; and
361     though alternative functional forms could be
362     used,\cite{Jones56,Heyes81} the complimentary error function was
363     chosen to mirror the effective screening used in the Ewald summation.
364     Incorporating this error function damping into the simple Coulomb
365     potential,
366 chrisfen 2612 \begin{equation}
367 gezelter 2624 v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
368 chrisfen 2601 \label{eq:dampCoulomb}
369     \end{equation}
370 chrisfen 2636 the shifted potential (Eq. (\ref{eq:SPPot})) can be reacquired using
371     eq. (\ref{eq:shiftingForm}),
372 chrisfen 2601 \begin{equation}
373 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},
374 chrisfen 2612 \label{eq:DSPPot}
375 chrisfen 2629 \end{equation}
376 gezelter 2624 with associated forces,
377 chrisfen 2612 \begin{equation}
378 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}.
379 chrisfen 2612 \label{eq:DSPForces}
380     \end{equation}
381 gezelter 2624 Again, this damped shifted potential suffers from a discontinuity and
382     a lack of the image charges in the forces. To remedy these concerns,
383 chrisfen 2629 one may derive a {\sc sf} variant by including the derivative
384     term in eq. (\ref{eq:shiftingForm}),
385 chrisfen 2612 \begin{equation}
386 chrisfen 2620 \begin{split}
387 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}.
388 chrisfen 2612 \label{eq:DSFPot}
389 chrisfen 2620 \end{split}
390 chrisfen 2612 \end{equation}
391 chrisfen 2636 The derivative of the above potential will lead to the following forces,
392 chrisfen 2612 \begin{equation}
393 chrisfen 2620 \begin{split}
394 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}.
395 chrisfen 2612 \label{eq:DSFForces}
396 chrisfen 2620 \end{split}
397 chrisfen 2612 \end{equation}
398 chrisfen 2636 If the damping parameter $(\alpha)$ is chosen to be zero, the undamped
399     case, eqs. (\ref{eq:SPPot}-\ref{eq:SFForces}) are correctly recovered
400     from eqs. (\ref{eq:DSPPot}-\ref{eq:DSFForces}).
401 chrisfen 2601
402 chrisfen 2636 This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
403     derived by Zahn \textit{et al.}; however, there are two important
404     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
405     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
406     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
407     in the Zahn potential, resulting in a potential discontinuity as
408     particles cross $R_\textrm{c}$. Second, the sign of the derivative
409     portion is different. The missing $v_\textrm{c}$ term would not
410     affect molecular dynamics simulations (although the computed energy
411     would be expected to have sudden jumps as particle distances crossed
412     $R_c$). The sign problem would be a potential source of errors,
413     however. In fact, it introduces a discontinuity in the forces at the
414     cutoff, because the force function is shifted in the wrong direction
415     and doesn't cross zero at $R_\textrm{c}$.
416 chrisfen 2602
417 gezelter 2624 Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
418     electrostatic summation method that is continuous in both the
419     potential and forces and which incorporates the damping function
420     proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of this
421     paper, we will evaluate exactly how good these methods ({\sc sp}, {\sc
422     sf}, damping) are at reproducing the correct electrostatic summation
423     performed by the Ewald sum.
424    
425     \subsection{Other alternatives}
426 chrisfen 2629 In addition to the methods described above, we will consider some
427     other techniques that commonly get used in molecular simulations. The
428     simplest of these is group-based cutoffs. Though of little use for
429     non-neutral molecules, collecting atoms into neutral groups takes
430     advantage of the observation that the electrostatic interactions decay
431     faster than those for monopolar pairs.\cite{Steinbach94} When
432     considering these molecules as groups, an orientational aspect is
433     introduced to the interactions. Consequently, as these molecular
434     particles move through $R_\textrm{c}$, the energy will drift upward
435     due to the anisotropy of the net molecular dipole
436     interactions.\cite{Rahman71} To maintain good energy conservation,
437     both the potential and derivative need to be smoothly switched to zero
438     at $R_\textrm{c}$.\cite{Adams79} This is accomplished using a
439     switching function,
440     \begin{equation}
441     S(r) = \begin{cases} 1 &\quad r\leqslant r_\textrm{sw} \\
442     \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} \\
443     0 &\quad r>R_\textrm{c}
444     \end{cases},
445     \end{equation}
446     where the above form is for a cubic function. If a smooth second
447     derivative is desired, a fifth (or higher) order polynomial can be
448     used.\cite{Andrea83}
449 gezelter 2624
450 chrisfen 2629 Group-based cutoffs neglect the surroundings beyond $R_\textrm{c}$,
451     and to incorporate their effect, a method like Reaction Field ({\sc
452 chrisfen 2630 rf}) can be used. The original theory for {\sc rf} was originally
453 chrisfen 2629 developed by Onsager,\cite{Onsager36} and it was applied in
454     simulations for the study of water by Barker and Watts.\cite{Barker73}
455     In application, it is simply an extension of the group-based cutoff
456     method where the net dipole within the cutoff sphere polarizes an
457     external dielectric, which reacts back on the central dipole. The
458     same switching function considerations for group-based cutoffs need to
459 chrisfen 2630 made for {\sc rf}, with the additional pre-specification of a
460 chrisfen 2629 dielectric constant.
461 gezelter 2624
462 chrisfen 2608 \section{Methods}
463    
464 chrisfen 2620 In classical molecular mechanics simulations, there are two primary
465     techniques utilized to obtain information about the system of
466     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
467     techniques utilize pairwise summations of interactions between
468     particle sites, but they use these summations in different ways.
469 chrisfen 2608
470 chrisfen 2620 In MC, the potential energy difference between two subsequent
471     configurations dictates the progression of MC sampling. Going back to
472 gezelter 2624 the origins of this method, the acceptance criterion for the canonical
473     ensemble laid out by Metropolis \textit{et al.} states that a
474     subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
475     \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
476     1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
477     alternate method for handling the long-range electrostatics will
478     ensure proper sampling from the ensemble.
479 chrisfen 2608
480 gezelter 2624 In MD, the derivative of the potential governs how the system will
481 chrisfen 2620 progress in time. Consequently, the force and torque vectors on each
482 gezelter 2624 body in the system dictate how the system evolves. If the magnitude
483     and direction of these vectors are similar when using alternate
484     electrostatic summation techniques, the dynamics in the short term
485     will be indistinguishable. Because error in MD calculations is
486     cumulative, one should expect greater deviation at longer times,
487     although methods which have large differences in the force and torque
488     vectors will diverge from each other more rapidly.
489 chrisfen 2608
490 chrisfen 2609 \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
491 gezelter 2624 The pairwise summation techniques (outlined in section
492     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
493     studying the energy differences between conformations. We took the
494     SPME-computed energy difference between two conformations to be the
495     correct behavior. An ideal performance by an alternative method would
496     reproduce these energy differences exactly. Since none of the methods
497     provide exact energy differences, we used linear least squares
498     regressions of the $\Delta E$ values between configurations using SPME
499     against $\Delta E$ values using tested methods provides a quantitative
500     comparison of this agreement. Unitary results for both the
501     correlation and correlation coefficient for these regressions indicate
502     equivalent energetic results between the method under consideration
503     and electrostatics handled using SPME. Sample correlation plots for
504     two alternate methods are shown in Fig. \ref{fig:linearFit}.
505 chrisfen 2608
506 chrisfen 2609 \begin{figure}
507     \centering
508 chrisfen 2619 \includegraphics[width = \linewidth]{./dualLinear.pdf}
509     \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.}
510 chrisfen 2609 \label{fig:linearFit}
511     \end{figure}
512    
513 gezelter 2624 Each system type (detailed in section \ref{sec:RepSims}) was
514     represented using 500 independent configurations. Additionally, we
515     used seven different system types, so each of the alternate
516     (non-Ewald) electrostatic summation methods was evaluated using
517     873,250 configurational energy differences.
518 chrisfen 2609
519 gezelter 2624 Results and discussion for the individual analysis of each of the
520     system types appear in the supporting information, while the
521     cumulative results over all the investigated systems appears below in
522     section \ref{sec:EnergyResults}.
523    
524 chrisfen 2609 \subsection{Molecular Dynamics and the Force and Torque Vectors}\label{sec:MDMethods}
525 gezelter 2624 We evaluated the pairwise methods (outlined in section
526     \ref{sec:ESMethods}) for use in MD simulations by
527     comparing the force and torque vectors with those obtained using the
528     reference Ewald summation (SPME). Both the magnitude and the
529     direction of these vectors on each of the bodies in the system were
530     analyzed. For the magnitude of these vectors, linear least squares
531     regression analyses were performed as described previously for
532     comparing $\Delta E$ values. Instead of a single energy difference
533     between two system configurations, we compared the magnitudes of the
534     forces (and torques) on each molecule in each configuration. For a
535     system of 1000 water molecules and 40 ions, there are 1040 force
536     vectors and 1000 torque vectors. With 500 configurations, this
537     results in 520,000 force and 500,000 torque vector comparisons.
538     Additionally, data from seven different system types was aggregated
539     before the comparison was made.
540 chrisfen 2609
541 gezelter 2624 The {\it directionality} of the force and torque vectors was
542     investigated through measurement of the angle ($\theta$) formed
543     between those computed from the particular method and those from SPME,
544 chrisfen 2610 \begin{equation}
545 gezelter 2624 \theta_f = \cos^{-1} \hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method},
546 chrisfen 2610 \end{equation}
547 gezelter 2624 where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
548     force vector computed using method $M$.
549    
550 chrisfen 2620 Each of these $\theta$ values was accumulated in a distribution
551 gezelter 2624 function, weighted by the area on the unit sphere. Non-linear
552     Gaussian fits were used to measure the width of the resulting
553     distributions.
554 chrisfen 2609
555     \begin{figure}
556     \centering
557 gezelter 2617 \includegraphics[width = \linewidth]{./gaussFit.pdf}
558 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.}
559     \label{fig:gaussian}
560     \end{figure}
561    
562 chrisfen 2620 Figure \ref{fig:gaussian} shows an example distribution with applied
563     non-linear fits. The solid line is a Gaussian profile, while the
564     dotted line is a Voigt profile, a convolution of a Gaussian and a
565     Lorentzian. Since this distribution is a measure of angular error
566 gezelter 2624 between two different electrostatic summation methods, there is no
567     {\it a priori} reason for the profile to adhere to any specific shape.
568     Gaussian fits was used to compare all the tested methods. The
569     variance ($\sigma^2$) was extracted from each of these fits and was
570     used to compare distribution widths. Values of $\sigma^2$ near zero
571     indicate vector directions indistinguishable from those calculated
572     when using the reference method (SPME).
573 chrisfen 2609
574 gezelter 2624 \subsection{Short-time Dynamics}
575 chrisfen 2638 Evaluation of the short-time dynamics of charged systems was performed
576     by considering the 1000 K NaCl crystal system while using a subset of the
577 chrisfen 2620 best performing pairwise methods. The NaCl crystal was chosen to
578     avoid possible complications involving the propagation techniques of
579 chrisfen 2638 orientational motion in molecular systems. All systems were started
580     with the same initial positions and velocities. Simulations were
581     performed under the microcanonical ensemble, and velocity
582     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
583     of the trajectories,
584 chrisfen 2609 \begin{equation}
585 chrisfen 2638 C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
586 chrisfen 2609 \label{eq:vCorr}
587     \end{equation}
588 chrisfen 2638 Velocity autocorrelation functions require detailed short time data,
589     thus velocity information was saved every 2 fs over 10 ps
590     trajectories. Because the NaCl crystal is composed of two different
591     atom types, the average of the two resulting velocity autocorrelation
592     functions was used for comparisons.
593    
594     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
595     Evaluation of the long-time dynamics of charged systems was performed
596     by considering the NaCl crystal system, again while using a subset of
597     the best performing pairwise methods. To enhance the atomic motion,
598     these crystals were equilibrated at 1000 K, near the experimental
599     $T_m$ for NaCl. Simulations were performed under the microcanonical
600     ensemble, and velocity information was saved every 5 fs over 100 ps
601     trajectories. The power spectrum ($I(\omega)$) was obtained via
602     Fourier transform of the velocity autocorrelation function
603 chrisfen 2609 \begin{equation}
604     I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
605     \label{eq:powerSpec}
606     \end{equation}
607 chrisfen 2638 where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
608     NaCl crystal is composed of two different atom types, the average of
609     the two resulting power spectra was used for comparisons.
610 chrisfen 2609
611     \subsection{Representative Simulations}\label{sec:RepSims}
612 chrisfen 2620 A variety of common and representative simulations were analyzed to
613     determine the relative effectiveness of the pairwise summation
614     techniques in reproducing the energetics and dynamics exhibited by
615     SPME. The studied systems were as follows:
616 chrisfen 2599 \begin{enumerate}
617 chrisfen 2586 \item Liquid Water
618     \item Crystalline Water (Ice I$_\textrm{c}$)
619 chrisfen 2595 \item NaCl Crystal
620     \item NaCl Melt
621 chrisfen 2599 \item Low Ionic Strength Solution of NaCl in Water
622     \item High Ionic Strength Solution of NaCl in Water
623 chrisfen 2586 \item 6 \AA\ Radius Sphere of Argon in Water
624 chrisfen 2599 \end{enumerate}
625 chrisfen 2620 By utilizing the pairwise techniques (outlined in section
626     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
627     charged particles, and mixtures of the two, we can comment on possible
628     system dependence and/or universal applicability of the techniques.
629 chrisfen 2586
630 chrisfen 2620 Generation of the system configurations was dependent on the system
631     type. For the solid and liquid water configurations, configuration
632     snapshots were taken at regular intervals from higher temperature 1000
633     SPC/E water molecule trajectories and each equilibrated individually.
634     The solid and liquid NaCl systems consisted of 500 Na+ and 500 Cl-
635     ions and were selected and equilibrated in the same fashion as the
636     water systems. For the low and high ionic strength NaCl solutions, 4
637     and 40 ions were first solvated in a 1000 water molecule boxes
638     respectively. Ion and water positions were then randomly swapped, and
639     the resulting configurations were again equilibrated individually.
640     Finally, for the Argon/Water "charge void" systems, the identities of
641     all the SPC/E waters within 6 \AA\ of the center of the equilibrated
642     water configurations were converted to argon
643     (Fig. \ref{fig:argonSlice}).
644 chrisfen 2586
645     \begin{figure}
646     \centering
647 gezelter 2617 \includegraphics[width = \linewidth]{./slice.pdf}
648 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.}
649 chrisfen 2601 \label{fig:argonSlice}
650 chrisfen 2586 \end{figure}
651    
652 chrisfen 2609 \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
653 chrisfen 2620 Electrostatic summation method comparisons were performed using SPME,
654 chrisfen 2629 the {\sc sp} and {\sc sf} methods - both with damping
655 chrisfen 2620 parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
656     moderate, and strong damping respectively), reaction field with an
657     infinite dielectric constant, and an unmodified cutoff. Group-based
658     cutoffs with a fifth-order polynomial switching function were
659     necessary for the reaction field simulations and were utilized in the
660     SP, SF, and pure cutoff methods for comparison to the standard lack of
661     group-based cutoffs with a hard truncation. The SPME calculations
662     were performed using the TINKER implementation of SPME,\cite{Ponder87}
663     while all other method calculations were performed using the OOPSE
664     molecular mechanics package.\cite{Meineke05}
665 chrisfen 2586
666 chrisfen 2620 These methods were additionally evaluated with three different cutoff
667     radii (9, 12, and 15 \AA) to investigate possible cutoff radius
668     dependence. It should be noted that the damping parameter chosen in
669     SPME, or so called ``Ewald Coefficient", has a significant effect on
670     the energies and forces calculated. Typical molecular mechanics
671     packages default this to a value dependent on the cutoff radius and a
672     tolerance (typically less than $1 \times 10^{-4}$ kcal/mol). Smaller
673 chrisfen 2636 tolerances are typically associated with increased accuracy, but this
674     usually means more time spent calculating the reciprocal-space portion
675     of the summation.\cite{Perram88,Essmann95} The default TINKER
676     tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
677 chrisfen 2620 calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
678     0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
679 chrisfen 2609
680 chrisfen 2575 \section{Results and Discussion}
681    
682 chrisfen 2609 \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
683 chrisfen 2620 In order to evaluate the performance of the pairwise electrostatic
684     summation methods for Monte Carlo simulations, the energy differences
685     between configurations were compared to the values obtained when using
686     SPME. The results for the subsequent regression analysis are shown in
687     figure \ref{fig:delE}.
688 chrisfen 2590
689     \begin{figure}
690     \centering
691 gezelter 2617 \includegraphics[width=5.5in]{./delEplot.pdf}
692 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).}
693 chrisfen 2601 \label{fig:delE}
694 chrisfen 2594 \end{figure}
695    
696 chrisfen 2620 In this figure, it is apparent that it is unreasonable to expect
697     realistic results using an unmodified cutoff. This is not all that
698     surprising since this results in large energy fluctuations as atoms
699     move in and out of the cutoff radius. These fluctuations can be
700     alleviated to some degree by using group based cutoffs with a
701     switching function.\cite{Steinbach94} The Group Switch Cutoff row
702     doesn't show a significant improvement in this plot because the salt
703     and salt solution systems contain non-neutral groups, see the
704     accompanying supporting information for a comparison where all groups
705     are neutral.
706 chrisfen 2594
707 chrisfen 2620 Correcting the resulting charged cutoff sphere is one of the purposes
708     of the damped Coulomb summation proposed by Wolf \textit{et
709     al.},\cite{Wolf99} and this correction indeed improves the results as
710     seen in the Shifted-Potental rows. While the undamped case of this
711     method is a significant improvement over the pure cutoff, it still
712     doesn't correlate that well with SPME. Inclusion of potential damping
713     improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
714     an excellent correlation and quality of fit with the SPME results,
715     particularly with a cutoff radius greater than 12 \AA . Use of a
716     larger damping parameter is more helpful for the shortest cutoff
717     shown, but it has a detrimental effect on simulations with larger
718 chrisfen 2629 cutoffs. In the {\sc sf} sets, increasing damping results in
719 chrisfen 2620 progressively poorer correlation. Overall, the undamped case is the
720     best performing set, as the correlation and quality of fits are
721     consistently superior regardless of the cutoff distance. This result
722     is beneficial in that the undamped case is less computationally
723     prohibitive do to the lack of complimentary error function calculation
724     when performing the electrostatic pair interaction. The reaction
725     field results illustrates some of that method's limitations, primarily
726     that it was developed for use in homogenous systems; although it does
727     provide results that are an improvement over those from an unmodified
728     cutoff.
729 chrisfen 2609
730 chrisfen 2608 \subsection{Magnitudes of the Force and Torque Vectors}
731 chrisfen 2599
732 chrisfen 2620 Evaluation of pairwise methods for use in Molecular Dynamics
733     simulations requires consideration of effects on the forces and
734     torques. Investigation of the force and torque vector magnitudes
735     provides a measure of the strength of these values relative to SPME.
736     Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
737     force and torque vector magnitude regression results for the
738     accumulated analysis over all the system types.
739 chrisfen 2594
740     \begin{figure}
741     \centering
742 gezelter 2617 \includegraphics[width=5.5in]{./frcMagplot.pdf}
743 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).}
744 chrisfen 2601 \label{fig:frcMag}
745 chrisfen 2594 \end{figure}
746    
747 chrisfen 2620 Figure \ref{fig:frcMag}, for the most part, parallels the results seen
748     in the previous $\Delta E$ section. The unmodified cutoff results are
749     poor, but using group based cutoffs and a switching function provides
750     a improvement much more significant than what was seen with $\Delta
751 chrisfen 2629 E$. Looking at the {\sc sp} sets, the slope and $R^2$
752 chrisfen 2620 improve with the use of damping to an optimal result of 0.2 \AA
753     $^{-1}$ for the 12 and 15 \AA\ cutoffs. Further increases in damping,
754     while beneficial for simulations with a cutoff radius of 9 \AA\ , is
755     detrimental to simulations with larger cutoff radii. The undamped
756 chrisfen 2629 {\sc sf} method gives forces in line with those obtained using
757 chrisfen 2620 SPME, and use of a damping function results in minor improvement. The
758     reaction field results are surprisingly good, considering the poor
759     quality of the fits for the $\Delta E$ results. There is still a
760     considerable degree of scatter in the data, but it correlates well in
761     general. To be fair, we again note that the reaction field
762     calculations do not encompass NaCl crystal and melt systems, so these
763     results are partly biased towards conditions in which the method
764     performs more favorably.
765 chrisfen 2594
766     \begin{figure}
767     \centering
768 gezelter 2617 \includegraphics[width=5.5in]{./trqMagplot.pdf}
769 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).}
770 chrisfen 2601 \label{fig:trqMag}
771 chrisfen 2594 \end{figure}
772    
773 chrisfen 2620 To evaluate the torque vector magnitudes, the data set from which
774     values are drawn is limited to rigid molecules in the systems
775     (i.e. water molecules). In spite of this smaller sampling pool, the
776     torque vector magnitude results in figure \ref{fig:trqMag} are still
777     similar to those seen for the forces; however, they more clearly show
778     the improved behavior that comes with increasing the cutoff radius.
779 chrisfen 2629 Moderate damping is beneficial to the {\sc sp} and helpful
780     yet possibly unnecessary with the {\sc sf} method, and they also
781 chrisfen 2620 show that over-damping adversely effects all cutoff radii rather than
782     showing an improvement for systems with short cutoffs. The reaction
783     field method performs well when calculating the torques, better than
784     the Shifted Force method over this limited data set.
785 chrisfen 2594
786 chrisfen 2608 \subsection{Directionality of the Force and Torque Vectors}
787 chrisfen 2599
788 chrisfen 2620 Having force and torque vectors with magnitudes that are well
789     correlated to SPME is good, but if they are not pointing in the proper
790     direction the results will be incorrect. These vector directions were
791     investigated through measurement of the angle formed between them and
792     those from SPME. The results (Fig. \ref{fig:frcTrqAng}) are compared
793     through the variance ($\sigma^2$) of the Gaussian fits of the angle
794     error distributions of the combined set over all system types.
795 chrisfen 2594
796     \begin{figure}
797     \centering
798 gezelter 2617 \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
799 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).}
800 chrisfen 2601 \label{fig:frcTrqAng}
801 chrisfen 2594 \end{figure}
802    
803 chrisfen 2620 Both the force and torque $\sigma^2$ results from the analysis of the
804     total accumulated system data are tabulated in figure
805     \ref{fig:frcTrqAng}. All of the sets, aside from the over-damped case
806     show the improvement afforded by choosing a longer simulation cutoff.
807     Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
808     of the distribution widths, with a similar improvement going from 12
809 chrisfen 2629 to 15 \AA . The undamped {\sc sf}, Group Based Cutoff, and
810 chrisfen 2620 Reaction Field methods all do equivalently well at capturing the
811     direction of both the force and torque vectors. Using damping
812 chrisfen 2629 improves the angular behavior significantly for the {\sc sp}
813     and moderately for the {\sc sf} methods. Increasing the damping
814 chrisfen 2620 too far is destructive for both methods, particularly to the torque
815     vectors. Again it is important to recognize that the force vectors
816     cover all particles in the systems, while torque vectors are only
817     available for neutral molecular groups. Damping appears to have a
818     more beneficial effect on non-neutral bodies, and this observation is
819     investigated further in the accompanying supporting information.
820 chrisfen 2594
821 chrisfen 2595 \begin{table}[htbp]
822     \centering
823     \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$^2$.}
824 chrisfen 2599 \begin{tabular}{@{} ccrrrrrrrr @{}}
825 chrisfen 2595 \\
826     \toprule
827     & & \multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted Force} \\
828     \cmidrule(lr){3-6}
829     \cmidrule(l){7-10}
830 chrisfen 2599 $R_\textrm{c}$ & Groups & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ & $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
831 chrisfen 2595 \midrule
832 chrisfen 2599
833     9 \AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
834     & \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} \\
835     12 \AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
836     & \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} \\
837     15 \AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
838     & \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} \\
839 chrisfen 2594
840 chrisfen 2595 \midrule
841    
842 chrisfen 2599 9 \AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
843     & \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} \\
844     12 \AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
845     & \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} \\
846     15 \AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
847     & \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} \\
848 chrisfen 2595
849     \bottomrule
850     \end{tabular}
851 chrisfen 2601 \label{tab:groupAngle}
852 chrisfen 2595 \end{table}
853    
854 chrisfen 2620 Although not discussed previously, group based cutoffs can be applied
855 chrisfen 2629 to both the {\sc sp} and {\sc sf} methods. Use off a
856 chrisfen 2620 switching function corrects for the discontinuities that arise when
857     atoms of a group exit the cutoff before the group's center of mass.
858     Though there are no significant benefit or drawbacks observed in
859     $\Delta E$ and vector magnitude results when doing this, there is a
860     measurable improvement in the vector angle results. Table
861     \ref{tab:groupAngle} shows the angular variance values obtained using
862     group based cutoffs and a switching function alongside the standard
863     results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
864 chrisfen 2629 The {\sc sp} shows much narrower angular distributions for
865 chrisfen 2620 both the force and torque vectors when using an $\alpha$ of 0.2
866 chrisfen 2629 \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
867 chrisfen 2620 undamped and lightly damped cases. Thus, by calculating the
868     electrostatic interactions in terms of molecular pairs rather than
869     atomic pairs, the direction of the force and torque vectors are
870     determined more accurately.
871 chrisfen 2595
872 chrisfen 2620 One additional trend to recognize in table \ref{tab:groupAngle} is
873 chrisfen 2629 that the $\sigma^2$ values for both {\sc sp} and
874     {\sc sf} converge as $\alpha$ increases, something that is easier
875 chrisfen 2620 to see when using group based cutoffs. Looking back on figures
876     \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
877     behavior clearly at large $\alpha$ and cutoff values. The reason for
878     this is that the complimentary error function inserted into the
879     potential weakens the electrostatic interaction as $\alpha$ increases.
880     Thus, at larger values of $\alpha$, both the summation method types
881     progress toward non-interacting functions, so care is required in
882     choosing large damping functions lest one generate an undesirable loss
883     in the pair interaction. Kast \textit{et al.} developed a method for
884     choosing appropriate $\alpha$ values for these types of electrostatic
885     summation methods by fitting to $g(r)$ data, and their methods
886     indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
887     values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
888     to be reasonable choices to obtain proper MC behavior
889     (Fig. \ref{fig:delE}); however, based on these findings, choices this
890     high would introduce error in the molecular torques, particularly for
891     the shorter cutoffs. Based on the above findings, empirical damping
892     up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
893 chrisfen 2629 unnecessary when using the {\sc sf} method.
894 chrisfen 2595
895 chrisfen 2638 \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
896 chrisfen 2601
897 chrisfen 2629 In the previous studies using a {\sc sf} variant of the damped
898 chrisfen 2620 Wolf coulomb potential, the structure and dynamics of water were
899     investigated rather extensively.\cite{Zahn02,Kast03} Their results
900 chrisfen 2629 indicated that the damped {\sc sf} method results in properties
901 chrisfen 2620 very similar to those obtained when using the Ewald summation.
902     Considering the statistical results shown above, the good performance
903     of this method is not that surprising. Rather than consider the same
904     systems and simply recapitulate their results, we decided to look at
905     the solid state dynamical behavior obtained using the best performing
906     summation methods from the above results.
907 chrisfen 2601
908     \begin{figure}
909     \centering
910 chrisfen 2638 \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
911     \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.}
912     \label{fig:vCorrPlot}
913     \end{figure}
914    
915     The short-time decays through the first collision are nearly identical
916     in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
917     functions show how the methods differ. The undamped {\sc sf} method
918     has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
919     peaks than any of the other methods. As the damping function is
920     increased, these peaks are smoothed out, and approach the SPME
921     curve. The damping acts as a distance dependent Gaussian screening of
922     the point charges for the pairwise summation methods; thus, the
923     collisions are more elastic in the undamped {\sc sf} potental, and the
924     stiffness of the potential is diminished as the electrostatic
925     interactions are softened by the damping function. With $\alpha$
926     values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
927     nearly identical and track the SPME features quite well. This is not
928     too surprising in that the differences between the {\sc sf} and {\sc
929     sp} potentials are mitigated with increased damping. However, this
930     appears to indicate that once damping is utilized, the form of the
931     potential seems to play a lesser role in the crystal dynamics.
932    
933     \subsection{Collective Motion: Power Spectra of NaCl Crystals}
934    
935     The short time dynamics were extended to evaluate how the differences
936     between the methods affect the collective long-time motion. The same
937     electrostatic summation methods were used as in the short time
938     velocity autocorrelation function evaluation, but the trajectories
939     were sampled over a much longer time. The power spectra of the
940     resulting velocity autocorrelation functions were calculated and are
941     displayed in figure \ref{fig:methodPS}.
942    
943     \begin{figure}
944     \centering
945 gezelter 2617 \includegraphics[width = \linewidth]{./spectraSquare.pdf}
946 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.}
947 chrisfen 2610 \label{fig:methodPS}
948 chrisfen 2601 \end{figure}
949    
950 chrisfen 2638 While high frequency peaks of the spectra in this figure overlap,
951     showing the same general features, the low frequency region shows how
952     the summation methods differ. Considering the low-frequency inset
953     (expanded in the upper frame of figure \ref{fig:dampInc}), at
954     frequencies below 100 cm$^{-1}$, the correlated motions are
955     blue-shifted when using undamped or weakly damped {\sc sf}. When
956     using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
957     and {\sc sp} methods give near identical correlated motion behavior as
958     the Ewald method (which has a damping value of 0.3119). This
959     weakening of the electrostatic interaction with increased damping
960     explains why the long-ranged correlated motions are at lower
961     frequencies for the moderately damped methods than for undamped or
962     weakly damped methods. To see this effect more clearly, we show how
963     damping strength alone affects a simple real-space electrostatic
964     potential,
965 chrisfen 2601 \begin{equation}
966 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),
967 chrisfen 2601 \end{equation}
968 chrisfen 2620 where $S(r)$ is a switching function that smoothly zeroes the
969     potential at the cutoff radius. Figure \ref{fig:dampInc} shows how
970     the low frequency motions are dependent on the damping used in the
971     direct electrostatic sum. As the damping increases, the peaks drop to
972     lower frequencies. Incidentally, use of an $\alpha$ of 0.25
973     \AA$^{-1}$ on a simple electrostatic summation results in low
974     frequency correlated dynamics equivalent to a simulation using SPME.
975     When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
976     shift to higher frequency in exponential fashion. Though not shown,
977     the spectrum for the simple undamped electrostatic potential is
978     blue-shifted such that the lowest frequency peak resides near 325
979 chrisfen 2638 cm$^{-1}$. In light of these results, the undamped {\sc sf} method
980     producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
981     respectable and shows that the shifted force procedure accounts for
982     most of the effect afforded through use of the Ewald summation.
983     However, it appears as though moderate damping is required for
984     accurate reproduction of crystal dynamics.
985 chrisfen 2601 \begin{figure}
986     \centering
987 gezelter 2617 \includegraphics[width = \linewidth]{./comboSquare.pdf}
988 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.}
989 chrisfen 2601 \label{fig:dampInc}
990     \end{figure}
991    
992 chrisfen 2575 \section{Conclusions}
993    
994 chrisfen 2620 This investigation of pairwise electrostatic summation techniques
995     shows that there are viable and more computationally efficient
996     electrostatic summation techniques than the Ewald summation, chiefly
997     methods derived from the damped Coulombic sum originally proposed by
998     Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
999 chrisfen 2629 {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1000 chrisfen 2620 shows a remarkable ability to reproduce the energetic and dynamic
1001     characteristics exhibited by simulations employing lattice summation
1002     techniques. The cumulative energy difference results showed the
1003 chrisfen 2629 undamped {\sc sf} and moderately damped {\sc sp} methods
1004 chrisfen 2620 produced results nearly identical to SPME. Similarly for the dynamic
1005 chrisfen 2629 features, the undamped or moderately damped {\sc sf} and
1006     moderately damped {\sc sp} methods produce force and torque
1007 chrisfen 2620 vector magnitude and directions very similar to the expected values.
1008     These results translate into long-time dynamic behavior equivalent to
1009     that produced in simulations using SPME.
1010 chrisfen 2604
1011 chrisfen 2620 Aside from the computational cost benefit, these techniques have
1012     applicability in situations where the use of the Ewald sum can prove
1013     problematic. Primary among them is their use in interfacial systems,
1014     where the unmodified lattice sum techniques artificially accentuate
1015     the periodicity of the system in an undesirable manner. There have
1016     been alterations to the standard Ewald techniques, via corrections and
1017     reformulations, to compensate for these systems; but the pairwise
1018     techniques discussed here require no modifications, making them
1019     natural tools to tackle these problems. Additionally, this
1020     transferability gives them benefits over other pairwise methods, like
1021     reaction field, because estimations of physical properties (e.g. the
1022     dielectric constant) are unnecessary.
1023 chrisfen 2605
1024 chrisfen 2620 We are not suggesting any flaw with the Ewald sum; in fact, it is the
1025     standard by which these simple pairwise sums are judged. However,
1026     these results do suggest that in the typical simulations performed
1027     today, the Ewald summation may no longer be required to obtain the
1028 chrisfen 2638 level of accuracy most researchers have come to expect
1029 chrisfen 2605
1030 chrisfen 2575 \section{Acknowledgments}
1031 chrisfen 2594 \newpage
1032    
1033 gezelter 2617 \bibliographystyle{jcp2}
1034 chrisfen 2575 \bibliography{electrostaticMethods}
1035    
1036    
1037     \end{document}