ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/dissertation.tex
(Generate patch)

Comparing trunk/fennellDissertation/dissertation.tex (file contents):
Revision 2971 by chrisfen, Sun Aug 6 22:29:27 2006 UTC vs.
Revision 2973 by chrisfen, Fri Aug 18 15:49:29 2006 UTC

# Line 43 | Line 43 | STUDY OF WATER}    
43  
44   \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
45  
46 < \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION
47 < TECHNIQUES}
48 <
49 < In molecular simulations, proper accumulation of the electrostatic
50 < interactions is essential and is one of the most
51 < computationally-demanding tasks.  The common molecular mechanics force
52 < fields represent atomic sites with full or partial charges protected
53 < by Lennard-Jones (short range) interactions.  This means that nearly
54 < every pair interaction involves a calculation of charge-charge forces.
55 < Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
56 < interactions quickly become the most expensive part of molecular
57 < simulations.  Historically, the electrostatic pair interaction would
58 < not have decayed appreciably within the typical box lengths that could
59 < be feasibly simulated.  In the larger systems that are more typical of
60 < modern simulations, large cutoffs should be used to incorporate
61 < electrostatics correctly.
62 <
63 < There have been many efforts to address the proper and practical
64 < handling of electrostatic interactions, and these have resulted in a
65 < variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
66 < typically classified as implicit methods (i.e., continuum dielectrics,
67 < static dipolar fields),\cite{Born20,Grossfield00} explicit methods
68 < (i.e., Ewald summations, interaction shifting or
69 < truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
70 < reaction field type methods, fast multipole
71 < methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
72 < often preferred because they physically incorporate solvent molecules
73 < in the system of interest, but these methods are sometimes difficult
74 < to utilize because of their high computational cost.\cite{Roux99} In
75 < addition to the computational cost, there have been some questions
76 < regarding possible artifacts caused by the inherent periodicity of the
77 < explicit Ewald summation.\cite{Tobias01}
78 <
79 < In this chapter, we focus on a new set of pairwise methods devised by
80 < Wolf {\it et al.},\cite{Wolf99} which we further extend.  These
81 < methods along with a few other mixed methods (i.e. reaction field) are
82 < compared with the smooth particle mesh Ewald
83 < sum,\cite{Onsager36,Essmann99} which is our reference method for
84 < handling long-range electrostatic interactions. The new methods for
85 < handling electrostatics have the potential to scale linearly with
86 < increasing system size since they involve only a simple modification
87 < to the direct pairwise sum.  They also lack the added periodicity of
88 < the Ewald sum, so they can be used for systems which are non-periodic
89 < or which have one- or two-dimensional periodicity.  Below, these
90 < methods are evaluated using a variety of model systems to
91 < establish their usability in molecular simulations.
92 <
93 < \section{The Ewald Sum}
94 <
95 < The complete accumulation of the electrostatic interactions in a system with
96 < periodic boundary conditions (PBC) requires the consideration of the
97 < effect of all charges within a (cubic) simulation box as well as those
98 < in the periodic replicas,
99 < \begin{equation}
100 < V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
101 < \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
102 < \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
103 < \right],
104 < \label{eq:PBCSum}
105 < \end{equation}
106 < where the sum over $\mathbf{n}$ is a sum over all periodic box
107 < replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
108 < prime indicates $i = j$ are neglected for $\mathbf{n} =
109 < 0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
110 < particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
111 < the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
112 < $j$, and $\phi$ is the solution to Poisson's equation
113 < ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
114 < charge-charge interactions). In the case of monopole electrostatics,
115 < eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
116 < non-neutral systems.
117 <
118 < The electrostatic summation problem was originally studied by Ewald
119 < for the case of an infinite crystal.\cite{Ewald21}. The approach he
120 < took was to convert this conditionally convergent sum into two
121 < absolutely convergent summations: a short-ranged real-space summation
122 < and a long-ranged reciprocal-space summation,
123 < \begin{equation}
124 < \begin{split}
125 < V_\textrm{elec} = \frac{1}{2}&
126 < \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
127 < \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
128 < {|\mathbf{r}_{ij}+\mathbf{n}|} \\
129 < &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
130 < \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
131 < \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
132 < &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
133 < + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
134 < \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
135 < \end{split}
136 < \label{eq:EwaldSum}
137 < \end{equation}
138 < where $\alpha$ is the damping or convergence parameter with units of
139 < \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
140 < $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
141 < constant of the surrounding medium. The final two terms of
142 < eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
143 < for interacting with a surrounding dielectric.\cite{Allen87} This
144 < dipolar term was neglected in early applications in molecular
145 < simulations,\cite{Brush66,Woodcock71} until it was introduced by de
146 < Leeuw {\it et al.} to address situations where the unit cell has a
147 < dipole moment which is magnified through replication of the periodic
148 < images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
149 < system is said to be using conducting (or ``tin-foil'') boundary
150 < conditions, $\epsilon_{\rm S} = \infty$. Figure
151 < \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
152 < time.  Initially, due to the small system sizes that could be
153 < simulated feasibly, the entire simulation box was replicated to
154 < convergence.  In more modern simulations, the systems have grown large
155 < enough that a real-space cutoff could potentially give convergent
156 < behavior.  Indeed, it has been observed that with the choice of a
157 < small $\alpha$, the reciprocal-space portion of the Ewald sum can be
158 < rapidly convergent and small relative to the real-space
159 < portion.\cite{Karasawa89,Kolafa92}
160 <
161 < \begin{figure}
162 < \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
163 < \caption{The change in the need for the Ewald sum with
164 < increasing computational power.  A:~Initially, only small systems
165 < could be studied, and the Ewald sum replicated the simulation box to
166 < convergence.  B:~Now, radial cutoff methods should be able to reach
167 < convergence for the larger systems of charges that are common today.}
168 < \label{fig:ewaldTime}
169 < \end{figure}
170 <
171 < The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The
172 < convergence parameter $(\alpha)$ plays an important role in balancing
173 < the computational cost between the direct and reciprocal-space
174 < portions of the summation.  The choice of this value allows one to
175 < select whether the real-space or reciprocal space portion of the
176 < summation is an $\mathscr{O}(N^2)$ calculation (with the other being
177 < $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
178 < $\alpha$ and thoughtful algorithm development, this cost can be
179 < reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
180 < taken to reduce the cost of the Ewald summation even further is to set
181 < $\alpha$ such that the real-space interactions decay rapidly, allowing
182 < for a short spherical cutoff. Then the reciprocal space summation is
183 < optimized.  These optimizations usually involve utilization of the
184 < fast Fourier transform (FFT),\cite{Hockney81} leading to the
185 < particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
186 < methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
187 < methods, the cost of the reciprocal-space portion of the Ewald
188 < summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
189 < \log N)$.
190 <
191 < These developments and optimizations have made the use of the Ewald
192 < summation routine in simulations with periodic boundary
193 < conditions. However, in certain systems, such as vapor-liquid
194 < interfaces and membranes, the intrinsic three-dimensional periodicity
195 < can prove problematic.  The Ewald sum has been reformulated to handle
196 < 2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
197 < methods are computationally expensive.\cite{Spohr97,Yeh99} More
198 < recently, there have been several successful efforts toward reducing
199 < the computational cost of 2-D lattice
200 < summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
201 < bringing them more in line with the cost of the full 3-D summation.
202 <
203 < Several studies have recognized that the inherent periodicity in the
204 < Ewald sum can also have an effect on three-dimensional
205 < systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
206 < Solvated proteins are essentially kept at high concentration due to
207 < the periodicity of the electrostatic summation method.  In these
208 < systems, the more compact folded states of a protein can be
209 < artificially stabilized by the periodic replicas introduced by the
210 < Ewald summation.\cite{Weber00} Thus, care must be taken when
211 < considering the use of the Ewald summation where the assumed
212 < periodicity would introduce spurious effects in the system dynamics.
213 <
214 <
215 < \section{The Wolf and Zahn Methods}
216 <
217 < In a recent paper by Wolf \textit{et al.}, a procedure was outlined
218 < for the accurate accumulation of electrostatic interactions in an
219 < efficient pairwise fashion.  This procedure lacks the inherent
220 < periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
221 < observed that the electrostatic interaction is effectively
222 < short-ranged in condensed phase systems and that neutralization of the
223 < charge contained within the cutoff radius is crucial for potential
224 < stability. They devised a pairwise summation method that ensures
225 < charge neutrality and gives results similar to those obtained with the
226 < Ewald summation.  The resulting shifted Coulomb potential includes
227 < image-charges subtracted out through placement on the cutoff sphere
228 < and a distance-dependent damping function (identical to that seen in
229 < the real-space portion of the Ewald sum) to aid convergence
230 < \begin{equation}
231 < V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
232 < - \lim_{r_{ij}\rightarrow R_\textrm{c}}
233 < \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
234 < \label{eq:WolfPot}
235 < \end{equation}
236 < Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
237 < potential.  However, neutralizing the charge contained within each
238 < cutoff sphere requires the placement of a self-image charge on the
239 < surface of the cutoff sphere.  This additional self-term in the total
240 < potential enabled Wolf {\it et al.}  to obtain excellent estimates of
241 < Madelung energies for many crystals.
242 <
243 < In order to use their charge-neutralized potential in molecular
244 < dynamics simulations, Wolf \textit{et al.} suggested taking the
245 < derivative of this potential prior to evaluation of the limit.  This
246 < procedure gives an expression for the forces,
247 < \begin{equation}
248 < \begin{split}
249 < F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
250 < \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
251 < + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
252 < \Biggr]\\
253 < &-\Biggr[
254 < \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
255 < + \frac{2\alpha}{\pi^{1/2}}
256 < \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
257 < \Biggr]\Biggr\},
258 < \end{split}
259 < \label{eq:WolfForces}
260 < \end{equation}
261 < that incorporates both image charges and damping of the electrostatic
262 < interaction.
263 <
264 < More recently, Zahn \textit{et al.} investigated these potential and
265 < force expressions for use in simulations involving water.\cite{Zahn02}
266 < In their work, they pointed out that the forces and derivative of
267 < the potential are not commensurate.  Attempts to use both
268 < eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
269 < to poor energy conservation.  They correctly observed that taking the
270 < limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
271 < derivatives gives forces for a different potential energy function
272 < than the one shown in eq. (\ref{eq:WolfPot}).
273 <
274 < Zahn \textit{et al.} introduced a modified form of this summation
275 < method as a way to use the technique in Molecular Dynamics
276 < simulations.  They proposed a new damped Coulomb potential,
277 < \begin{equation}
278 < \begin{split}
279 < V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
280 < \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
281 < &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
282 < + \frac{2\alpha}{\pi^{1/2}}
283 < \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
284 < \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
285 < \end{split}
286 < \label{eq:ZahnPot}
287 < \end{equation}
288 < and showed that this potential does fairly well at capturing the
289 < structural and dynamic properties of water compared the same
290 < properties obtained using the Ewald sum.
291 <
292 < \section{Simple Forms for Pairwise Electrostatics}\label{sec:PairwiseDerivation}
293 <
294 < The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
295 < al.} are constructed using two different (and separable) computational
296 < tricks:
297 <
298 < \begin{enumerate}[itemsep=0pt]
299 < \item shifting through the use of image charges, and
300 < \item damping the electrostatic interaction.
301 < \end{enumerate}  
302 < Wolf \textit{et al.} treated the development of their summation method
303 < as a progressive application of these techniques,\cite{Wolf99} while
304 < Zahn \textit{et al.} founded their damped Coulomb modification
305 < (Eq. (\ref{eq:ZahnPot})) on the post-limit forces
306 < (Eq. (\ref{eq:WolfForces})) which were derived using both techniques.
307 < It is possible, however, to separate these tricks and study their
308 < effects independently.
309 <
310 < Starting with the original observation that the effective range of the
311 < electrostatic interaction in condensed phases is considerably less
312 < than $r^{-1}$, either the cutoff sphere neutralization or the
313 < distance-dependent damping technique could be used as a foundation for
314 < a new pairwise summation method.  Wolf \textit{et al.} made the
315 < observation that charge neutralization within the cutoff sphere plays
316 < a significant role in energy convergence; therefore we will begin our
317 < analysis with the various shifted forms that maintain this charge
318 < neutralization.  We can evaluate the methods of Wolf
319 < \textit{et al.}  and Zahn \textit{et al.} by considering the standard
320 < shifted potential,
321 < \begin{equation}
322 < V_\textrm{SP}(r) =      \begin{cases}
323 < v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
324 < R_\textrm{c}  
325 < \end{cases},
326 < \label{eq:shiftingPotForm}
327 < \end{equation}
328 < and shifted force,
329 < \begin{equation}
330 < V_\textrm{SF}(r) = \begin{cases}
331 < v(r) - v_\textrm{c}
332 < - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
333 < &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
334 < \end{cases},
335 < \label{eq:shiftingForm}
336 < \end{equation}
337 < functions where $v(r)$ is the unshifted form of the potential, and
338 < $v_c$ is $v(R_\textrm{c})$.  The Shifted Force ({\sc sf}) form ensures
339 < that both the potential and the forces goes to zero at the cutoff
340 < radius, while the Shifted Potential ({\sc sp}) form only ensures the
341 < potential is smooth at the cutoff radius
342 < ($R_\textrm{c}$).\cite{Allen87}
343 <
344 < The forces associated with the shifted potential are simply the forces
345 < of the unshifted potential itself (when inside the cutoff sphere),
346 < \begin{equation}
347 < F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
348 < \end{equation}
349 < and are zero outside.  Inside the cutoff sphere, the forces associated
350 < with the shifted force form can be written,
351 < \begin{equation}
352 < F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
353 < v(r)}{dr} \right)_{r=R_\textrm{c}}.
354 < \end{equation}
355 <
356 < If the potential, $v(r)$, is taken to be the normal Coulomb potential,
357 < \begin{equation}
358 < v(r) = \frac{q_i q_j}{r},
359 < \label{eq:Coulomb}
360 < \end{equation}
361 < then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
362 < al.}'s undamped prescription:
363 < \begin{equation}
364 < V_\textrm{SP}(r) =
365 < q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
366 < r\leqslant R_\textrm{c},
367 < \label{eq:SPPot}
368 < \end{equation}
369 < with associated forces,
370 < \begin{equation}
371 < F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
372 < \quad r\leqslant R_\textrm{c}.
373 < \label{eq:SPForces}
374 < \end{equation}
375 < These forces are identical to the forces of the standard Coulomb
376 < interaction, and cutting these off at $R_c$ was addressed by Wolf
377 < \textit{et al.} as undesirable.  They pointed out that the effect of
378 < the image charges is neglected in the forces when this form is
379 < used,\cite{Wolf99} thereby eliminating any benefit from the method in
380 < molecular dynamics.  Additionally, there is a discontinuity in the
381 < forces at the cutoff radius which results in energy drift during MD
382 < simulations.
383 <
384 < The shifted force ({\sc sf}) form using the normal Coulomb potential
385 < will give,
386 < \begin{equation}
387 < V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
388 < + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
389 < \quad r\leqslant R_\textrm{c}.
390 < \label{eq:SFPot}
391 < \end{equation}
392 < with associated forces,
393 < \begin{equation}
394 < F_\textrm{SF}(r) =  q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
395 < \quad r\leqslant R_\textrm{c}.
396 < \label{eq:SFForces}
397 < \end{equation}
398 < This formulation has the benefits that there are no discontinuities at
399 < the cutoff radius, while the neutralizing image charges are present in
400 < both the energy and force expressions.  It would be simple to add the
401 < self-neutralizing term back when computing the total energy of the
402 < system, thereby maintaining the agreement with the Madelung energies.
403 < A side effect of this treatment is the alteration in the shape of the
404 < potential that comes from the derivative term.  Thus, a degree of
405 < clarity about agreement with the empirical potential is lost in order
406 < to gain functionality in dynamics simulations.
407 <
408 < Wolf \textit{et al.} originally discussed the energetics of the
409 < shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
410 < insufficient for accurate determination of the energy with reasonable
411 < cutoff distances.  The calculated Madelung energies fluctuated around
412 < the expected value as the cutoff radius was increased, but the
413 < oscillations converged toward the correct value.\cite{Wolf99} A
414 < damping function was incorporated to accelerate the convergence; and
415 < though alternative forms for the damping function could be
416 < used,\cite{Jones56,Heyes81} the complimentary error function was
417 < chosen to mirror the effective screening used in the Ewald summation.
418 < Incorporating this error function damping into the simple Coulomb
419 < potential,
420 < \begin{equation}
421 < v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
422 < \label{eq:dampCoulomb}
423 < \end{equation}
424 < the shifted potential (eq. (\ref{eq:SPPot})) becomes
425 < \begin{equation}
426 < V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
427 < - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
428 < \quad r\leqslant R_\textrm{c},
429 < \label{eq:DSPPot}
430 < \end{equation}
431 < with associated forces,
432 < \begin{equation}
433 < F_{\textrm{DSP}}(r) = q_iq_j
434 < \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
435 < + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
436 < \quad r\leqslant R_\textrm{c}.
437 < \label{eq:DSPForces}
438 < \end{equation}
439 < Again, this damped shifted potential suffers from a
440 < force-discontinuity at the cutoff radius, and the image charges play
441 < no role in the forces.  To remedy these concerns, one may derive a
442 < {\sc sf} variant by including the derivative term in
443 < eq. (\ref{eq:shiftingForm}),
444 < \begin{equation}
445 < \begin{split}
446 < V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
447 < \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
448 < - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
449 < &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
450 < + \frac{2\alpha}{\pi^{1/2}}
451 < \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
452 < \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
453 < \quad r\leqslant R_\textrm{c}.
454 < \label{eq:DSFPot}
455 < \end{split}
456 < \end{equation}
457 < The derivative of the above potential will lead to the following forces,
458 < \begin{equation}
459 < \begin{split}
460 < F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
461 < \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
462 < + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
463 < &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
464 < {R_{\textrm{c}}^2}
465 < + \frac{2\alpha}{\pi^{1/2}}
466 < \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
467 < \right)\Biggr{]}
468 < \quad r\leqslant R_\textrm{c}.
469 < \label{eq:DSFForces}
470 < \end{split}
471 < \end{equation}
472 < If the damping parameter $(\alpha)$ is set to zero, the undamped case,
473 < eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
474 < recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
475 <
476 < This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
477 < derived by Zahn \textit{et al.}; however, there are two important
478 < differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
479 < eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
480 < with $r$ replaced by $R_\textrm{c}$.  This term is {\it not} present
481 < in the Zahn potential, resulting in a potential discontinuity as
482 < particles cross $R_\textrm{c}$.  Second, the sign of the derivative
483 < portion is different.  The missing $v_\textrm{c}$ term would not
484 < affect molecular dynamics simulations (although the computed energy
485 < would be expected to have sudden jumps as particle distances crossed
486 < $R_c$).  The sign problem is a potential source of errors, however.
487 < In fact, it introduces a discontinuity in the forces at the cutoff,
488 < because the force function is shifted in the wrong direction and
489 < doesn't cross zero at $R_\textrm{c}$.
490 <
491 < Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
492 < electrostatic summation method in which the potential and forces are
493 < continuous at the cutoff radius and which incorporates the damping
494 < function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
495 < this paper, we will evaluate exactly how good these methods ({\sc sp},
496 < {\sc sf}, damping) are at reproducing the correct electrostatic
497 < summation performed by the Ewald sum.
498 <
499 <
500 < \section{Evaluating Pairwise Summation Techniques}
501 <
502 < In classical molecular mechanics simulations, there are two primary
503 < techniques utilized to obtain information about the system of
504 < interest: Monte Carlo (MC) and Molecular Dynamics (MD).  Both of these
505 < techniques utilize pairwise summations of interactions between
506 < particle sites, but they use these summations in different ways.
507 <
508 < In MC, the potential energy difference between configurations dictates
509 < the progression of MC sampling.  Going back to the origins of this
510 < method, the acceptance criterion for the canonical ensemble laid out
511 < by Metropolis \textit{et al.} states that a subsequent configuration
512 < is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
513 < $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
514 < Maintaining the correct $\Delta E$ when using an alternate method for
515 < handling the long-range electrostatics will ensure proper sampling
516 < from the ensemble.
517 <
518 < In MD, the derivative of the potential governs how the system will
519 < progress in time.  Consequently, the force and torque vectors on each
520 < body in the system dictate how the system evolves.  If the magnitude
521 < and direction of these vectors are similar when using alternate
522 < electrostatic summation techniques, the dynamics in the short term
523 < will be indistinguishable.  Because error in MD calculations is
524 < cumulative, one should expect greater deviation at longer times,
525 < although methods which have large differences in the force and torque
526 < vectors will diverge from each other more rapidly.
527 <
528 < \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
529 <
530 < The pairwise summation techniques (outlined in section
531 < \ref{sec:ESMethods}) were evaluated for use in MC simulations by
532 < studying the energy differences between conformations.  We took the
533 < {\sc spme}-computed energy difference between two conformations to be the
534 < correct behavior. An ideal performance by an alternative method would
535 < reproduce these energy differences exactly (even if the absolute
536 < energies calculated by the methods are different).  Since none of the
537 < methods provide exact energy differences, we used linear least squares
538 < regressions of energy gap data to evaluate how closely the methods
539 < mimicked the Ewald energy gaps.  Unitary results for both the
540 < correlation (slope) and correlation coefficient for these regressions
541 < indicate perfect agreement between the alternative method and {\sc spme}.
542 < Sample correlation plots for two alternate methods are shown in
543 < Fig. \ref{fig:linearFit}.
544 <
545 < \begin{figure}
546 < \centering
547 < \includegraphics[width = \linewidth]{./figures/dualLinear.pdf}
548 < \caption{Example least squares regressions of the configuration energy
549 < differences for SPC/E water systems. The upper plot shows a data set
550 < with a poor correlation coefficient ($R^2$), while the lower plot
551 < shows a data set with a good correlation coefficient.}
552 < \label{fig:linearFit}
553 < \end{figure}
554 <
555 < Each of the seven system types (detailed in section \ref{sec:RepSims})
556 < were represented using 500 independent configurations.  Thus, each of
557 < the alternative (non-Ewald) electrostatic summation methods was
558 < evaluated using an accumulated 873,250 configurational energy
559 < differences.
560 <
561 < Results and discussion for the individual analysis of each of the
562 < system types appear in sections \ref{sec:IndividualResults}, while the
563 < cumulative results over all the investigated systems appear below in
564 < sections \ref{sec:EnergyResults}.
565 <
566 < \subsection{Molecular Dynamics and the Force and Torque
567 < Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
568 < (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
569 < comparing the force and torque vectors with those obtained using the
570 < reference Ewald summation ({\sc spme}).  Both the magnitude and the
571 < direction of these vectors on each of the bodies in the system were
572 < analyzed.  For the magnitude of these vectors, linear least squares
573 < regression analyses were performed as described previously for
574 < comparing $\Delta E$ values.  Instead of a single energy difference
575 < between two system configurations, we compared the magnitudes of the
576 < forces (and torques) on each molecule in each configuration.  For a
577 < system of 1000 water molecules and 40 ions, there are 1040 force
578 < vectors and 1000 torque vectors.  With 500 configurations, this
579 < results in 520,000 force and 500,000 torque vector comparisons.
580 < Additionally, data from seven different system types was aggregated
581 < before the comparison was made.
582 <
583 < The {\it directionality} of the force and torque vectors was
584 < investigated through measurement of the angle ($\theta$) formed
585 < between those computed from the particular method and those from {\sc spme},
586 < \begin{equation}
587 < \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
588 < \cdot \hat{F}_\textrm{M}\right),
589 < \end{equation}
590 < where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
591 < vector computed using method M.  Each of these $\theta$ values was
592 < accumulated in a distribution function and weighted by the area on the
593 < unit sphere.  Since this distribution is a measure of angular error
594 < between two different electrostatic summation methods, there is no
595 < {\it a priori} reason for the profile to adhere to any specific
596 < shape. Thus, gaussian fits were used to measure the width of the
597 < resulting distributions. The variance ($\sigma^2$) was extracted from
598 < each of these fits and was used to compare distribution widths.
599 < Values of $\sigma^2$ near zero indicate vector directions
600 < indistinguishable from those calculated when using the reference
601 < method ({\sc spme}).
602 <
603 < \subsection{Short-time Dynamics}
604 <
605 < The effects of the alternative electrostatic summation methods on the
606 < short-time dynamics of charged systems were evaluated by considering a
607 < NaCl crystal at a temperature of 1000 K.  A subset of the best
608 < performing pairwise methods was used in this comparison.  The NaCl
609 < crystal was chosen to avoid possible complications from the treatment
610 < of orientational motion in molecular systems.  All systems were
611 < started with the same initial positions and velocities.  Simulations
612 < were performed under the microcanonical ensemble, and velocity
613 < autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
614 < of the trajectories,
615 < \begin{equation}
616 < C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
617 < \label{eq:vCorr}
618 < \end{equation}
619 < Velocity autocorrelation functions require detailed short time data,
620 < thus velocity information was saved every 2 fs over 10 ps
621 < trajectories. Because the NaCl crystal is composed of two different
622 < atom types, the average of the two resulting velocity autocorrelation
623 < functions was used for comparisons.
624 <
625 < \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
626 <
627 < The effects of the same subset of alternative electrostatic methods on
628 < the {\it long-time} dynamics of charged systems were evaluated using
629 < the same model system (NaCl crystals at 1000K).  The power spectrum
630 < ($I(\omega)$) was obtained via Fourier transform of the velocity
631 < autocorrelation function, \begin{equation} I(\omega) =
632 < \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
633 < \label{eq:powerSpec}
634 < \end{equation}
635 < where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
636 < NaCl crystal is composed of two different atom types, the average of
637 < the two resulting power spectra was used for comparisons. Simulations
638 < were performed under the microcanonical ensemble, and velocity
639 < information was saved every 5~fs over 100~ps trajectories.
640 <
641 < \subsection{Representative Simulations}\label{sec:RepSims}
642 < A variety of representative molecular simulations were analyzed to
643 < determine the relative effectiveness of the pairwise summation
644 < techniques in reproducing the energetics and dynamics exhibited by
645 < {\sc spme}.  We wanted to span the space of typical molecular
646 < simulations (i.e. from liquids of neutral molecules to ionic
647 < crystals), so the systems studied were:
648 <
649 < \begin{enumerate}[itemsep=0pt]
650 < \item liquid water (SPC/E),\cite{Berendsen87}
651 < \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
652 < \item NaCl crystals,
653 < \item NaCl melts,
654 < \item a low ionic strength solution of NaCl in water (0.11 M),
655 < \item a high ionic strength solution of NaCl in water (1.1 M), and
656 < \item a 6\AA\  radius sphere of Argon in water.
657 < \end{enumerate}
658 <
659 < By utilizing the pairwise techniques (outlined in section
660 < \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
661 < charged particles, and mixtures of the two, we hope to discern under
662 < which conditions it will be possible to use one of the alternative
663 < summation methodologies instead of the Ewald sum.
664 <
665 < For the solid and liquid water configurations, configurations were
666 < taken at regular intervals from high temperature trajectories of 1000
667 < SPC/E water molecules.  Each configuration was equilibrated
668 < independently at a lower temperature (300K for the liquid, 200K for
669 < the crystal).  The solid and liquid NaCl systems consisted of 500
670 < $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions.  Configurations for
671 < these systems were selected and equilibrated in the same manner as the
672 < water systems. In order to introduce measurable fluctuations in the
673 < configuration energy differences, the crystalline simulations were
674 < equilibrated at 1000K, near the $T_\textrm{m}$ for NaCl. The liquid
675 < NaCl configurations needed to represent a fully disordered array of
676 < point charges, so the high temperature of 7000K was selected for
677 < equilibration. The ionic solutions were made by solvating 4 (or 40)
678 < ions in a periodic box containing 1000 SPC/E water molecules.  Ion and
679 < water positions were then randomly swapped, and the resulting
680 < configurations were again equilibrated individually.  Finally, for the
681 < Argon / Water ``charge void'' systems, the identities of all the SPC/E
682 < waters within 6\AA\ of the center of the equilibrated water
683 < configurations were converted to argon.
684 <
685 < These procedures guaranteed us a set of representative configurations
686 < from chemically-relevant systems sampled from appropriate
687 < ensembles. Force field parameters for the ions and Argon were taken
688 < from the force field utilized by {\sc oopse}.\cite{Meineke05}
689 <
690 < \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
691 < We compared the following alternative summation methods with results
692 < from the reference method ({\sc spme}):
693 <
694 < \begin{enumerate}[itemsep=0pt]
695 < \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
696 < and 0.3\AA$^{-1}$,
697 < \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
698 < and 0.3\AA$^{-1}$,
699 < \item reaction field with an infinite dielectric constant, and
700 < \item an unmodified cutoff.
701 < \end{enumerate}
702 <
703 < Group-based cutoffs with a fifth-order polynomial switching function
704 < were utilized for the reaction field simulations.  Additionally, we
705 < investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
706 < cutoff.  The {\sc spme} electrostatics were performed using the {\sc tinker}
707 < implementation of {\sc spme},\cite{Ponder87} while all other calculations
708 < were performed using the {\sc oopse} molecular mechanics
709 < package.\cite{Meineke05} All other portions of the energy calculation
710 < (i.e. Lennard-Jones interactions) were handled in exactly the same
711 < manner across all systems and configurations.
712 <
713 < The alternative methods were also evaluated with three different
714 < cutoff radii (9, 12, and 15\AA).  As noted previously, the
715 < convergence parameter ($\alpha$) plays a role in the balance of the
716 < real-space and reciprocal-space portions of the Ewald calculation.
717 < Typical molecular mechanics packages set this to a value dependent on
718 < the cutoff radius and a tolerance (typically less than $1 \times
719 < 10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with
720 < increasing accuracy at the expense of computational time spent on the
721 < reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
722 < The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
723 < in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
724 < 0.3119, and 0.2476\AA$^{-1}$ for cutoff radii of 9, 12, and 15\AA\
725 < respectively.
726 <
727 < \section{Configuration Energy Difference Results}\label{sec:EnergyResults}
728 < In order to evaluate the performance of the pairwise electrostatic
729 < summation methods for Monte Carlo (MC) simulations, the energy
730 < differences between configurations were compared to the values
731 < obtained when using {\sc spme}.  The results for the combined
732 < regression analysis of all of the systems are shown in figure
733 < \ref{fig:delE}.
734 <
735 < \begin{figure}
736 < \centering
737 < \includegraphics[width=4.75in]{./figures/delEplot.pdf}
738 < \caption{Statistical analysis of the quality of configurational energy
739 < differences for a given electrostatic method compared with the
740 < reference Ewald sum.  Results with a value equal to 1 (dashed line)
741 < indicate $\Delta E$ values indistinguishable from those obtained using
742 < {\sc spme}.  Different values of the cutoff radius are indicated with
743 < different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
744 < inverted triangles).}
745 < \label{fig:delE}
746 < \end{figure}
747 <
748 < The most striking feature of this plot is how well the Shifted Force
749 < ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
750 < differences.  For the undamped {\sc sf} method, and the
751 < moderately-damped {\sc sp} methods, the results are nearly
752 < indistinguishable from the Ewald results.  The other common methods do
753 < significantly less well.
754 <
755 < The unmodified cutoff method is essentially unusable.  This is not
756 < surprising since hard cutoffs give large energy fluctuations as atoms
757 < or molecules move in and out of the cutoff
758 < radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
759 < some degree by using group based cutoffs with a switching
760 < function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
761 < significant improvement using the group-switched cutoff because the
762 < salt and salt solution systems contain non-neutral groups.  Section
763 < \ref{sec:IndividualResults} includes results for systems comprised entirely
764 < of neutral groups.
765 <
766 < For the {\sc sp} method, inclusion of electrostatic damping improves
767 < the agreement with Ewald, and using an $\alpha$ of 0.2\AA $^{-1}$
768 < shows an excellent correlation and quality of fit with the {\sc spme}
769 < results, particularly with a cutoff radius greater than 12
770 < \AA .  Use of a larger damping parameter is more helpful for the
771 < shortest cutoff shown, but it has a detrimental effect on simulations
772 < with larger cutoffs.  
773 <
774 < In the {\sc sf} sets, increasing damping results in progressively {\it
775 < worse} correlation with Ewald.  Overall, the undamped case is the best
776 < performing set, as the correlation and quality of fits are
777 < consistently superior regardless of the cutoff distance.  The undamped
778 < case is also less computationally demanding (because no evaluation of
779 < the complementary error function is required).
780 <
781 < The reaction field results illustrates some of that method's
782 < limitations, primarily that it was developed for use in homogeneous
783 < systems; although it does provide results that are an improvement over
784 < those from an unmodified cutoff.
785 <
786 < \section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults}
787 <
788 < Evaluation of pairwise methods for use in Molecular Dynamics
789 < simulations requires consideration of effects on the forces and
790 < torques.  Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
791 < regression results for the force and torque vector magnitudes,
792 < respectively.  The data in these figures was generated from an
793 < accumulation of the statistics from all of the system types.
794 <
795 < \begin{figure}
796 < \centering
797 < \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
798 < \caption{Statistical analysis of the quality of the force vector
799 < magnitudes for a given electrostatic method compared with the
800 < reference Ewald sum.  Results with a value equal to 1 (dashed line)
801 < indicate force magnitude values indistinguishable from those obtained
802 < using {\sc spme}.  Different values of the cutoff radius are indicated with
803 < different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
804 < inverted triangles).}
805 < \label{fig:frcMag}
806 < \end{figure}
807 <
808 < Again, it is striking how well the Shifted Potential and Shifted Force
809 < methods are doing at reproducing the {\sc spme} forces.  The undamped and
810 < weakly-damped {\sc sf} method gives the best agreement with Ewald.
811 < This is perhaps expected because this method explicitly incorporates a
812 < smooth transition in the forces at the cutoff radius as well as the
813 < neutralizing image charges.
814 <
815 < Figure \ref{fig:frcMag}, for the most part, parallels the results seen
816 < in the previous $\Delta E$ section.  The unmodified cutoff results are
817 < poor, but using group based cutoffs and a switching function provides
818 < an improvement much more significant than what was seen with $\Delta
819 < E$.
820 <
821 < With moderate damping and a large enough cutoff radius, the {\sc sp}
822 < method is generating usable forces.  Further increases in damping,
823 < while beneficial for simulations with a cutoff radius of 9\AA\ , is
824 < detrimental to simulations with larger cutoff radii.
825 <
826 < The reaction field results are surprisingly good, considering the poor
827 < quality of the fits for the $\Delta E$ results.  There is still a
828 < considerable degree of scatter in the data, but the forces correlate
829 < well with the Ewald forces in general.  We note that the reaction
830 < field calculations do not include the pure NaCl systems, so these
831 < results are partly biased towards conditions in which the method
832 < performs more favorably.
833 <
834 < \begin{figure}
835 < \centering
836 < \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
837 < \caption{Statistical analysis of the quality of the torque vector
838 < magnitudes for a given electrostatic method compared with the
839 < reference Ewald sum.  Results with a value equal to 1 (dashed line)
840 < indicate torque magnitude values indistinguishable from those obtained
841 < using {\sc spme}.  Different values of the cutoff radius are indicated with
842 < different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
843 < inverted triangles).}
844 < \label{fig:trqMag}
845 < \end{figure}
846 <
847 < Molecular torques were only available from the systems which contained
848 < rigid molecules (i.e. the systems containing water).  The data in
849 < fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
850 <
851 < Torques appear to be much more sensitive to charges at a longer
852 < distance.   The striking feature in comparing the new electrostatic
853 < methods with {\sc spme} is how much the agreement improves with increasing
854 < cutoff radius.  Again, the weakly damped and undamped {\sc sf} method
855 < appears to be reproducing the {\sc spme} torques most accurately.  
856 <
857 < Water molecules are dipolar, and the reaction field method reproduces
858 < the effect of the surrounding polarized medium on each of the
859 < molecular bodies. Therefore it is not surprising that reaction field
860 < performs best of all of the methods on molecular torques.
861 <
862 < \section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults}
863 <
864 < It is clearly important that a new electrostatic method can reproduce
865 < the magnitudes of the force and torque vectors obtained via the Ewald
866 < sum. However, the {\it directionality} of these vectors will also be
867 < vital in calculating dynamical quantities accurately.  Force and
868 < torque directionalities were investigated by measuring the angles
869 < formed between these vectors and the same vectors calculated using
870 < {\sc spme}.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the
871 < variance ($\sigma^2$) of the Gaussian fits of the angle error
872 < distributions of the combined set over all system types.
873 <
874 < \begin{figure}
875 < \centering
876 < \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
877 < \caption{Statistical analysis of the width of the angular distribution
878 < that the force and torque vectors from a given electrostatic method
879 < make with their counterparts obtained using the reference Ewald sum.
880 < Results with a variance ($\sigma^2$) equal to zero (dashed line)
881 < indicate force and torque directions indistinguishable from those
882 < obtained using {\sc spme}.  Different values of the cutoff radius are
883 < indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
884 < and 15\AA\ = inverted triangles).}
885 < \label{fig:frcTrqAng}
886 < \end{figure}
887 <
888 < Both the force and torque $\sigma^2$ results from the analysis of the
889 < total accumulated system data are tabulated in figure
890 < \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
891 < sp}) method would be essentially unusable for molecular dynamics
892 < unless the damping function is added.  The Shifted Force ({\sc sf})
893 < method, however, is generating force and torque vectors which are
894 < within a few degrees of the Ewald results even with weak (or no)
895 < damping.
896 <
897 < All of the sets (aside from the over-damped case) show the improvement
898 < afforded by choosing a larger cutoff radius.  Increasing the cutoff
899 < from 9 to 12\AA\ typically results in a halving of the width of the
900 < distribution, with a similar improvement when going from 12 to 15
901 < \AA .
902 <
903 < The undamped {\sc sf}, group-based cutoff, and reaction field methods
904 < all do equivalently well at capturing the direction of both the force
905 < and torque vectors.  Using the electrostatic damping improves the
906 < angular behavior significantly for the {\sc sp} and moderately for the
907 < {\sc sf} methods.  Over-damping is detrimental to both methods.  Again
908 < it is important to recognize that the force vectors cover all
909 < particles in all seven systems, while torque vectors are only
910 < available for neutral molecular groups.  Damping is more beneficial to
911 < charged bodies, and this observation is investigated further in
912 < section \ref{sec:IndividualResults}.
913 <
914 < Although not discussed previously, group based cutoffs can be applied
915 < to both the {\sc sp} and {\sc sf} methods.  The group-based cutoffs
916 < will reintroduce small discontinuities at the cutoff radius, but the
917 < effects of these can be minimized by utilizing a switching function.
918 < Though there are no significant benefits or drawbacks observed in
919 < $\Delta E$ and the force and torque magnitudes when doing this, there
920 < is a measurable improvement in the directionality of the forces and
921 < torques. Table \ref{tab:groupAngle} shows the angular variances
922 < obtained both without (N) and with (Y) group based cutoffs and a
923 < switching function.  Note that the $\alpha$ values have units of
924 < \AA$^{-1}$ and the variance values have units of degrees$^2$.  The
925 < {\sc sp} (with an $\alpha$ of 0.2\AA$^{-1}$ or smaller) shows much
926 < narrower angular distributions when using group-based cutoffs. The
927 < {\sc sf} method likewise shows improvement in the undamped and lightly
928 < damped cases.
929 <
930 < \begin{table}
931 < \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
932 < THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
933 < GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
934 < THE REFERENCE EWALD SUMMATION}
935 <
936 < \footnotesize
937 < \begin{center}
938 < \begin{tabular}{@{} ccrrrrrrrr @{}}
939 < \toprule
940 < \toprule
941 < & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
942 < Force} \\
943 < \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
944 < $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
945 < $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
946 <
947 < \midrule
948 <    
949 < 9\AA   & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
950 <       & \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} \\
951 < 12\AA  & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
952 <       & \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} \\
953 < 15\AA  & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
954 <       & \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} \\
955 <
956 < \midrule
957 <      
958 < 9\AA   & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
959 <       & \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} \\
960 < 12\AA  & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
961 <       & \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} \\
962 < 15\AA  & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
963 <       & \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} \\
964 <
965 < \bottomrule
966 < \end{tabular}
967 < \end{center}
968 < \label{tab:groupAngle}
969 < \end{table}
970 <
971 < One additional trend in table \ref{tab:groupAngle} is that the
972 < $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
973 < increases, something that is more obvious with group-based cutoffs.
974 < The complimentary error function inserted into the potential weakens
975 < the electrostatic interaction as the value of $\alpha$ is increased.
976 < However, at larger values of $\alpha$, it is possible to over-damp the
977 < electrostatic interaction and to remove it completely.  Kast
978 < \textit{et al.}  developed a method for choosing appropriate $\alpha$
979 < values for these types of electrostatic summation methods by fitting
980 < to $g(r)$ data, and their methods indicate optimal values of 0.34,
981 < 0.25, and 0.16\AA$^{-1}$ for cutoff values of 9, 12, and 15\AA\
982 < respectively.\cite{Kast03} These appear to be reasonable choices to
983 < obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
984 < these findings, choices this high would introduce error in the
985 < molecular torques, particularly for the shorter cutoffs.  Based on our
986 < observations, empirical damping up to 0.2\AA$^{-1}$ is beneficial,
987 < but damping may be unnecessary when using the {\sc sf} method.
988 <
989 < \section{Individual System Analysis Results}\label{sec:IndividualResults}
990 <
991 < The combined results of the previous sections show how the pairwise
992 < methods compare to the Ewald summation in the general sense over all
993 < of the system types.  It is also useful to consider each of the
994 < studied systems in an individual fashion, so that we can identify
995 < conditions that are particularly difficult for a selected pairwise
996 < method to address. This allows us to further establish the limitations
997 < of these pairwise techniques.  Below, the energy difference, force
998 < vector, and torque vector analyses are presented on an individual
999 < system basis.
1000 <
1001 < \subsection{SPC/E Water Results}\label{sec:WaterResults}
1002 <
1003 < The first system considered was liquid water at 300K using the SPC/E
1004 < model of water.\cite{Berendsen87} The results for the energy gap
1005 < comparisons and the force and torque vector magnitude comparisons are
1006 < shown in table \ref{tab:spce}.  The force and torque vector
1007 < directionality results are displayed separately in table
1008 < \ref{tab:spceAng}, where the effect of group-based cutoffs and
1009 < switching functions on the {\sc sp} and {\sc sf} potentials are also
1010 < investigated.  In all of the individual results table, the method
1011 < abbreviations are as follows:
1012 <
1013 < \begin{itemize}[itemsep=0pt]
1014 < \item PC = Pure Cutoff,
1015 < \item SP = Shifted Potential,
1016 < \item SF = Shifted Force,
1017 < \item GSC = Group Switched Cutoff,
1018 < \item RF = Reaction Field (where $\varepsilon \approx\infty$),
1019 < \item GSSP = Group Switched Shifted Potential, and
1020 < \item GSSF = Group Switched Shifted Force.
1021 < \end{itemize}
1022 <
1023 < \begin{table}[htbp]
1024 < \centering
1025 < \caption{REGRESSION RESULTS OF THE LIQUID WATER SYSTEM FOR THE
1026 < $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it middle})
1027 < AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1028 <
1029 < \footnotesize
1030 < \begin{tabular}{@{} ccrrrrrr @{}}
1031 < \toprule
1032 < \toprule
1033 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1034 < \cmidrule(lr){3-4}
1035 < \cmidrule(lr){5-6}
1036 < \cmidrule(l){7-8}
1037 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1038 < \midrule
1039 < PC  &     & 3.046 & 0.002 & -3.018 & 0.002 & 4.719 & 0.005 \\
1040 < SP  & 0.0 & 1.035 & 0.218 & 0.908 & 0.313 & 1.037 & 0.470 \\
1041 <    & 0.1 & 1.021 & 0.387 & 0.965 & 0.752 & 1.006 & 0.947 \\
1042 <    & 0.2 & 0.997 & 0.962 & 1.001 & 0.994 & 0.994 & 0.996 \\
1043 <    & 0.3 & 0.984 & 0.980 & 0.997 & 0.985 & 0.982 & 0.987 \\
1044 < SF  & 0.0 & 0.977 & 0.974 & 0.996 & 0.992 & 0.991 & 0.997 \\
1045 <    & 0.1 & 0.983 & 0.974 & 1.001 & 0.994 & 0.996 & 0.998 \\
1046 <    & 0.2 & 0.992 & 0.989 & 1.001 & 0.995 & 0.994 & 0.996 \\
1047 <    & 0.3 & 0.984 & 0.980 & 0.996 & 0.985 & 0.982 & 0.987 \\
1048 < GSC &     & 0.918 & 0.862 & 0.852 & 0.756 & 0.801 & 0.700 \\
1049 < RF  &     & 0.971 & 0.958 & 0.975 & 0.987 & 0.959 & 0.983 \\                
1050 < \midrule
1051 < PC  &     & -1.647 & 0.000 & -0.127 & 0.000 & -0.979 & 0.000 \\
1052 < SP  & 0.0 & 0.735 & 0.368 & 0.813 & 0.537 & 0.865 & 0.659 \\
1053 <    & 0.1 & 0.850 & 0.612 & 0.956 & 0.887 & 0.992 & 0.979 \\
1054 <    & 0.2 & 0.996 & 0.989 & 1.000 & 1.000 & 1.000 & 1.000 \\
1055 <    & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1056 < SF  & 0.0 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 0.999 \\
1057 <    & 0.1 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1058 <    & 0.2 & 0.999 & 0.998 & 1.000 & 1.000 & 1.000 & 1.000 \\
1059 <    & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1060 < GSC &     & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1061 < RF  &     & 0.999 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\          
1062 < \midrule
1063 < PC  &     & 2.387 & 0.000 & 0.183 & 0.000 & 1.282 & 0.000 \\
1064 < SP  & 0.0 & 0.847 & 0.543 & 0.904 & 0.694 & 0.935 & 0.786 \\
1065 <    & 0.1 & 0.922 & 0.749 & 0.980 & 0.934 & 0.996 & 0.988 \\
1066 <    & 0.2 & 0.987 & 0.985 & 0.989 & 0.992 & 0.990 & 0.993 \\
1067 <    & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1068 < SF  & 0.0 & 0.978 & 0.990 & 0.988 & 0.997 & 0.993 & 0.999 \\
1069 <    & 0.1 & 0.983 & 0.991 & 0.993 & 0.997 & 0.997 & 0.999 \\
1070 <    & 0.2 & 0.986 & 0.989 & 0.989 & 0.992 & 0.990 & 0.993 \\
1071 <    & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1072 < GSC &     & 0.995 & 0.981 & 0.999 & 0.991 & 1.001 & 0.994 \\
1073 < RF  &     & 0.993 & 0.989 & 0.998 & 0.996 & 1.000 & 0.999 \\
1074 < \bottomrule
1075 < \end{tabular}
1076 < \label{tab:spce}
1077 < \end{table}
1078 <
1079 < \begin{table}[htbp]
1080 < \centering
1081 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1082 < DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE LIQUID WATER
1083 < SYSTEM}
1084 <
1085 < \footnotesize
1086 < \begin{tabular}{@{} ccrrrrrr @{}}
1087 < \toprule
1088 < \toprule
1089 < & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1090 < \cmidrule(lr){3-5}
1091 < \cmidrule(l){6-8}
1092 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1093 < \midrule
1094 < PC  &     & 783.759 & 481.353 & 332.677 & 248.674 & 144.382 & 98.535 \\
1095 < SP  & 0.0 & 659.440 & 380.699 & 250.002 & 235.151 & 134.661 & 88.135 \\
1096 <    & 0.1 & 293.849 & 67.772 & 11.609 & 105.090 & 23.813 & 4.369 \\
1097 <    & 0.2 & 5.975 & 0.136 & 0.094 & 5.553 & 1.784 & 1.536 \\
1098 <    & 0.3 & 0.725 & 0.707 & 0.693 & 7.293 & 6.933 & 6.748 \\
1099 < SF  & 0.0 & 2.238 & 0.713 & 0.292 & 3.290 & 1.090 & 0.416 \\
1100 <    & 0.1 & 2.238 & 0.524 & 0.115 & 3.184 & 0.945 & 0.326 \\
1101 <    & 0.2 & 0.374 & 0.102 & 0.094 & 2.598 & 1.755 & 1.537 \\
1102 <    & 0.3 & 0.721 & 0.707 & 0.693 & 7.322 & 6.933 & 6.748 \\
1103 < GSC &     & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1104 < RF  &     & 2.091 & 0.403 & 0.113 & 3.583 & 1.071 & 0.399 \\      
1105 < \midrule
1106 < GSSP  & 0.0 & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1107 <      & 0.1 & 1.879 & 0.291 & 0.057 & 3.983 & 1.117 & 0.370 \\
1108 <      & 0.2 & 0.443 & 0.103 & 0.093 & 2.821 & 1.794 & 1.532 \\
1109 <      & 0.3 & 0.728 & 0.694 & 0.692 & 7.387 & 6.942 & 6.748 \\
1110 < GSSF  & 0.0 & 1.298 & 0.270 & 0.083 & 3.098 & 0.992 & 0.375 \\
1111 <      & 0.1 & 1.296 & 0.210 & 0.044 & 3.055 & 0.922 & 0.330 \\
1112 <      & 0.2 & 0.433 & 0.104 & 0.093 & 2.895 & 1.797 & 1.532 \\
1113 <      & 0.3 & 0.728 & 0.694 & 0.692 & 7.410 & 6.942 & 6.748 \\
1114 < \bottomrule
1115 < \end{tabular}
1116 < \label{tab:spceAng}
1117 < \end{table}
1118 <
1119 < The water results parallel the combined results seen in sections
1120 < \ref{sec:EnergyResults} through \ref{sec:FTDirResults}.  There is good
1121 < agreement with {\sc spme} in both energetic and dynamic behavior when
1122 < using the {\sc sf} method with and without damping. The {\sc sp}
1123 < method does well with an $\alpha$ around 0.2\AA$^{-1}$, particularly
1124 < with cutoff radii greater than 12\AA. Over-damping the electrostatics
1125 < reduces the agreement between both these methods and {\sc spme}.
1126 <
1127 < The pure cutoff ({\sc pc}) method performs poorly, again mirroring the
1128 < observations from the combined results.  In contrast to these results, however, the use of a switching function and group
1129 < based cutoffs greatly improves the results for these neutral water
1130 < molecules.  The group switched cutoff ({\sc gsc}) does not mimic the
1131 < energetics of {\sc spme} as well as the {\sc sp} (with moderate
1132 < damping) and {\sc sf} methods, but the dynamics are quite good.  The
1133 < switching functions correct discontinuities in the potential and
1134 < forces, leading to these improved results.  Such improvements with the
1135 < use of a switching function have been recognized in previous
1136 < studies,\cite{Andrea83,Steinbach94} and this proves to be a useful
1137 < tactic for stably incorporating local area electrostatic effects.
1138 <
1139 < The reaction field ({\sc rf}) method simply extends upon the results
1140 < observed in the {\sc gsc} case.  Both methods are similar in form
1141 < (i.e. neutral groups, switching function), but {\sc rf} incorporates
1142 < an added effect from the external dielectric. This similarity
1143 < translates into the same good dynamic results and improved energetic
1144 < agreement with {\sc spme}.  Though this agreement is not to the level
1145 < of the moderately damped {\sc sp} and {\sc sf} methods, these results
1146 < show how incorporating some implicit properties of the surroundings
1147 < (i.e. $\epsilon_\textrm{S}$) can improve the solvent depiction.
1148 <
1149 < As a final note for the liquid water system, use of group cutoffs and a
1150 < switching function leads to noticeable improvements in the {\sc sp}
1151 < and {\sc sf} methods, primarily in directionality of the force and
1152 < torque vectors (table \ref{tab:spceAng}). The {\sc sp} method shows
1153 < significant narrowing of the angle distribution when using little to
1154 < no damping and only modest improvement for the recommended conditions
1155 < ($\alpha = 0.2$\AA$^{-1}$ and $R_\textrm{c}~\geqslant~12$\AA).  The
1156 < {\sc sf} method shows modest narrowing across all damping and cutoff
1157 < ranges of interest.  When over-damping these methods, group cutoffs and
1158 < the switching function do not improve the force and torque
1159 < directionalities.
1160 <
1161 < \subsection{SPC/E Ice I$_\textrm{c}$ Results}\label{sec:IceResults}
1162 <
1163 < In addition to the disordered molecular system above, the ordered
1164 < molecular system of ice I$_\textrm{c}$ was also considered.  Ice
1165 < polymorph could have been used to fit this role; however, ice
1166 < I$_\textrm{c}$ was chosen because it can form an ideal periodic
1167 < lattice with the same number of water molecules used in the disordered
1168 < liquid state case.  The results for the energy gap comparisons and the
1169 < force and torque vector magnitude comparisons are shown in table
1170 < \ref{tab:ice}.  The force and torque vector directionality results are
1171 < displayed separately in table \ref{tab:iceAng}, where the effect of
1172 < group-based cutoffs and switching functions on the {\sc sp} and {\sc
1173 < sf} potentials are also displayed.
1174 <
1175 < \begin{table}[htbp]
1176 < \centering
1177 < \caption{REGRESSION RESULTS OF THE ICE I$_\textrm{c}$ SYSTEM FOR
1178 < $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it
1179 < middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1180 <
1181 < \footnotesize
1182 < \begin{tabular}{@{} ccrrrrrr @{}}
1183 < \toprule
1184 < \toprule
1185 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1186 < \cmidrule(lr){3-4}
1187 < \cmidrule(lr){5-6}
1188 < \cmidrule(l){7-8}
1189 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1190 < \midrule
1191 < PC  &     & 19.897 & 0.047 & -29.214 & 0.048 & -3.771 & 0.001 \\
1192 < SP  & 0.0 & -0.014 & 0.000 & 2.135 & 0.347 & 0.457 & 0.045 \\
1193 <    & 0.1 & 0.321 & 0.017 & 1.490 & 0.584 & 0.886 & 0.796 \\
1194 <    & 0.2 & 0.896 & 0.872 & 1.011 & 0.998 & 0.997 & 0.999 \\
1195 <    & 0.3 & 0.983 & 0.997 & 0.992 & 0.997 & 0.991 & 0.997 \\
1196 < SF  & 0.0 & 0.943 & 0.979 & 1.048 & 0.978 & 0.995 & 0.999 \\
1197 <    & 0.1 & 0.948 & 0.979 & 1.044 & 0.983 & 1.000 & 0.999 \\
1198 <    & 0.2 & 0.982 & 0.997 & 0.969 & 0.960 & 0.997 & 0.999 \\
1199 <    & 0.3 & 0.985 & 0.997 & 0.961 & 0.961 & 0.991 & 0.997 \\
1200 < GSC &     & 0.983 & 0.985 & 0.966 & 0.994 & 1.003 & 0.999 \\
1201 < RF  &     & 0.924 & 0.944 & 0.990 & 0.996 & 0.991 & 0.998 \\
1202 < \midrule
1203 < PC  &     & -4.375 & 0.000 & 6.781 & 0.000 & -3.369 & 0.000 \\
1204 < SP  & 0.0 & 0.515 & 0.164 & 0.856 & 0.426 & 0.743 & 0.478 \\
1205 <    & 0.1 & 0.696 & 0.405 & 0.977 & 0.817 & 0.974 & 0.964 \\
1206 <    & 0.2 & 0.981 & 0.980 & 1.001 & 1.000 & 1.000 & 1.000 \\
1207 <    & 0.3 & 0.996 & 0.998 & 0.997 & 0.999 & 0.997 & 0.999 \\
1208 < SF  & 0.0 & 0.991 & 0.995 & 1.003 & 0.998 & 0.999 & 1.000 \\
1209 <    & 0.1 & 0.992 & 0.995 & 1.003 & 0.998 & 1.000 & 1.000 \\
1210 <    & 0.2 & 0.998 & 0.998 & 0.981 & 0.962 & 1.000 & 1.000 \\
1211 <    & 0.3 & 0.996 & 0.998 & 0.976 & 0.957 & 0.997 & 0.999 \\
1212 < GSC &     & 0.997 & 0.996 & 0.998 & 0.999 & 1.000 & 1.000 \\
1213 < RF  &     & 0.988 & 0.989 & 1.000 & 0.999 & 1.000 & 1.000 \\
1214 < \midrule
1215 < PC  &     & -6.367 & 0.000 & -3.552 & 0.000 & -3.447 & 0.000 \\
1216 < SP  & 0.0 & 0.643 & 0.409 & 0.833 & 0.607 & 0.961 & 0.805 \\
1217 <    & 0.1 & 0.791 & 0.683 & 0.957 & 0.914 & 1.000 & 0.989 \\
1218 <    & 0.2 & 0.974 & 0.991 & 0.993 & 0.998 & 0.993 & 0.998 \\
1219 <    & 0.3 & 0.976 & 0.992 & 0.977 & 0.992 & 0.977 & 0.992 \\
1220 < SF  & 0.0 & 0.979 & 0.997 & 0.992 & 0.999 & 0.994 & 1.000 \\
1221 <    & 0.1 & 0.984 & 0.997 & 0.996 & 0.999 & 0.998 & 1.000 \\
1222 <    & 0.2 & 0.991 & 0.997 & 0.974 & 0.958 & 0.993 & 0.998 \\
1223 <    & 0.3 & 0.977 & 0.992 & 0.956 & 0.948 & 0.977 & 0.992 \\
1224 < GSC &     & 0.999 & 0.997 & 0.996 & 0.999 & 1.002 & 1.000 \\
1225 < RF  &     & 0.994 & 0.997 & 0.997 & 0.999 & 1.000 & 1.000 \\
1226 < \bottomrule
1227 < \end{tabular}
1228 < \label{tab:ice}
1229 < \end{table}
1230 <
1231 < \begin{table}[htbp]
1232 < \centering
1233 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1234 < OF THE FORCE AND TORQUE VECTORS IN THE ICE I$_\textrm{c}$ SYSTEM}      
1235 <
1236 < \footnotesize
1237 < \begin{tabular}{@{} ccrrrrrr @{}}
1238 < \toprule
1239 < \toprule
1240 < & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque
1241 < $\sigma^2$} \\
1242 < \cmidrule(lr){3-5}
1243 < \cmidrule(l){6-8}
1244 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1245 < \midrule
1246 < PC  &     & 2128.921 & 603.197 & 715.579 & 329.056 & 221.397 & 81.042 \\
1247 < SP  & 0.0 & 1429.341 & 470.320 & 447.557 & 301.678 & 197.437 & 73.840 \\
1248 <    & 0.1 & 590.008 & 107.510 & 18.883 & 118.201 & 32.472 & 3.599 \\
1249 <    & 0.2 & 10.057 & 0.105 & 0.038 & 2.875 & 0.572 & 0.518 \\
1250 <    & 0.3 & 0.245 & 0.260 & 0.262 & 2.365 & 2.396 & 2.327 \\
1251 < SF  & 0.0 & 1.745 & 1.161 & 0.212 & 1.135 & 0.426 & 0.155 \\
1252 <    & 0.1 & 1.721 & 0.868 & 0.082 & 1.118 & 0.358 & 0.118 \\
1253 <    & 0.2 & 0.201 & 0.040 & 0.038 & 0.786 & 0.555 & 0.518 \\
1254 <    & 0.3 & 0.241 & 0.260 & 0.262 & 2.368 & 2.400 & 2.327 \\
1255 < GSC &     & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1256 < RF  &     & 2.887 & 0.217 & 0.107 & 1.006 & 0.281 & 0.085 \\
1257 < \midrule
1258 < GSSP  & 0.0 & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1259 <      & 0.1 & 1.341 & 0.123 & 0.037 & 0.835 & 0.234 & 0.085 \\
1260 <      & 0.2 & 0.558 & 0.040 & 0.037 & 0.823 & 0.557 & 0.519 \\
1261 <      & 0.3 & 0.250 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1262 < GSSF  & 0.0 & 2.124 & 0.132 & 0.069 & 0.919 & 0.263 & 0.099 \\
1263 <      & 0.1 & 2.165 & 0.101 & 0.035 & 0.895 & 0.244 & 0.096 \\
1264 <      & 0.2 & 0.706 & 0.040 & 0.037 & 0.870 & 0.559 & 0.519 \\
1265 <      & 0.3 & 0.251 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1266 < \bottomrule
1267 < \end{tabular}
1268 < \label{tab:iceAng}
1269 < \end{table}
1270 <
1271 < Highly ordered systems are a difficult test for the pairwise methods
1272 < in that they lack the implicit periodicity of the Ewald summation.  As
1273 < expected, the energy gap agreement with {\sc spme} is reduced for the
1274 < {\sc sp} and {\sc sf} methods with parameters that were ideal for the
1275 < disordered liquid system.  Moving to higher $R_\textrm{c}$ helps
1276 < improve the agreement, though at an increase in computational cost.
1277 < The dynamics of this crystalline system (both in magnitude and
1278 < direction) are little affected. Both methods still reproduce the Ewald
1279 < behavior with the same parameter recommendations from the previous
1280 < section.
1281 <
1282 < It is also worth noting that {\sc rf} exhibits improved energy gap
1283 < results over the liquid water system.  One possible explanation is
1284 < that the ice I$_\textrm{c}$ crystal is ordered such that the net
1285 < dipole moment of the crystal is zero.  With $\epsilon_\textrm{S} =
1286 < \infty$, the reaction field incorporates this structural organization
1287 < by actively enforcing a zeroed dipole moment within each cutoff
1288 < sphere.
1289 <
1290 < \subsection{NaCl Melt Results}\label{sec:SaltMeltResults}
1291 <
1292 < A high temperature NaCl melt was tested to gauge the accuracy of the
1293 < pairwise summation methods in a disordered system of charges. The
1294 < results for the energy gap comparisons and the force vector magnitude
1295 < comparisons are shown in table \ref{tab:melt}.  The force vector
1296 < directionality results are displayed separately in table
1297 < \ref{tab:meltAng}.
1298 <
1299 < \begin{table}[htbp]
1300 < \centering
1301 < \caption{REGRESSION RESULTS OF THE MOLTEN SODIUM CHLORIDE SYSTEM FOR
1302 < $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES ({\it
1303 < lower})}
1304 <
1305 < \footnotesize
1306 < \begin{tabular}{@{} ccrrrrrr @{}}
1307 < \toprule
1308 < \toprule
1309 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1310 < \cmidrule(lr){3-4}
1311 < \cmidrule(lr){5-6}
1312 < \cmidrule(l){7-8}
1313 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1314 < \midrule
1315 < PC  &     & -0.008 & 0.000 & -0.049 & 0.005 & -0.136 & 0.020 \\
1316 < SP  & 0.0 & 0.928 & 0.996 & 0.931 & 0.998 & 0.950 & 0.999 \\
1317 <    & 0.1 & 0.977 & 0.998 & 0.998 & 1.000 & 0.997 & 1.000 \\
1318 <    & 0.2 & 0.960 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1319 <    & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1320 < SF  & 0.0 & 0.996 & 1.000 & 0.995 & 1.000 & 0.997 & 1.000 \\
1321 <    & 0.1 & 1.021 & 1.000 & 1.024 & 1.000 & 1.007 & 1.000 \\
1322 <    & 0.2 & 0.966 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1323 <    & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1324 <            \midrule
1325 < PC  &     & 1.103 & 0.000 & 0.989 & 0.000 & 0.802 & 0.000 \\
1326 < SP  & 0.0 & 0.973 & 0.981 & 0.975 & 0.988 & 0.979 & 0.992 \\
1327 <    & 0.1 & 0.987 & 0.992 & 0.993 & 0.998 & 0.997 & 0.999 \\
1328 <    & 0.2 & 0.993 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1329 <    & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1330 < SF  & 0.0 & 0.996 & 0.997 & 0.997 & 0.999 & 0.998 & 1.000 \\
1331 <    & 0.1 & 1.000 & 0.997 & 1.001 & 0.999 & 1.000 & 1.000 \\
1332 <    & 0.2 & 0.994 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1333 <    & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1334 < \bottomrule
1335 < \end{tabular}
1336 < \label{tab:melt}
1337 < \end{table}
1338 <
1339 < \begin{table}[htbp]
1340 < \centering
1341 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1342 < OF THE FORCE VECTORS IN THE MOLTEN SODIUM CHLORIDE SYSTEM}      
1343 <
1344 < \footnotesize
1345 < \begin{tabular}{@{} ccrrrrrr @{}}
1346 < \toprule
1347 < \toprule
1348 < & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1349 < \cmidrule(lr){3-5}
1350 < \cmidrule(l){6-8}
1351 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1352 < \midrule
1353 < PC  &     & 13.294 & 8.035 & 5.366 \\
1354 < SP  & 0.0 & 13.316 & 8.037 & 5.385 \\
1355 <    & 0.1 & 5.705 & 1.391 & 0.360 \\
1356 <    & 0.2 & 2.415 & 7.534 & 13.927 \\
1357 <    & 0.3 & 23.769 & 67.306 & 57.252 \\
1358 < SF  & 0.0 & 1.693 & 0.603 & 0.256 \\
1359 <    & 0.1 & 1.687 & 0.653 & 0.272 \\
1360 <    & 0.2 & 2.598 & 7.523 & 13.930 \\
1361 <    & 0.3 & 23.734 & 67.305 & 57.252 \\
1362 < \bottomrule
1363 < \end{tabular}
1364 < \label{tab:meltAng}
1365 < \end{table}
1366 <
1367 < The molten NaCl system shows more sensitivity to the electrostatic
1368 < damping than the water systems. The most noticeable point is that the
1369 < undamped {\sc sf} method does very well at replicating the {\sc spme}
1370 < configurational energy differences and forces. Light damping appears
1371 < to minimally improve the dynamics, but this comes with a deterioration
1372 < of the energy gap results. In contrast, this light damping improves
1373 < the {\sc sp} energy gaps and forces. Moderate and heavy electrostatic
1374 < damping reduce the agreement with {\sc spme} for both methods. From
1375 < these observations, the undamped {\sc sf} method is the best choice
1376 < for disordered systems of charges.
1377 <
1378 < \subsection{NaCl Crystal Results}\label{sec:SaltCrystalResults}
1379 <
1380 < Similar to the use of ice I$_\textrm{c}$ to investigate the role of
1381 < order in molecular systems on the effectiveness of the pairwise
1382 < methods, the 1000K NaCl crystal system was used to investigate the
1383 < accuracy of the pairwise summation methods in an ordered system of
1384 < charged particles. The results for the energy gap comparisons and the
1385 < force vector magnitude comparisons are shown in table \ref{tab:salt}.
1386 < The force vector directionality results are displayed separately in
1387 < table \ref{tab:saltAng}.
1388 <
1389 < \begin{table}[htbp]
1390 < \centering
1391 < \caption{REGRESSION RESULTS OF THE CRYSTALLINE SODIUM CHLORIDE
1392 < SYSTEM FOR $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES
1393 < ({\it lower})}
1394 <
1395 < \footnotesize
1396 < \begin{tabular}{@{} ccrrrrrr @{}}
1397 < \toprule
1398 < \toprule
1399 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1400 < \cmidrule(lr){3-4}
1401 < \cmidrule(lr){5-6}
1402 < \cmidrule(l){7-8}
1403 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1404 < \midrule
1405 < PC  &     & -20.241 & 0.228 & -20.248 & 0.229 & -20.239 & 0.228 \\
1406 < SP  & 0.0 & 1.039 & 0.733 & 2.037 & 0.565 & 1.225 & 0.743 \\
1407 <    & 0.1 & 1.049 & 0.865 & 1.424 & 0.784 & 1.029 & 0.980 \\
1408 <    & 0.2 & 0.982 & 0.976 & 0.969 & 0.980 & 0.960 & 0.980 \\
1409 <    & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.945 \\
1410 < SF  & 0.0 & 1.041 & 0.967 & 0.994 & 0.989 & 0.957 & 0.993 \\
1411 <    & 0.1 & 1.050 & 0.968 & 0.996 & 0.991 & 0.972 & 0.995 \\
1412 <    & 0.2 & 0.982 & 0.975 & 0.959 & 0.980 & 0.960 & 0.980 \\
1413 <    & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.944 \\
1414 < \midrule
1415 < PC  &     & 0.795 & 0.000 & 0.792 & 0.000 & 0.793 & 0.000 \\
1416 < SP  & 0.0 & 0.916 & 0.829 & 1.086 & 0.791 & 1.010 & 0.936 \\
1417 <    & 0.1 & 0.958 & 0.917 & 1.049 & 0.943 & 1.001 & 0.995 \\
1418 <    & 0.2 & 0.981 & 0.981 & 0.982 & 0.984 & 0.981 & 0.984 \\
1419 <    & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1420 < SF  & 0.0 & 1.002 & 0.983 & 0.997 & 0.994 & 0.991 & 0.997 \\
1421 <    & 0.1 & 1.003 & 0.984 & 0.996 & 0.995 & 0.993 & 0.997 \\
1422 <    & 0.2 & 0.983 & 0.980 & 0.981 & 0.984 & 0.981 & 0.984 \\
1423 <    & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1424 < \bottomrule
1425 < \end{tabular}
1426 < \label{tab:salt}
1427 < \end{table}
1428 <
1429 < \begin{table}[htbp]
1430 < \centering
1431 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1432 < DISTRIBUTIONS OF THE FORCE VECTORS IN THE CRYSTALLINE SODIUM CHLORIDE
1433 < SYSTEM}
1434 <
1435 < \footnotesize
1436 < \begin{tabular}{@{} ccrrrrrr @{}}
1437 < \toprule
1438 < \toprule
1439 < & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1440 < \cmidrule(lr){3-5}
1441 < \cmidrule(l){6-8}
1442 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1443 < \midrule
1444 < PC  &     & 111.945 & 111.824 & 111.866 \\
1445 < SP  & 0.0 & 112.414 & 152.215 & 38.087 \\
1446 <    & 0.1 & 52.361 & 42.574 & 2.819 \\
1447 <    & 0.2 & 10.847 & 9.709 & 9.686 \\
1448 <    & 0.3 & 31.128 & 31.104 & 31.029 \\
1449 < SF  & 0.0 & 10.025 & 3.555 & 1.648 \\
1450 <    & 0.1 & 9.462 & 3.303 & 1.721 \\
1451 <    & 0.2 & 11.454 & 9.813 & 9.701 \\
1452 <    & 0.3 & 31.120 & 31.105 & 31.029 \\
1453 < \bottomrule
1454 < \end{tabular}
1455 < \label{tab:saltAng}
1456 < \end{table}
1457 <
1458 < The crystalline NaCl system is the most challenging test case for the
1459 < pairwise summation methods, as evidenced by the results in tables
1460 < \ref{tab:salt} and \ref{tab:saltAng}. The undamped and weakly damped
1461 < {\sc sf} methods seem to be the best choices. These methods match well
1462 < with {\sc spme} across the energy gap, force magnitude, and force
1463 < directionality tests.  The {\sc sp} method struggles in all cases,
1464 < with the exception of good dynamics reproduction when using weak
1465 < electrostatic damping with a large cutoff radius.
1466 <
1467 < The moderate electrostatic damping case is not as good as we would
1468 < expect given the long-time dynamics results observed for this system
1469 < (see section \ref{sec:LongTimeDynamics}). Since the data tabulated in
1470 < tables \ref{tab:salt} and \ref{tab:saltAng} are a test of
1471 < instantaneous dynamics, this indicates that good long-time dynamics
1472 < comes in part at the expense of short-time dynamics.
1473 <
1474 < \subsection{0.11M NaCl Solution Results}
1475 <
1476 < In an effort to bridge the charged atomic and neutral molecular
1477 < systems, Na$^+$ and Cl$^-$ ion charge defects were incorporated into
1478 < the liquid water system. This low ionic strength system consists of 4
1479 < ions in the 1000 SPC/E water solvent ($\approx$0.11 M). The results
1480 < for the energy gap comparisons and the force and torque vector
1481 < magnitude comparisons are shown in table \ref{tab:solnWeak}.  The
1482 < force and torque vector directionality results are displayed
1483 < separately in table \ref{tab:solnWeakAng}, where the effect of
1484 < group-based cutoffs and switching functions on the {\sc sp} and {\sc
1485 < sf} potentials are investigated.
1486 <
1487 < \begin{table}[htbp]
1488 < \centering
1489 < \caption{REGRESSION RESULTS OF THE WEAK SODIUM CHLORIDE SOLUTION
1490 < SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1491 < ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1492 <
1493 < \footnotesize
1494 < \begin{tabular}{@{} ccrrrrrr @{}}
1495 < \toprule
1496 < \toprule
1497 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1498 < \cmidrule(lr){3-4}
1499 < \cmidrule(lr){5-6}
1500 < \cmidrule(l){7-8}
1501 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1502 < \midrule
1503 < PC  &     & 0.247 & 0.000 & -1.103 & 0.001 & 5.480 & 0.015 \\
1504 < SP  & 0.0 & 0.935 & 0.388 & 0.984 & 0.541 & 1.010 & 0.685 \\
1505 <    & 0.1 & 0.951 & 0.603 & 0.993 & 0.875 & 1.001 & 0.979 \\
1506 <    & 0.2 & 0.969 & 0.968 & 0.996 & 0.997 & 0.994 & 0.997 \\
1507 <    & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1508 < SF  & 0.0 & 0.963 & 0.971 & 0.989 & 0.996 & 0.991 & 0.998 \\
1509 <    & 0.1 & 0.970 & 0.971 & 0.995 & 0.997 & 0.997 & 0.999 \\
1510 <    & 0.2 & 0.972 & 0.975 & 0.996 & 0.997 & 0.994 & 0.997 \\
1511 <    & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1512 < GSC &     & 0.964 & 0.731 & 0.984 & 0.704 & 1.005 & 0.770 \\
1513 < RF  &     & 0.968 & 0.605 & 0.974 & 0.541 & 1.014 & 0.614 \\
1514 < \midrule
1515 < PC  &     & 1.354 & 0.000 & -1.190 & 0.000 & -0.314 & 0.000 \\
1516 < SP  & 0.0 & 0.720 & 0.338 & 0.808 & 0.523 & 0.860 & 0.643 \\
1517 <    & 0.1 & 0.839 & 0.583 & 0.955 & 0.882 & 0.992 & 0.978 \\
1518 <    & 0.2 & 0.995 & 0.987 & 0.999 & 1.000 & 0.999 & 1.000 \\
1519 <    & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1520 < SF  & 0.0 & 0.998 & 0.994 & 1.000 & 0.998 & 1.000 & 0.999 \\
1521 <    & 0.1 & 0.997 & 0.994 & 1.000 & 0.999 & 1.000 & 1.000 \\
1522 <    & 0.2 & 0.999 & 0.998 & 0.999 & 1.000 & 0.999 & 1.000 \\
1523 <    & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1524 < GSC &     & 0.995 & 0.990 & 0.998 & 0.997 & 0.998 & 0.996 \\
1525 < RF  &     & 0.998 & 0.993 & 0.999 & 0.998 & 0.999 & 0.996 \\
1526 < \midrule
1527 < PC  &     & 2.437 & 0.000 & -1.872 & 0.000 & 2.138 & 0.000 \\
1528 < SP  & 0.0 & 0.838 & 0.525 & 0.901 & 0.686 & 0.932 & 0.779 \\
1529 <    & 0.1 & 0.914 & 0.733 & 0.979 & 0.932 & 0.995 & 0.987 \\
1530 <    & 0.2 & 0.977 & 0.969 & 0.988 & 0.990 & 0.989 & 0.990 \\
1531 <    & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1532 < SF  & 0.0 & 0.969 & 0.977 & 0.987 & 0.996 & 0.993 & 0.998 \\
1533 <    & 0.1 & 0.975 & 0.978 & 0.993 & 0.996 & 0.997 & 0.998 \\
1534 <    & 0.2 & 0.976 & 0.973 & 0.988 & 0.990 & 0.989 & 0.990 \\
1535 <    & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1536 < GSC &     & 0.980 & 0.959 & 0.990 & 0.983 & 0.992 & 0.989 \\
1537 < RF  &     & 0.984 & 0.975 & 0.996 & 0.995 & 0.998 & 0.998 \\
1538 < \bottomrule
1539 < \end{tabular}
1540 < \label{tab:solnWeak}
1541 < \end{table}
1542 <
1543 < \begin{table}[htbp]
1544 < \centering
1545 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1546 < DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE WEAK SODIUM
1547 < CHLORIDE SOLUTION SYSTEM}
1548 <
1549 < \footnotesize
1550 < \begin{tabular}{@{} ccrrrrrr @{}}
1551 < \toprule
1552 < \toprule
1553 < & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1554 < \cmidrule(lr){3-5}
1555 < \cmidrule(l){6-8}
1556 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1557 < \midrule
1558 < PC  &     & 882.863 & 510.435 & 344.201 & 277.691 & 154.231 & 100.131 \\
1559 < SP  & 0.0 & 732.569 & 405.704 & 257.756 & 261.445 & 142.245 & 91.497 \\
1560 <    & 0.1 & 329.031 & 70.746 & 12.014 & 118.496 & 25.218 & 4.711 \\
1561 <    & 0.2 & 6.772 & 0.153 & 0.118 & 9.780 & 2.101 & 2.102 \\
1562 <    & 0.3 & 0.951 & 0.774 & 0.784 & 12.108 & 7.673 & 7.851 \\
1563 < SF  & 0.0 & 2.555 & 0.762 & 0.313 & 6.590 & 1.328 & 0.558 \\
1564 <    & 0.1 & 2.561 & 0.560 & 0.123 & 6.464 & 1.162 & 0.457 \\
1565 <    & 0.2 & 0.501 & 0.118 & 0.118 & 5.698 & 2.074 & 2.099 \\
1566 <    & 0.3 & 0.943 & 0.774 & 0.784 & 12.118 & 7.674 & 7.851 \\
1567 < GSC &     & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1568 < RF  &     & 2.415 & 0.452 & 0.130 & 6.915 & 1.423 & 0.507 \\
1569 < \midrule
1570 < GSSP  & 0.0 & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1571 <      & 0.1 & 2.251 & 0.324 & 0.064 & 7.628 & 1.639 & 0.497 \\
1572 <      & 0.2 & 0.590 & 0.118 & 0.116 & 6.080 & 2.096 & 2.103 \\
1573 <      & 0.3 & 0.953 & 0.759 & 0.780 & 12.347 & 7.683 & 7.849 \\
1574 < GSSF  & 0.0 & 1.541 & 0.301 & 0.096 & 6.407 & 1.316 & 0.496 \\
1575 <      & 0.1 & 1.541 & 0.237 & 0.050 & 6.356 & 1.202 & 0.457 \\
1576 <      & 0.2 & 0.568 & 0.118 & 0.116 & 6.166 & 2.105 & 2.105 \\
1577 <      & 0.3 & 0.954 & 0.759 & 0.780 & 12.337 & 7.684 & 7.849 \\
1578 < \bottomrule
1579 < \end{tabular}
1580 < \label{tab:solnWeakAng}
1581 < \end{table}
1582 <
1583 < Because this system is a perturbation of the pure liquid water system,
1584 < comparisons are best drawn between these two sets. The {\sc sp} and
1585 < {\sc sf} methods are not significantly affected by the inclusion of a
1586 < few ions. The aspect of cutoff sphere neutralization aids in the
1587 < smooth incorporation of these ions; thus, all of the observations
1588 < regarding these methods carry over from section
1589 < \ref{sec:WaterResults}. The differences between these systems are more
1590 < visible for the {\sc rf} method. Though good force agreement is still
1591 < maintained, the energy gaps show a significant increase in the scatter
1592 < of the data.
1593 <
1594 < \subsection{1.1M NaCl Solution Results}
1595 <
1596 < The bridging of the charged atomic and neutral molecular systems was
1597 < further developed by considering a high ionic strength system
1598 < consisting of 40 ions in the 1000 SPC/E water solvent ($\approx$1.1
1599 < M). The results for the energy gap comparisons and the force and
1600 < torque vector magnitude comparisons are shown in table
1601 < \ref{tab:solnStr}.  The force and torque vector directionality
1602 < results are displayed separately in table \ref{tab:solnStrAng}, where
1603 < the effect of group-based cutoffs and switching functions on the {\sc
1604 < sp} and {\sc sf} potentials are investigated.
1605 <
1606 < \begin{table}[htbp]
1607 < \centering
1608 < \caption{REGRESSION RESULTS OF THE STRONG SODIUM CHLORIDE SOLUTION
1609 < SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1610 < ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1611 <
1612 < \footnotesize
1613 < \begin{tabular}{@{} ccrrrrrr @{}}
1614 < \toprule
1615 < \toprule
1616 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1617 < \cmidrule(lr){3-4}
1618 < \cmidrule(lr){5-6}
1619 < \cmidrule(l){7-8}
1620 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1621 < \midrule
1622 < PC  &     & -0.081 & 0.000 & 0.945 & 0.001 & 0.073 & 0.000 \\
1623 < SP  & 0.0 & 0.978 & 0.469 & 0.996 & 0.672 & 0.975 & 0.668 \\
1624 <    & 0.1 & 0.944 & 0.645 & 0.997 & 0.886 & 0.991 & 0.978 \\
1625 <    & 0.2 & 0.873 & 0.896 & 0.985 & 0.993 & 0.980 & 0.993 \\
1626 <    & 0.3 & 0.831 & 0.860 & 0.960 & 0.979 & 0.955 & 0.977 \\
1627 < SF  & 0.0 & 0.858 & 0.905 & 0.985 & 0.970 & 0.990 & 0.998 \\
1628 <    & 0.1 & 0.865 & 0.907 & 0.992 & 0.974 & 0.994 & 0.999 \\
1629 <    & 0.2 & 0.862 & 0.894 & 0.985 & 0.993 & 0.980 & 0.993 \\
1630 <    & 0.3 & 0.831 & 0.859 & 0.960 & 0.979 & 0.955 & 0.977 \\
1631 < GSC &     & 1.985 & 0.152 & 0.760 & 0.031 & 1.106 & 0.062 \\
1632 < RF  &     & 2.414 & 0.116 & 0.813 & 0.017 & 1.434 & 0.047 \\
1633 < \midrule
1634 < PC  &     & -7.028 & 0.000 & -9.364 & 0.000 & 0.925 & 0.865 \\
1635 < SP  & 0.0 & 0.701 & 0.319 & 0.909 & 0.773 & 0.861 & 0.665 \\
1636 <    & 0.1 & 0.824 & 0.565 & 0.970 & 0.930 & 0.990 & 0.979 \\
1637 <    & 0.2 & 0.988 & 0.981 & 0.995 & 0.998 & 0.991 & 0.998 \\
1638 <    & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1639 < SF  & 0.0 & 0.993 & 0.988 & 0.992 & 0.984 & 0.998 & 0.999 \\
1640 <    & 0.1 & 0.993 & 0.989 & 0.993 & 0.986 & 0.998 & 1.000 \\
1641 <    & 0.2 & 0.993 & 0.992 & 0.995 & 0.998 & 0.991 & 0.998 \\
1642 <    & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1643 < GSC &     & 0.964 & 0.897 & 0.970 & 0.917 & 0.925 & 0.865 \\
1644 < RF  &     & 0.994 & 0.864 & 0.988 & 0.865 & 0.980 & 0.784 \\
1645 < \midrule
1646 < PC  &     & -2.212 & 0.000 & -0.588 & 0.000 & 0.953 & 0.925 \\
1647 < SP  & 0.0 & 0.800 & 0.479 & 0.930 & 0.804 & 0.924 & 0.759 \\
1648 <    & 0.1 & 0.883 & 0.694 & 0.976 & 0.942 & 0.993 & 0.986 \\
1649 <    & 0.2 & 0.952 & 0.943 & 0.980 & 0.984 & 0.980 & 0.983 \\
1650 <    & 0.3 & 0.914 & 0.909 & 0.943 & 0.948 & 0.944 & 0.946 \\
1651 < SF  & 0.0 & 0.945 & 0.953 & 0.980 & 0.984 & 0.991 & 0.998 \\
1652 <    & 0.1 & 0.951 & 0.954 & 0.987 & 0.986 & 0.995 & 0.998 \\
1653 <    & 0.2 & 0.951 & 0.946 & 0.980 & 0.984 & 0.980 & 0.983 \\
1654 <    & 0.3 & 0.914 & 0.908 & 0.943 & 0.948 & 0.944 & 0.946 \\
1655 < GSC &     & 0.882 & 0.818 & 0.939 & 0.902 & 0.953 & 0.925 \\
1656 < RF  &     & 0.949 & 0.939 & 0.988 & 0.988 & 0.992 & 0.993 \\
1657 < \bottomrule
1658 < \end{tabular}
1659 < \label{tab:solnStr}
1660 < \end{table}
1661 <
1662 < \begin{table}[htbp]
1663 < \centering
1664 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1665 < OF THE FORCE AND TORQUE VECTORS IN THE STRONG SODIUM CHLORIDE SOLUTION
1666 < SYSTEM}
1667 <
1668 < \footnotesize
1669 < \begin{tabular}{@{} ccrrrrrr @{}}
1670 < \toprule
1671 < \toprule
1672 < & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1673 < \cmidrule(lr){3-5}
1674 < \cmidrule(l){6-8}
1675 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1676 < \midrule
1677 < PC  &     & 957.784 & 513.373 & 2.260 & 340.043 & 179.443 & 13.079 \\
1678 < SP  & 0.0 & 786.244 & 139.985 & 259.289 & 311.519 & 90.280 & 105.187 \\
1679 <    & 0.1 & 354.697 & 38.614 & 12.274 & 144.531 & 23.787 & 5.401 \\
1680 <    & 0.2 & 7.674 & 0.363 & 0.215 & 16.655 & 3.601 & 3.634 \\
1681 <    & 0.3 & 1.745 & 1.456 & 1.449 & 23.669 & 14.376 & 14.240 \\
1682 < SF  & 0.0 & 3.282 & 8.567 & 0.369 & 11.904 & 6.589 & 0.717 \\
1683 <    & 0.1 & 3.263 & 7.479 & 0.142 & 11.634 & 5.750 & 0.591 \\
1684 <    & 0.2 & 0.686 & 0.324 & 0.215 & 10.809 & 3.580 & 3.635 \\
1685 <    & 0.3 & 1.749 & 1.456 & 1.449 & 23.635 & 14.375 & 14.240 \\
1686 < GSC &     & 6.181 & 2.904 & 2.263 & 44.349 & 19.442 & 12.873 \\
1687 < RF  &     & 3.891 & 0.847 & 0.323 & 18.628 & 3.995 & 2.072 \\
1688 < \midrule
1689 < GSSP  & 0.0 & 6.197 & 2.929 & 2.290 & 44.441 & 19.442 & 12.873 \\
1690 <      & 0.1 & 4.688 & 1.064 & 0.260 & 31.208 & 6.967 & 2.303 \\
1691 <      & 0.2 & 1.021 & 0.218 & 0.213 & 14.425 & 3.629 & 3.649 \\
1692 <      & 0.3 & 1.752 & 1.454 & 1.451 & 23.540 & 14.390 & 14.245 \\
1693 < GSSF  & 0.0 & 2.494 & 0.546 & 0.217 & 16.391 & 3.230 & 1.613 \\
1694 <      & 0.1 & 2.448 & 0.429 & 0.106 & 16.390 & 2.827 & 1.159 \\
1695 <      & 0.2 & 0.899 & 0.214 & 0.213 & 13.542 & 3.583 & 3.645 \\
1696 <      & 0.3 & 1.752 & 1.454 & 1.451 & 23.587 & 14.390 & 14.245 \\
1697 < \bottomrule
1698 < \end{tabular}
1699 < \label{tab:solnStrAng}
1700 < \end{table}
1701 <
1702 < The {\sc rf} method struggles with the jump in ionic strength. The
1703 < configuration energy differences degrade to unusable levels while the
1704 < forces and torques show a more modest reduction in the agreement with
1705 < {\sc spme}. The {\sc rf} method was designed for homogeneous systems,
1706 < and this attribute is apparent in these results.
1707 <
1708 < The {\sc sp} and {\sc sf} methods require larger cutoffs to maintain
1709 < their agreement with {\sc spme}. With these results, we still
1710 < recommend undamped to moderate damping for the {\sc sf} method and
1711 < moderate damping for the {\sc sp} method, both with cutoffs greater
1712 < than 12\AA.
1713 <
1714 < \subsection{6\AA\ Argon Sphere in SPC/E Water Results}
1715 <
1716 < The final model system studied was a 6\AA\ sphere of Argon solvated
1717 < by SPC/E water. This serves as a test case of a specifically sized
1718 < electrostatic defect in a disordered molecular system. The results for
1719 < the energy gap comparisons and the force and torque vector magnitude
1720 < comparisons are shown in table \ref{tab:argon}.  The force and torque
1721 < vector directionality results are displayed separately in table
1722 < \ref{tab:argonAng}, where the effect of group-based cutoffs and
1723 < switching functions on the {\sc sp} and {\sc sf} potentials are
1724 < investigated.
1725 <
1726 < \begin{table}[htbp]
1727 < \centering
1728 < \caption{REGRESSION RESULTS OF THE 6\AA\ ARGON SPHERE IN LIQUID
1729 < WATER SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR
1730 < MAGNITUDES ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1731 <
1732 < \footnotesize
1733 < \begin{tabular}{@{} ccrrrrrr @{}}
1734 < \toprule
1735 < \toprule
1736 < & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1737 < \cmidrule(lr){3-4}
1738 < \cmidrule(lr){5-6}
1739 < \cmidrule(l){7-8}
1740 < Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1741 < \midrule
1742 < PC  &     & 2.320 & 0.008 & -0.650 & 0.001 & 3.848 & 0.029 \\
1743 < SP  & 0.0 & 1.053 & 0.711 & 0.977 & 0.820 & 0.974 & 0.882 \\
1744 <    & 0.1 & 1.032 & 0.846 & 0.989 & 0.965 & 0.992 & 0.994 \\
1745 <    & 0.2 & 0.993 & 0.995 & 0.982 & 0.998 & 0.986 & 0.998 \\
1746 <    & 0.3 & 0.968 & 0.995 & 0.954 & 0.992 & 0.961 & 0.994 \\
1747 < SF  & 0.0 & 0.982 & 0.996 & 0.992 & 0.999 & 0.993 & 1.000 \\
1748 <    & 0.1 & 0.987 & 0.996 & 0.996 & 0.999 & 0.997 & 1.000 \\
1749 <    & 0.2 & 0.989 & 0.998 & 0.984 & 0.998 & 0.989 & 0.998 \\
1750 <    & 0.3 & 0.971 & 0.995 & 0.957 & 0.992 & 0.965 & 0.994 \\
1751 < GSC &     & 1.002 & 0.983 & 0.992 & 0.973 & 0.996 & 0.971 \\
1752 < RF  &     & 0.998 & 0.995 & 0.999 & 0.998 & 0.998 & 0.998 \\
1753 < \midrule
1754 < PC  &     & -36.559 & 0.002 & -44.917 & 0.004 & -52.945 & 0.006 \\
1755 < SP  & 0.0 & 0.890 & 0.786 & 0.927 & 0.867 & 0.949 & 0.909 \\
1756 <    & 0.1 & 0.942 & 0.895 & 0.984 & 0.974 & 0.997 & 0.995 \\
1757 <    & 0.2 & 0.999 & 0.997 & 1.000 & 1.000 & 1.000 & 1.000 \\
1758 <    & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1759 < SF  & 0.0 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1760 <    & 0.1 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1761 <    & 0.2 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 \\
1762 <    & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1763 < GSC &     & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1764 < RF  &     & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1765 < \midrule
1766 < PC  &     & 1.984 & 0.000 & 0.012 & 0.000 & 1.357 & 0.000 \\
1767 < SP  & 0.0 & 0.850 & 0.552 & 0.907 & 0.703 & 0.938 & 0.793 \\
1768 <    & 0.1 & 0.924 & 0.755 & 0.980 & 0.936 & 0.995 & 0.988 \\
1769 <    & 0.2 & 0.985 & 0.983 & 0.986 & 0.988 & 0.987 & 0.988 \\
1770 <    & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1771 < SF  & 0.0 & 0.977 & 0.989 & 0.987 & 0.995 & 0.992 & 0.998 \\
1772 <    & 0.1 & 0.982 & 0.989 & 0.992 & 0.996 & 0.997 & 0.998 \\
1773 <    & 0.2 & 0.984 & 0.987 & 0.986 & 0.987 & 0.987 & 0.988 \\
1774 <    & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1775 < GSC &     & 0.995 & 0.981 & 0.999 & 0.990 & 1.000 & 0.993 \\
1776 < RF  &     & 0.993 & 0.988 & 0.997 & 0.995 & 0.999 & 0.998 \\
1777 < \bottomrule
1778 < \end{tabular}
1779 < \label{tab:argon}
1780 < \end{table}
1781 <
1782 < \begin{table}[htbp]
1783 < \centering
1784 < \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1785 < DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE 6\AA\ SPHERE OF
1786 < ARGON IN LIQUID WATER SYSTEM}  
1787 <
1788 < \footnotesize
1789 < \begin{tabular}{@{} ccrrrrrr @{}}
1790 < \toprule
1791 < \toprule
1792 < & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1793 < \cmidrule(lr){3-5}
1794 < \cmidrule(l){6-8}
1795 < Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1796 < \midrule
1797 < PC  &     & 568.025 & 265.993 & 195.099 & 246.626 & 138.600 & 91.654 \\
1798 < SP  & 0.0 & 504.578 & 251.694 & 179.932 & 231.568 & 131.444 & 85.119 \\
1799 <    & 0.1 & 224.886 & 49.746 & 9.346 & 104.482 & 23.683 & 4.480 \\
1800 <    & 0.2 & 4.889 & 0.197 & 0.155 & 6.029 & 2.507 & 2.269 \\
1801 <    & 0.3 & 0.817 & 0.833 & 0.812 & 8.286 & 8.436 & 8.135 \\
1802 < SF  & 0.0 & 1.924 & 0.675 & 0.304 & 3.658 & 1.448 & 0.600 \\
1803 <    & 0.1 & 1.937 & 0.515 & 0.143 & 3.565 & 1.308 & 0.546 \\
1804 <    & 0.2 & 0.407 & 0.166 & 0.156 & 3.086 & 2.501 & 2.274 \\
1805 <    & 0.3 & 0.815 & 0.833 & 0.812 & 8.330 & 8.437 & 8.135 \\
1806 < GSC &     & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1807 < RF  &     & 1.822 & 0.408 & 0.142 & 3.799 & 1.362 & 0.550 \\
1808 < \midrule
1809 < GSSP  & 0.0 & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1810 <      & 0.1 & 1.652 & 0.309 & 0.087 & 4.197 & 1.401 & 0.590 \\
1811 <      & 0.2 & 0.465 & 0.165 & 0.153 & 3.323 & 2.529 & 2.273 \\
1812 <      & 0.3 & 0.813 & 0.825 & 0.816 & 8.316 & 8.447 & 8.132 \\
1813 < GSSF  & 0.0 & 1.173 & 0.292 & 0.113 & 3.452 & 1.347 & 0.583 \\
1814 <      & 0.1 & 1.166 & 0.240 & 0.076 & 3.381 & 1.281 & 0.575 \\
1815 <      & 0.2 & 0.459 & 0.165 & 0.153 & 3.430 & 2.542 & 2.273 \\
1816 <      & 0.3 & 0.814 & 0.825 & 0.816 & 8.325 & 8.447 & 8.132 \\
1817 < \bottomrule
1818 < \end{tabular}
1819 < \label{tab:argonAng}
1820 < \end{table}
1821 <
1822 < This system does not appear to show any significant deviations from
1823 < the previously observed results. The {\sc sp} and {\sc sf} methods
1824 < have agreements similar to those observed in section
1825 < \ref{sec:WaterResults}. The only significant difference is the
1826 < improvement in the configuration energy differences for the {\sc rf}
1827 < method. This is surprising in that we are introducing an inhomogeneity
1828 < to the system; however, this inhomogeneity is charge-neutral and does
1829 < not result in charged cutoff spheres. The charge-neutrality of the
1830 < cutoff spheres, which the {\sc sp} and {\sc sf} methods explicitly
1831 < enforce, seems to play a greater role in the stability of the {\sc rf}
1832 < method than the required homogeneity of the environment.
1833 <
1834 <
1835 < \section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics}
1836 <
1837 < Zahn {\it et al.} investigated the structure and dynamics of water
1838 < using eqs. (\ref{eq:ZahnPot}) and
1839 < (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
1840 < that a method similar (but not identical with) the damped {\sc sf}
1841 < method resulted in properties very similar to those obtained when
1842 < using the Ewald summation.  The properties they studied (pair
1843 < distribution functions, diffusion constants, and velocity and
1844 < orientational correlation functions) may not be particularly sensitive
1845 < to the long-range and collective behavior that governs the
1846 < low-frequency behavior in crystalline systems.  Additionally, the
1847 < ionic crystals are the worst case scenario for the pairwise methods
1848 < because they lack the reciprocal space contribution contained in the
1849 < Ewald summation.  
1850 <
1851 < We are using two separate measures to probe the effects of these
1852 < alternative electrostatic methods on the dynamics in crystalline
1853 < materials.  For short- and intermediate-time dynamics, we are
1854 < computing the velocity autocorrelation function, and for long-time
1855 < and large length-scale collective motions, we are looking at the
1856 < low-frequency portion of the power spectrum.
1857 <
1858 < \begin{figure}
1859 < \centering
1860 < \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
1861 < \caption{Velocity autocorrelation functions of NaCl crystals at
1862 < 1000K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \&
1863 < 0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2\AA$^{-1}$). The inset is
1864 < a magnification of the area around the first minimum.  The times to
1865 < first collision are nearly identical, but differences can be seen in
1866 < the peaks and troughs, where the undamped and weakly damped methods
1867 < are stiffer than the moderately damped and {\sc spme} methods.}
1868 < \label{fig:vCorrPlot}
1869 < \end{figure}
1870 <
1871 < The short-time decay of the velocity autocorrelation function through
1872 < the first collision are nearly identical in figure
1873 < \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1874 < how the methods differ.  The undamped {\sc sf} method has deeper
1875 < troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1876 < any of the other methods.  As the damping parameter ($\alpha$) is
1877 < increased, these peaks are smoothed out, and the {\sc sf} method
1878 < approaches the {\sc spme} results.  With $\alpha$ values of 0.2\AA$^{-1}$,
1879 < the {\sc sf} and {\sc sp} functions are nearly identical and track the
1880 < {\sc spme} features quite well.  This is not surprising because the {\sc sf}
1881 < and {\sc sp} potentials become nearly identical with increased
1882 < damping.  However, this appears to indicate that once damping is
1883 < utilized, the details of the form of the potential (and forces)
1884 < constructed out of the damped electrostatic interaction are less
1885 < important.
1886 <
1887 < \section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics}
1888 <
1889 < To evaluate how the differences between the methods affect the
1890 < collective long-time motion, we computed power spectra from long-time
1891 < traces of the velocity autocorrelation function. The power spectra for
1892 < the best-performing alternative methods are shown in
1893 < fig. \ref{fig:methodPS}.  Apodization of the correlation functions via
1894 < a cubic switching function between 40 and 50 ps was used to reduce the
1895 < ringing resulting from data truncation.  This procedure had no
1896 < noticeable effect on peak location or magnitude.
1897 <
1898 < \begin{figure}
1899 < \centering
1900 < \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
1901 < \caption{Power spectra obtained from the velocity auto-correlation
1902 < functions of NaCl crystals at 1000K while using {\sc spme}, {\sc sf}
1903 < ($\alpha$ = 0, 0.1, \& 0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ =
1904 < 0.2\AA$^{-1}$).  The inset shows the frequency region below 100
1905 < cm$^{-1}$ to highlight where the spectra differ.}
1906 < \label{fig:methodPS}
1907 < \end{figure}
1908 <
1909 < While the high frequency regions of the power spectra for the
1910 < alternative methods are quantitatively identical with Ewald spectrum,
1911 < the low frequency region shows how the summation methods differ.
1912 < Considering the low-frequency inset (expanded in the upper frame of
1913 < figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1914 < correlated motions are blue-shifted when using undamped or weakly
1915 < damped {\sc sf}.  When using moderate damping ($\alpha = 0.2$
1916 < \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1917 < correlated motion to the Ewald method (which has a convergence
1918 < parameter of 0.3119\AA$^{-1}$).  This weakening of the electrostatic
1919 < interaction with increased damping explains why the long-ranged
1920 < correlated motions are at lower frequencies for the moderately damped
1921 < methods than for undamped or weakly damped methods.
1922 <
1923 < To isolate the role of the damping constant, we have computed the
1924 < spectra for a single method ({\sc sf}) with a range of damping
1925 < constants and compared this with the {\sc spme} spectrum.
1926 < Fig. \ref{fig:dampInc} shows more clearly that increasing the
1927 < electrostatic damping red-shifts the lowest frequency phonon modes.
1928 < However, even without any electrostatic damping, the {\sc sf} method
1929 < has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1930 < Without the {\sc sf} modifications, an undamped (pure cutoff) method
1931 < would predict the lowest frequency peak near 325 cm$^{-1}$.  {\it
1932 < Most} of the collective behavior in the crystal is accurately captured
1933 < using the {\sc sf} method.  Quantitative agreement with Ewald can be
1934 < obtained using moderate damping in addition to the shifting at the
1935 < cutoff distance.
1936 <
1937 < \begin{figure}
1938 < \centering
1939 < \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1940 < \caption{Effect of damping on the two lowest-frequency phonon modes in
1941 < the NaCl crystal at 1000K.  The undamped shifted force ({\sc sf})
1942 < method is off by less than 10 cm$^{-1}$, and increasing the
1943 < electrostatic damping to 0.25\AA$^{-1}$ gives quantitative agreement
1944 < with the power spectrum obtained using the Ewald sum.  Over-damping can
1945 < result in underestimates of frequencies of the long-wavelength
1946 < motions.}
1947 < \label{fig:dampInc}
1948 < \end{figure}
1949 <
1950 < \section{An Application: TIP5P-E Water}\label{sec:t5peApplied}
1951 <
1952 < The above sections focused on the energetics and dynamics of a variety
1953 < of systems when utilizing the {\sc sp} and {\sc sf} pairwise
1954 < techniques.  A unitary correlation with results obtained using the
1955 < Ewald summation should result in a successful reproduction of both the
1956 < static and dynamic properties of any selected system.  To test this,
1957 < we decided to calculate a series of properties for the TIP5P-E water
1958 < model when using the {\sc sf} technique.
1959 <
1960 < The TIP5P-E water model is a variant of Mahoney and Jorgensen's
1961 < five-point transferable intermolecular potential (TIP5P) model for
1962 < water.\cite{Mahoney00} TIP5P was developed to reproduce the density
1963 < maximum anomaly present in liquid water near 4$^\circ$C. As with many
1964 < previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
1965 < and SPC/E), TIP5P was parametrized using a simple cutoff with no
1966 < long-range electrostatic
1967 < correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
1968 < Without this correction, the pressure term on the central particle
1969 < from the surroundings is missing. Because they expand to compensate
1970 < for this added pressure term when this correction is included, systems
1971 < composed of these particles tend to underpredict the density of water
1972 < under standard conditions. When using any form of long-range
1973 < electrostatic correction, it has become common practice to develop or
1974 < utilize a reparametrized water model that corrects for this
1975 < effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
1976 < this practice and was optimized specifically for use with the Ewald
1977 < summation.\cite{Rick04} In his publication, Rick preserved the
1978 < geometry and point charge magnitudes in TIP5P and focused on altering
1979 < the Lennard-Jones parameters to correct the density at
1980 < 298K.\cite{Rick04} With the density corrected, he compared common
1981 < water properties for TIP5P-E using the Ewald sum with TIP5P using a
1982 < 9\AA\ cutoff.
1983 <
1984 < In the following sections, we compared these same water properties
1985 < calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
1986 < {\sc sf} technique.  In the above evaluation of the pairwise
1987 < techniques, we observed some flexibility in the choice of parameters.
1988 < Because of this, the following comparisons include the {\sc sf}
1989 < technique with a 12\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
1990 < 0.2\AA$^{-1}$, as well as a 9\AA\ cutoff with an $\alpha$ =
1991 < 0.2\AA$^{-1}$.
1992 <
1993 < \subsection{Density}\label{sec:t5peDensity}
1994 <
1995 < As stated previously, the property that prompted the development of
1996 < TIP5P-E was the density at 1 atm.  The density depends upon the
1997 < internal pressure of the system in the $NPT$ ensemble, and the
1998 < calculation of the pressure includes a components from both the
1999 < kinetic energy and the virial. More specifically, the instantaneous
2000 < molecular pressure ($p(t)$) is given by
2001 < \begin{equation}
2002 < p(t) =  \frac{1}{\textrm{d}V}\sum_\mu
2003 <        \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
2004 <        + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
2005 < \label{eq:MolecularPressure}
2006 < \end{equation}
2007 < where $V$ is the volume, $\mathbf{P}_{\mu}$ is the momentum of
2008 < molecule $\mu$, $\mathbf{R}_\mu$ is the position of the center of mass
2009 < ($M_\mu$) of molecule $\mu$, and $\mathbf{F}_{\mu i}$ is the force on
2010 < atom $i$ of molecule $\mu$.\cite{Melchionna93} The virial term (the
2011 < right term in the brackets of eq. \ref{eq:MolecularPressure}) is
2012 < directly dependent on the interatomic forces.  Since the {\sc sp}
2013 < method does not modify the forces (see
2014 < sec. \ref{sec:PairwiseDerivation}), the pressure using {\sc sp} will
2015 < be identical to that obtained without an electrostatic correction.
2016 < The {\sc sf} method does alter the virial component and, by way of the
2017 < modified pressures, should provide densities more in line with those
2018 < obtained using the Ewald summation.
2019 <
2020 < To compare densities, $NPT$ simulations were performed with the same
2021 < temperatures as those selected by Rick in his Ewald summation
2022 < simulations.\cite{Rick04} In order to improve statistics around the
2023 < density maximum, 3ns trajectories were accumulated at 0, 12.5, and
2024 < 25$^\circ$C, while 2ns trajectories were obtained at all other
2025 < temperatures. The average densities were calculated from the later
2026 < three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
2027 < method for accumulating statistics, these sequences were spliced into
2028 < 200 segments to calculate the average density and standard deviation
2029 < at each temperature.\cite{Mahoney00}
2030 <
2031 < \begin{figure}
2032 < \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
2033 < \caption{Density versus temperature for the TIP5P-E water model when
2034 < using the Ewald summation (Ref. \cite{Rick04}) and the {\sc sf} method
2035 < with various parameters. The pressure term from the image-charge shell
2036 < is larger than that provided by the reciprocal-space portion of the
2037 < Ewald summation, leading to slightly lower densities. This effect is
2038 < more visible with the 9\AA\ cutoff, where the image charges exert a
2039 < greater force on the central particle. The error bars for the {\sc sf}
2040 < methods show plus or minus the standard deviation of the density
2041 < measurement at each temperature.}
2042 < \label{fig:t5peDensities}
2043 < \end{figure}
2044 <
2045 < Figure \ref{fig:t5peDensities} shows the densities calculated for
2046 < TIP5P-E using differing electrostatic corrections overlaid on the
2047 < experimental values.\cite{CRC80} The densities when using the {\sc sf}
2048 < technique are close to, though typically lower than, those calculated
2049 < while using the Ewald summation. These slightly reduced densities
2050 < indicate that the pressure component from the image charges at
2051 < R$_\textrm{c}$ is larger than that exerted by the reciprocal-space
2052 < portion of the Ewald summation. Bringing the image charges closer to
2053 < the central particle by choosing a 9\AA\ R$_\textrm{c}$ (rather than
2054 < the preferred 12\AA\ R$_\textrm{c}$) increases the strength of their
2055 < interactions, resulting in a further reduction of the densities.
2056 <
2057 < Because the strength of the image charge interactions has a noticable
2058 < effect on the density, we would expect the use of electrostatic
2059 < damping to also play a role in these calculations. Larger values of
2060 < $\alpha$ weaken the pair-interactions; and since electrostatic damping
2061 < is distance-dependent, force components from the image charges will be
2062 < reduced more than those from particles close the the central
2063 < charge. This effect is visible in figure \ref{fig:t5peDensities} with
2064 < the damped {\sc sf} sums showing slightly higher densities; however,
2065 < it is apparent that the choice of cutoff radius plays a much more
2066 < important role in the resulting densities.
2067 <
2068 < As a final note, all of the above density calculations were performed
2069 < with systems of 512 water molecules. Rick observed a system sized
2070 < dependence of the computed densities when using the Ewald summation,
2071 < most likely due to his tying of the convergence parameter to the box
2072 < dimensions.\cite{Rick04} For systems of 256 water molecules, the
2073 < calculated densities were on average 0.002 to 0.003 g/cm$^3$ lower. A
2074 < system size of 256 molecules would force the use of a shorter
2075 < R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
2076 < lower the densities. Moving to larger systems, as long as the
2077 < R$_\textrm{c}$ remains at a fixed value, we would expect the densities
2078 < to remain constant.
2079 <
2080 < \subsection{Liquid Structure}\label{sec:t5peLiqStructure}
2081 <
2082 < A common function considered when developing and comparing water
2083 < models is the oxygen-oxygen radial distribution function
2084 < ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
2085 < finding a pair of oxygen atoms some distance ($r$) apart relative to a
2086 < random distribution at the same density.\cite{Allen87} It is
2087 < calculated via
2088 < \begin{equation}
2089 < g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
2090 <        \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
2091 < \label{eq:GOOofR}
2092 < \end{equation}
2093 < where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
2094 < atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
2095 < neutron scattering experiments through the oxygen-oxygen structure
2096 < factor ($S_\textrm{OO}(k)$) by the following relationship:
2097 < \begin{equation}
2098 < S_\textrm{OO}(k) = 1 + 4\pi\rho
2099 <        \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
2100 < \label{eq:SOOofK}
2101 < \end{equation}
2102 < Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
2103 < of $g_\textrm{OO}(r)$.
2104 <
2105 < The expermentally determined $g_\textrm{OO}(r)$ for liquid water has
2106 < been compared in great detail with the various common water models,
2107 < and TIP5P was found to be in better agreement than other rigid,
2108 < non-polarizable models.\cite{Sorenson00} This excellent agreement with
2109 < experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
2110 < check whether the choice of using the Ewald summation or the {\sc sf}
2111 < technique alters the liquid structure, the $g_\textrm{OO}(r)$s at 298K
2112 < and 1atm were determined for the systems compared in the previous
2113 < section.
2114 <
2115 < \begin{figure}
2116 < \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
2117 < \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298K and
2118 < 1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
2119 < sf} technique with varying parameters. Even with the reduced densities
2120 < using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
2121 < identical.}
2122 < \label{fig:t5peGofRs}
2123 < \end{figure}
2124 <
2125 < The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
2126 < sf} technique with a various parameters are overlaid on the
2127 < $g_\textrm{OO}(r)$ while using the Ewald summation. The differences in
2128 < density do not appear to have any effect on the liquid structure as
2129 < the $g_\textrm{OO}(r)$s are indistinquishable. These results indicate
2130 < that the $g_\textrm{OO}(r)$ is insensitive to the choice of
2131 < electrostatic correction.
2132 <
2133 < \subsection{Thermodynamic Properties}\label{sec:t5peThermo}
2134 <
2135 < In addition to the density, there are a variety of thermodynamic
2136 < quantities that can be calculated for water and compared directly to
2137 < experimental values. Some of these additional quatities include the
2138 < latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
2139 < pressure heat capacity ($C_p$), the isothermal compressibility
2140 < ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
2141 < dielectric constant ($\epsilon$). All of these properties were
2142 < calculated for TIP5P-E with the Ewald summation, so they provide a
2143 < good set for comparisons involving the {\sc sf} technique.
2144 <
2145 < The $\Delta H_\textrm{vap}$ is the enthalpy change required to
2146 < transform one mol of substance from the liquid phase to the gas
2147 < phase.\cite{Berry00} In molecular simulations, this quantity can be
2148 < determined via
2149 < \begin{equation}
2150 < \begin{split}
2151 < \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
2152 <                      &= E_\textrm{gas} - E_\textrm{liq.}
2153 <                        + p(V_\textrm{gas} - V_\textrm{liq.}) \\
2154 <                      &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
2155 < \end{split}
2156 < \label{eq:DeltaHVap}
2157 < \end{equation}
2158 < where $E$ is the total energy, $U$ is the potential energy, $p$ is the
2159 < pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
2160 < the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
2161 < seen in the last line of equation \ref{eq:DeltaHVap}, we can
2162 < approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
2163 < state. This allows us to cancel the kinetic energy terms, leaving only
2164 < the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
2165 < several orders of magnitude larger than that of the liquid, so we can
2166 < neglect the liquid $pV$ term.
2167 <
2168 < The remaining thermodynamic properties can all be calculated from
2169 < fluctuations of the enthalpy, volume, and system dipole
2170 < moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
2171 < enthalpy in constant pressure simulations via
2172 < \begin{equation}
2173 < \begin{split}
2174 < C_p     = \left(\frac{\partial H}{\partial T}\right)_{N,p}
2175 <        = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
2176 < \end{split}
2177 < \label{eq:Cp}
2178 < \end{equation}
2179 < where $k_B$ is Boltzmann's constant.  $\kappa_T$ can be calculated via
2180 < \begin{equation}
2181 < \begin{split}
2182 < \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
2183 <         = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
2184 <                {k_BT\langle V\rangle_{N,P,T}},
2185 < \end{split}
2186 < \label{eq:kappa}
2187 < \end{equation}
2188 < and $\alpha_p$ can be calculated via
2189 < \begin{equation}
2190 < \begin{split}
2191 < \alpha_p        = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
2192 <                = \frac{(\langle VH\rangle_{N,P,T}
2193 <                        - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
2194 <                        {k_BT^2\langle V\rangle_{N,P,T}}.
2195 < \end{split}
2196 < \label{eq:alpha}
2197 < \end{equation}
2198 < Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
2199 < be calculated for systems of non-polarizable substances via
2200 < \begin{equation}
2201 < \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
2202 < \label{eq:staticDielectric}
2203 < \end{equation}
2204 < where $\epsilon_0$ is the permittivity of free space and $\langle
2205 < M^2\rangle$ is the fluctuation of the system dipole
2206 < moment.\cite{Allen87} The numerator in the fractional term in equation
2207 < \ref{eq:staticDielectric} is the fluctuation of the simulation-box
2208 < dipole moment, identical to the quantity calculated in the
2209 < finite-system Kirkwood $g$ factor ($G_k$):
2210 < \begin{equation}
2211 < G_k = \frac{\langle M^2\rangle}{N\mu^2},
2212 < \label{eq:KirkwoodFactor}
2213 < \end{equation}
2214 < where $\mu$ is the dipole moment of a single molecule of the
2215 < homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
2216 < fluctuation term in both equation \ref{eq:staticDielectric} and
2217 < \ref{eq:KirkwoodFactor} is calculated as follows,
2218 < \begin{equation}
2219 < \begin{split}
2220 < \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
2221 <                        - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
2222 <                   &= \langle M_x^2+M_y^2+M_z^2\rangle
2223 <                        - (\langle M_x\rangle^2 + \langle M_x\rangle^2
2224 <                                + \langle M_x\rangle^2).        
2225 < \end{split}
2226 < \label{eq:fluctBoxDipole}
2227 < \end{equation}
2228 < This fluctuation term can be accumulated during the simulation;
2229 < however, it converges rather slowly, thus requiring multi-nanosecond
2230 < simulation times.\cite{Horn04} In the case of tin-foil boundary
2231 < conditions, the dielectric/surface term of equation \ref{eq:EwaldSum}
2232 < is equal to zero. Since the {\sc sf} method also lacks this
2233 < dielectric/surface term, equation \ref{eq:staticDielectric} is still
2234 < valid for determining static dielectric constants.
2235 <
2236 < All of the above properties were calculated from the same trajectories
2237 < used to determine the densities in section \ref{sec:t5peDensity}
2238 < except for the static dielectric constants. The $\epsilon$ values were
2239 < accumulated from 2ns $NVE$ ensemble trajectories with system densities
2240 < fixed at the average values from the $NPT$ simulations at each of the
2241 < temperatures. The resulting values are displayed in figure
2242 < \ref{fig:t5peThermo}.
2243 < \begin{figure}
2244 < \centering
2245 < \includegraphics[width=5.5in]{./figures/t5peThermo.pdf}
2246 < \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
2247 < and the {\sc sf} techniques along with the experimental values. Units
2248 < for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
2249 < cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
2250 < and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
2251 < reference \cite{Rick04}. Experimental values for $\Delta
2252 < H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
2253 < \cite{Kell75}. Experimental values for $C_p$ are from reference
2254 < \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
2255 < \cite{Malmberg56}.}
2256 < \label{fig:t5peThermo}
2257 < \end{figure}
2258 <
2259 < As observed for the density in section \ref{sec:t5peDensity}, the
2260 < property trends with temperature seen when using the Ewald summation
2261 < are reproduced with the {\sc sf} technique. Differences include the
2262 < calculated values of $\Delta H_\textrm{vap}$ underpredicting the Ewald
2263 < values. This is to be expected due to the direct weakening of the
2264 < electrostatic interaction through forced neutralization in {\sc
2265 < sf}. This results in an increase of the intermolecular potential
2266 < producing lower values from equation \ref{eq:DeltaHVap}. The slopes of
2267 < these values with temperature are similar to that seen using the Ewald
2268 < summation; however, they are both steeper than the experimental trend,
2269 < indirectly resulting in the inflated $C_p$ values at all temperatures.
2270 <
2271 < Above the supercooled regim\'{e}, $C_p$, $\kappa_T$, and $\alpha_p$
2272 < values all overlap within error. As indicated for the $\Delta
2273 < H_\textrm{vap}$ and $C_p$ results discussed in the previous paragraph,
2274 < the deviations between experiment and simulation in this region are
2275 < not the fault of the electrostatic summation methods but are due to
2276 < the TIP5P class model itself. Like most rigid, non-polarizable,
2277 < point-charge water models, the density decreases with temperature at a
2278 < much faster rate than experiment (see figure
2279 < \ref{fig:t5peDensities}). The reduced density leads to the inflated
2280 < compressibility and expansivity values at higher temperatures seen
2281 < here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
2282 < and many-body effects are required in order for simulation to overcome
2283 < these differences with experiment.\cite{Laasonen93,Donchev06}
2284 <
2285 < At temperatures below the freezing point for experimental water, the
2286 < differences between {\sc sf} and the Ewald summation results are more
2287 < apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
2288 < indicate a more pronounced transition in the supercooled regim\'{e},
2289 < particularly in the case of {\sc sf} without damping. This points to
2290 < the onset of a more frustrated or glassy behavior for TIP5P-E at
2291 < temperatures below 250K in these simulations. Because the systems are
2292 < locked in different regions of phase-space, comparisons between
2293 < properties at these temperatures are not exactly fair. This
2294 < observation is explored in more detail in section
2295 < \ref{sec:t5peDynamics}.
2296 <
2297 < The final thermodynamic property displayed in figure
2298 < \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
2299 < between the Ewald summation and the {\sc sf} technique (and experiment
2300 < for that matter). It is known that the dielectric constant is
2301 < dependent upon and quite sensitive to the imposed boundary
2302 < conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
2303 < converged $\epsilon$ values accumulated for the {\sc sf}
2304 < simulations. Lack of a damping function results in dielectric
2305 < constants significantly smaller than that obtained using the Ewald
2306 < sum. Increasing the damping coefficient to 0.2\AA$^{-1}$ improves the
2307 < agreement considerably. It should be noted that the choice of the
2308 < ``Ewald coefficient'' value also has a significant effect on the
2309 < calculated value when using the Ewald summation. In the simulations of
2310 < TIP5P-E with the Ewald sum, this screening parameter was tethered to
2311 < the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04}
2312 < Systems with larger screening parameters reported larger dielectric
2313 < constant values, the same behavior we see here with {\sc sf}. In
2314 < section \ref{sec:dampingDielectric}, this connection is further
2315 < explored as optimal damping coefficients are determined for {\sc
2316 < sf} for capturing the dielectric behavior.
2317 <
2318 < \subsection{Dynamic Properties}\label{sec:t5peDynamics}
2319 <
2320 < To look at the dynamic properties of TIP5P-E when using the {\sc sf}
2321 < method, 200ps $NVE$ simulations were performed for each temperature at
2322 < the average density reported by the $NPT$ simulations. The
2323 < self-diffusion constants ($D$) were calculated with the Einstein
2324 < relation using the mean square displacement (MSD),
2325 < \begin{equation}
2326 < D = \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
2327 < \label{eq:MSD}
2328 < \end{equation}
2329 < where $t$ is time, and $\mathbf{r}_i$ is the position of particle
2330 < $i$.\cite{Allen87} Figure \ref{fig:ExampleMSD} shows an example MSD
2331 < plot. As labeled in the figure, MSD plots consist of three distinct
2332 < regions:
2333 <
2334 < \begin{enumerate}[itemsep=0pt]
2335 < \item parabolic short-time ballistic motion,
2336 < \item linear diffusive regime, and
2337 < \item poor statistic region at long-time.
2338 < \end{enumerate}
2339 < The slope from the linear region (region 2) is used to calculate $D$.
2340 < \begin{figure}
2341 < \centering
2342 < \includegraphics[width=3.5in]{./figures/ExampleMSD.pdf}
2343 < \caption{Example plot of mean square displacement verses time. The
2344 < left red region is the ballistic motion regime, the middle green
2345 < region is the linear diffusive regime, and the right blue region is
2346 < the region with poor statistics.}
2347 < \label{fig:ExampleMSD}
2348 < \end{figure}
2349 <
2350 < \begin{figure}
2351 < \centering
2352 < \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
2353 < \caption{Body-fixed coordinate frame for a water molecule. The
2354 < respective molecular principle axes point in the direction of the
2355 < labeled frame axes.}
2356 < \label{fig:waterFrame}
2357 < \end{figure}
2358 < In addition to translational diffusion, reorientational time constants
2359 < were calculated for comparisons with the Ewald simulations and with
2360 < experiments. These values were determined from 25ps $NVE$ trajectories
2361 < through calculation of the orientational time correlation function,
2362 < \begin{equation}
2363 < C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
2364 <                \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
2365 < \label{eq:OrientCorr}
2366 < \end{equation}
2367 < where $P_l$ is the Legendre polynomial of order $l$ and
2368 < $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
2369 < principle axis $\alpha$. The principle axis frame for these water
2370 < molecules is shown in figure \ref{fig:waterFrame}. As an example,
2371 < $C_l^y$ is calculated from the time evolution of the unit vector
2372 < connecting the two hydrogen atoms.
2373 <
2374 < \begin{figure}
2375 < \centering
2376 < \includegraphics[width=3.5in]{./figures/ExampleOrientCorr.pdf}
2377 < \caption{Example plots of the orientational autocorrelation functions
2378 < for the first and second Legendre polynomials. These curves show the
2379 < time decay of the unit vector along the $y$ principle axis.}
2380 < \label{fig:OrientCorr}
2381 < \end{figure}
2382 < From the orientation autocorrelation functions, we can obtain time
2383 < constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
2384 < some example plots of orientational autocorrelation functions for the
2385 < first and second Legendre polynomials. The relatively short time
2386 < portions (between 1 and 3ps for water) of these curves can be fit to
2387 < an exponential decay to obtain these constants, and they are directly
2388 < comparable to water orientational relaxation times from nuclear
2389 < magnetic resonance (NMR). The relaxation constant obtained from
2390 < $C_2^y(t)$ is of particular interest because it is about the principle
2391 < axis with the minimum moment of inertia and should thereby dominate
2392 < the orientational relaxation of the molecule.\cite{Impey82} This means
2393 < that $C_2^y(t)$ should provide the best comparison to the NMR
2394 < relaxation data.
2395 <
2396 < \begin{figure}
2397 < \centering
2398 < \includegraphics[width=5.5in]{./figures/t5peDynamics.pdf}
2399 < \caption{Diffusion constants ({\it upper}) and reorientational time
2400 < constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
2401 < technique compared with experiment. Data at temperatures less that
2402 < 0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
2403 < easier comparisons in the more relavent temperature regime.}
2404 < \label{fig:t5peDynamics}
2405 < \end{figure}
2406 < Results for the diffusion constants and reorientational time constants
2407 < are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
2408 < apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
2409 < the Ewald sum are reproduced with the {\sc sf} techinque. The enhanced
2410 < diffusion at high temperatures are again the product of the lower
2411 < densities in comparison with experiment and do not provide any special
2412 < insight into differences between the electrostatic summation
2413 < techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
2414 < diffuse a little faster than with the Ewald sum; however, use of light
2415 < to moderate damping results in indistiguishable $D$ values. Though not
2416 < apparent in this figure, {\sc sf} values at the lowest temperature are
2417 < approximately an order of magnitude lower than with Ewald. These
2418 < values support the observation from section \ref{sec:t5peThermo} that
2419 < there appeared to be a change to a more glassy-like phase with the
2420 < {\sc sf} technique at these lower temperatures.
2421 <
2422 < The $\tau_2^y$ results in the lower frame of figure
2423 < \ref{fig:t5peDynamics} show a much greater difference between the {\sc
2424 < sf} results and the Ewald results. At all temperatures shown, TIP5P-E
2425 < relaxes faster than experiment with the Ewald sum while tracking
2426 < experiment fairly well when using the {\sc sf} technique, independent
2427 < of the choice of damping constant. Their are several possible reasons
2428 < for this deviation between techniques. The Ewald results were taken
2429 < shorter (10ps) trajectories than the {\sc sf} results (25ps). A quick
2430 < calculation from a 10ps trajectory with {\sc sf} with an $\alpha$ of
2431 < 0.2\AA$^-1$ at 25$^\circ$C showed a 0.4ps drop in $\tau_2^y$, placing
2432 < the result more in line with that obtained using the Ewald sum. These
2433 < results support this explanation; however, recomputing the results to
2434 < meet a poorer statistical standard is counter-productive. Assuming the
2435 < Ewald results are not the product of poor statistics, differences in
2436 < techniques to integrate the orientational motion could also play a
2437 < role. {\sc shake} is the most commonly used technique for
2438 < approximating rigid-body orientational motion,\cite{Ryckaert77} where
2439 < as in {\sc oopse}, we maintain and integrate the entire rotation
2440 < matrix using the {\sc dlm} method.\cite{Meineke05} Since {\sc shake}
2441 < is an iterative constraint technique, if the convergence tolerances
2442 < are raised for increased performance, error will accumulate in the
2443 < orientational motion. Finally, the Ewald results were calculated using
2444 < the $NVT$ ensemble, while the $NVE$ ensemble was used for {\sc sf}
2445 < calculations. The additional mode of motion due to the thermostat will
2446 < alter the dynamics, resulting in differences between $NVT$ and $NVE$
2447 < results. These differences are increasingly noticable as the
2448 < thermostat time constant decreases.
2449 <
2450 < \section{Damping of Point Multipoles}\label{sec:dampingMultipoles}
2451 <
2452 <
2453 <
2454 < \section{Damping and Dielectric Constants}\label{sec:dampingDielectric}
2455 <
2456 < \section{Conclusions}\label{sec:PairwiseConclusions}
2457 <
2458 < The above investigation of pairwise electrostatic summation techniques
2459 < shows that there are viable and computationally efficient alternatives
2460 < to the Ewald summation.  These methods are derived from the damped and
2461 < cutoff-neutralized Coulombic sum originally proposed by Wolf
2462 < \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
2463 < method, reformulated above as eqs. (\ref{eq:DSFPot}) and
2464 < (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
2465 < energetic and dynamic characteristics exhibited by simulations
2466 < employing lattice summation techniques.  The cumulative energy
2467 < difference results showed the undamped {\sc sf} and moderately damped
2468 < {\sc sp} methods produced results nearly identical to {\sc spme}.
2469 < Similarly for the dynamic features, the undamped or moderately damped
2470 < {\sc sf} and moderately damped {\sc sp} methods produce force and
2471 < torque vector magnitude and directions very similar to the expected
2472 < values.  These results translate into long-time dynamic behavior
2473 < equivalent to that produced in simulations using {\sc spme}.
2474 <
2475 < As in all purely-pairwise cutoff methods, these methods are expected
2476 < to scale approximately {\it linearly} with system size, and they are
2477 < easily parallelizable.  This should result in substantial reductions
2478 < in the computational cost of performing large simulations.
2479 <
2480 < Aside from the computational cost benefit, these techniques have
2481 < applicability in situations where the use of the Ewald sum can prove
2482 < problematic.  Of greatest interest is their potential use in
2483 < interfacial systems, where the unmodified lattice sum techniques
2484 < artificially accentuate the periodicity of the system in an
2485 < undesirable manner.  There have been alterations to the standard Ewald
2486 < techniques, via corrections and reformulations, to compensate for
2487 < these systems; but the pairwise techniques discussed here require no
2488 < modifications, making them natural tools to tackle these problems.
2489 < Additionally, this transferability gives them benefits over other
2490 < pairwise methods, like reaction field, because estimations of physical
2491 < properties (e.g. the dielectric constant) are unnecessary.
2492 <
2493 < If a researcher is using Monte Carlo simulations of large chemical
2494 < systems containing point charges, most structural features will be
2495 < accurately captured using the undamped {\sc sf} method or the {\sc sp}
2496 < method with an electrostatic damping of 0.2\AA$^{-1}$.  These methods
2497 < would also be appropriate for molecular dynamics simulations where the
2498 < data of interest is either structural or short-time dynamical
2499 < quantities.  For long-time dynamics and collective motions, the safest
2500 < pairwise method we have evaluated is the {\sc sf} method with an
2501 < electrostatic damping between 0.2 and 0.25\AA$^{-1}$.
2502 <
2503 < We are not suggesting that there is any flaw with the Ewald sum; in
2504 < fact, it is the standard by which these simple pairwise sums have been
2505 < judged.  However, these results do suggest that in the typical
2506 < simulations performed today, the Ewald summation may no longer be
2507 < required to obtain the level of accuracy most researchers have come to
2508 < expect.
46 > \input{electrostaticsChapter}
47  
48   \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
49  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines