ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/dissertation.tex
Revision: 2927
Committed: Sat Jul 8 13:13:27 2006 UTC (17 years, 11 months ago) by chrisfen
Content type: application/x-tex
File size: 96775 byte(s)
Log Message:
Incorporated supporting info into pairwise chapter

File Contents

# User Rev Content
1 chrisfen 2918 \documentclass[12pt]{ndthesis}
2    
3     % some packages for things like equations and graphics
4     \usepackage{amsmath,bm}
5     \usepackage{amssymb}
6     \usepackage{mathrsfs}
7     \usepackage{tabularx}
8     \usepackage{graphicx}
9     \usepackage{booktabs}
10    
11     \begin{document}
12    
13     \frontmatter
14    
15     \title{APPLICATION AND DEVELOPMENT OF MOLECULAR DYNAMICS TECHNIQUES FOR THE
16     STUDY OF WATER}
17     \author{Christopher Joseph Fennell}
18     \work{Dissertation}
19     \degprior{B.Sc.}
20     \degaward{Doctor of Philosophy}
21     \advisor{J. Daniel Gezelter}
22     \department{Chemistry and Biochemistry}
23    
24     \maketitle
25    
26     \begin{abstract}
27     \end{abstract}
28    
29     \begin{dedication}
30     \end{dedication}
31    
32     \tableofcontents
33     \listoffigures
34     \listoftables
35    
36     \begin{acknowledge}
37     \end{acknowledge}
38    
39     \mainmatter
40    
41     \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
42    
43     \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION
44     TECHNIQUES}
45    
46     In molecular simulations, proper accumulation of the electrostatic
47     interactions is essential and is one of the most
48     computationally-demanding tasks. The common molecular mechanics force
49     fields represent atomic sites with full or partial charges protected
50     by Lennard-Jones (short range) interactions. This means that nearly
51     every pair interaction involves a calculation of charge-charge forces.
52     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
53     interactions quickly become the most expensive part of molecular
54     simulations. Historically, the electrostatic pair interaction would
55     not have decayed appreciably within the typical box lengths that could
56     be feasibly simulated. In the larger systems that are more typical of
57     modern simulations, large cutoffs should be used to incorporate
58     electrostatics correctly.
59    
60     There have been many efforts to address the proper and practical
61     handling of electrostatic interactions, and these have resulted in a
62     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
63     typically classified as implicit methods (i.e., continuum dielectrics,
64     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
65     (i.e., Ewald summations, interaction shifting or
66     truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
67     reaction field type methods, fast multipole
68     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
69     often preferred because they physically incorporate solvent molecules
70     in the system of interest, but these methods are sometimes difficult
71     to utilize because of their high computational cost.\cite{Roux99} In
72     addition to the computational cost, there have been some questions
73     regarding possible artifacts caused by the inherent periodicity of the
74     explicit Ewald summation.\cite{Tobias01}
75    
76     In this chapter, we focus on a new set of pairwise methods devised by
77     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
78     methods along with a few other mixed methods (i.e. reaction field) are
79     compared with the smooth particle mesh Ewald
80     sum,\cite{Onsager36,Essmann99} which is our reference method for
81     handling long-range electrostatic interactions. The new methods for
82     handling electrostatics have the potential to scale linearly with
83     increasing system size since they involve only a simple modification
84     to the direct pairwise sum. They also lack the added periodicity of
85     the Ewald sum, so they can be used for systems which are non-periodic
86     or which have one- or two-dimensional periodicity. Below, these
87     methods are evaluated using a variety of model systems to
88     establish their usability in molecular simulations.
89    
90     \section{The Ewald Sum}
91    
92     The complete accumulation of the electrostatic interactions in a system with
93     periodic boundary conditions (PBC) requires the consideration of the
94     effect of all charges within a (cubic) simulation box as well as those
95     in the periodic replicas,
96     \begin{equation}
97     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
98     \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
99     \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
100     \right],
101     \label{eq:PBCSum}
102     \end{equation}
103     where the sum over $\mathbf{n}$ is a sum over all periodic box
104     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
105     prime indicates $i = j$ are neglected for $\mathbf{n} =
106     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
107     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
108     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
109     $j$, and $\phi$ is the solution to Poisson's equation
110     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
111     charge-charge interactions). In the case of monopole electrostatics,
112     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
113     non-neutral systems.
114    
115     The electrostatic summation problem was originally studied by Ewald
116     for the case of an infinite crystal.\cite{Ewald21}. The approach he
117     took was to convert this conditionally convergent sum into two
118     absolutely convergent summations: a short-ranged real-space summation
119     and a long-ranged reciprocal-space summation,
120     \begin{equation}
121     \begin{split}
122     V_\textrm{elec} = \frac{1}{2}&
123     \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
124     \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
125     {|\mathbf{r}_{ij}+\mathbf{n}|} \\
126     &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
127     \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
128     \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
129     &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
130     + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
131     \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
132     \end{split}
133     \label{eq:EwaldSum}
134     \end{equation}
135     where $\alpha$ is the damping or convergence parameter with units of
136     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
137     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
138     constant of the surrounding medium. The final two terms of
139     eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
140     for interacting with a surrounding dielectric.\cite{Allen87} This
141     dipolar term was neglected in early applications in molecular
142     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
143     Leeuw {\it et al.} to address situations where the unit cell has a
144     dipole moment which is magnified through replication of the periodic
145     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
146     system is said to be using conducting (or ``tin-foil'') boundary
147     conditions, $\epsilon_{\rm S} = \infty$. Figure
148     \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
149     time. Initially, due to the small system sizes that could be
150     simulated feasibly, the entire simulation box was replicated to
151     convergence. In more modern simulations, the systems have grown large
152     enough that a real-space cutoff could potentially give convergent
153     behavior. Indeed, it has been observed that with the choice of a
154     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
155     rapidly convergent and small relative to the real-space
156     portion.\cite{Karasawa89,Kolafa92}
157    
158     \begin{figure}
159     \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
160     \caption{The change in the need for the Ewald sum with
161     increasing computational power. A:~Initially, only small systems
162     could be studied, and the Ewald sum replicated the simulation box to
163     convergence. B:~Now, radial cutoff methods should be able to reach
164     convergence for the larger systems of charges that are common today.}
165     \label{fig:ewaldTime}
166     \end{figure}
167    
168     The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
169     convergence parameter $(\alpha)$ plays an important role in balancing
170     the computational cost between the direct and reciprocal-space
171     portions of the summation. The choice of this value allows one to
172     select whether the real-space or reciprocal space portion of the
173     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
174     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
175     $\alpha$ and thoughtful algorithm development, this cost can be
176     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
177     taken to reduce the cost of the Ewald summation even further is to set
178     $\alpha$ such that the real-space interactions decay rapidly, allowing
179     for a short spherical cutoff. Then the reciprocal space summation is
180     optimized. These optimizations usually involve utilization of the
181     fast Fourier transform (FFT),\cite{Hockney81} leading to the
182     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
183     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
184     methods, the cost of the reciprocal-space portion of the Ewald
185     summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
186     \log N)$.
187    
188     These developments and optimizations have made the use of the Ewald
189     summation routine in simulations with periodic boundary
190     conditions. However, in certain systems, such as vapor-liquid
191     interfaces and membranes, the intrinsic three-dimensional periodicity
192     can prove problematic. The Ewald sum has been reformulated to handle
193     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
194     methods are computationally expensive.\cite{Spohr97,Yeh99} More
195     recently, there have been several successful efforts toward reducing
196     the computational cost of 2-D lattice
197     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
198     bringing them more in line with the cost of the full 3-D summation.
199    
200     Several studies have recognized that the inherent periodicity in the
201     Ewald sum can also have an effect on three-dimensional
202     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
203     Solvated proteins are essentially kept at high concentration due to
204     the periodicity of the electrostatic summation method. In these
205     systems, the more compact folded states of a protein can be
206     artificially stabilized by the periodic replicas introduced by the
207     Ewald summation.\cite{Weber00} Thus, care must be taken when
208     considering the use of the Ewald summation where the assumed
209     periodicity would introduce spurious effects in the system dynamics.
210    
211    
212     \section{The Wolf and Zahn Methods}
213    
214     In a recent paper by Wolf \textit{et al.}, a procedure was outlined
215     for the accurate accumulation of electrostatic interactions in an
216     efficient pairwise fashion. This procedure lacks the inherent
217     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
218     observed that the electrostatic interaction is effectively
219     short-ranged in condensed phase systems and that neutralization of the
220     charge contained within the cutoff radius is crucial for potential
221     stability. They devised a pairwise summation method that ensures
222     charge neutrality and gives results similar to those obtained with the
223     Ewald summation. The resulting shifted Coulomb potential includes
224     image-charges subtracted out through placement on the cutoff sphere
225     and a distance-dependent damping function (identical to that seen in
226     the real-space portion of the Ewald sum) to aid convergence
227     \begin{equation}
228     V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
229     - \lim_{r_{ij}\rightarrow R_\textrm{c}}
230     \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
231     \label{eq:WolfPot}
232     \end{equation}
233     Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
234     potential. However, neutralizing the charge contained within each
235     cutoff sphere requires the placement of a self-image charge on the
236     surface of the cutoff sphere. This additional self-term in the total
237     potential enabled Wolf {\it et al.} to obtain excellent estimates of
238     Madelung energies for many crystals.
239    
240     In order to use their charge-neutralized potential in molecular
241     dynamics simulations, Wolf \textit{et al.} suggested taking the
242     derivative of this potential prior to evaluation of the limit. This
243     procedure gives an expression for the forces,
244     \begin{equation}
245     \begin{split}
246     F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
247     \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
248     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
249     \Biggr]\\
250     &-\Biggr[
251     \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
252     + \frac{2\alpha}{\pi^{1/2}}
253     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
254     \Biggr]\Biggr\},
255     \end{split}
256     \label{eq:WolfForces}
257     \end{equation}
258     that incorporates both image charges and damping of the electrostatic
259     interaction.
260    
261     More recently, Zahn \textit{et al.} investigated these potential and
262     force expressions for use in simulations involving water.\cite{Zahn02}
263     In their work, they pointed out that the forces and derivative of
264     the potential are not commensurate. Attempts to use both
265     eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
266     to poor energy conservation. They correctly observed that taking the
267     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
268     derivatives gives forces for a different potential energy function
269     than the one shown in eq. (\ref{eq:WolfPot}).
270    
271     Zahn \textit{et al.} introduced a modified form of this summation
272     method as a way to use the technique in Molecular Dynamics
273     simulations. They proposed a new damped Coulomb potential,
274     \begin{equation}
275     \begin{split}
276     V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
277     \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
278     &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
279     + \frac{2\alpha}{\pi^{1/2}}
280     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
281     \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
282     \end{split}
283     \label{eq:ZahnPot}
284     \end{equation}
285     and showed that this potential does fairly well at capturing the
286     structural and dynamic properties of water compared the same
287     properties obtained using the Ewald sum.
288    
289     \section{Simple Forms for Pairwise Electrostatics}
290    
291     The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
292     al.} are constructed using two different (and separable) computational
293     tricks:
294    
295     \begin{enumerate}
296     \item shifting through the use of image charges, and
297     \item damping the electrostatic interaction.
298     \end{enumerate}
299     Wolf \textit{et al.} treated the
300     development of their summation method as a progressive application of
301     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
302     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
303     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
304     both techniques. It is possible, however, to separate these
305     tricks and study their effects independently.
306    
307     Starting with the original observation that the effective range of the
308     electrostatic interaction in condensed phases is considerably less
309     than $r^{-1}$, either the cutoff sphere neutralization or the
310     distance-dependent damping technique could be used as a foundation for
311     a new pairwise summation method. Wolf \textit{et al.} made the
312     observation that charge neutralization within the cutoff sphere plays
313     a significant role in energy convergence; therefore we will begin our
314     analysis with the various shifted forms that maintain this charge
315     neutralization. We can evaluate the methods of Wolf
316     \textit{et al.} and Zahn \textit{et al.} by considering the standard
317     shifted potential,
318     \begin{equation}
319     V_\textrm{SP}(r) = \begin{cases}
320     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
321     R_\textrm{c}
322     \end{cases},
323     \label{eq:shiftingPotForm}
324     \end{equation}
325     and shifted force,
326     \begin{equation}
327     V_\textrm{SF}(r) = \begin{cases}
328     v(r) - v_\textrm{c}
329     - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
330     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
331     \end{cases},
332     \label{eq:shiftingForm}
333     \end{equation}
334     functions where $v(r)$ is the unshifted form of the potential, and
335     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
336     that both the potential and the forces goes to zero at the cutoff
337     radius, while the Shifted Potential ({\sc sp}) form only ensures the
338     potential is smooth at the cutoff radius
339     ($R_\textrm{c}$).\cite{Allen87}
340    
341     The forces associated with the shifted potential are simply the forces
342     of the unshifted potential itself (when inside the cutoff sphere),
343     \begin{equation}
344     F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
345     \end{equation}
346     and are zero outside. Inside the cutoff sphere, the forces associated
347     with the shifted force form can be written,
348     \begin{equation}
349     F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
350     v(r)}{dr} \right)_{r=R_\textrm{c}}.
351     \end{equation}
352    
353     If the potential, $v(r)$, is taken to be the normal Coulomb potential,
354     \begin{equation}
355     v(r) = \frac{q_i q_j}{r},
356     \label{eq:Coulomb}
357     \end{equation}
358     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
359     al.}'s undamped prescription:
360     \begin{equation}
361     V_\textrm{SP}(r) =
362     q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
363     r\leqslant R_\textrm{c},
364     \label{eq:SPPot}
365     \end{equation}
366     with associated forces,
367     \begin{equation}
368     F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
369     \quad r\leqslant R_\textrm{c}.
370     \label{eq:SPForces}
371     \end{equation}
372     These forces are identical to the forces of the standard Coulomb
373     interaction, and cutting these off at $R_c$ was addressed by Wolf
374     \textit{et al.} as undesirable. They pointed out that the effect of
375     the image charges is neglected in the forces when this form is
376     used,\cite{Wolf99} thereby eliminating any benefit from the method in
377     molecular dynamics. Additionally, there is a discontinuity in the
378     forces at the cutoff radius which results in energy drift during MD
379     simulations.
380    
381     The shifted force ({\sc sf}) form using the normal Coulomb potential
382     will give,
383     \begin{equation}
384     V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
385     + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
386     \quad r\leqslant R_\textrm{c}.
387     \label{eq:SFPot}
388     \end{equation}
389     with associated forces,
390     \begin{equation}
391     F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
392     \quad r\leqslant R_\textrm{c}.
393     \label{eq:SFForces}
394     \end{equation}
395     This formulation has the benefits that there are no discontinuities at
396     the cutoff radius, while the neutralizing image charges are present in
397     both the energy and force expressions. It would be simple to add the
398     self-neutralizing term back when computing the total energy of the
399     system, thereby maintaining the agreement with the Madelung energies.
400     A side effect of this treatment is the alteration in the shape of the
401     potential that comes from the derivative term. Thus, a degree of
402     clarity about agreement with the empirical potential is lost in order
403     to gain functionality in dynamics simulations.
404    
405     Wolf \textit{et al.} originally discussed the energetics of the
406     shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
407     insufficient for accurate determination of the energy with reasonable
408     cutoff distances. The calculated Madelung energies fluctuated around
409     the expected value as the cutoff radius was increased, but the
410     oscillations converged toward the correct value.\cite{Wolf99} A
411     damping function was incorporated to accelerate the convergence; and
412     though alternative forms for the damping function could be
413     used,\cite{Jones56,Heyes81} the complimentary error function was
414     chosen to mirror the effective screening used in the Ewald summation.
415     Incorporating this error function damping into the simple Coulomb
416     potential,
417     \begin{equation}
418     v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
419     \label{eq:dampCoulomb}
420     \end{equation}
421     the shifted potential (eq. (\ref{eq:SPPot})) becomes
422     \begin{equation}
423     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
424     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
425     \quad r\leqslant R_\textrm{c},
426     \label{eq:DSPPot}
427     \end{equation}
428     with associated forces,
429     \begin{equation}
430     F_{\textrm{DSP}}(r) = q_iq_j
431     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
432     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
433     \quad r\leqslant R_\textrm{c}.
434     \label{eq:DSPForces}
435     \end{equation}
436     Again, this damped shifted potential suffers from a
437     force-discontinuity at the cutoff radius, and the image charges play
438     no role in the forces. To remedy these concerns, one may derive a
439     {\sc sf} variant by including the derivative term in
440     eq. (\ref{eq:shiftingForm}),
441     \begin{equation}
442     \begin{split}
443     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
444     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
445     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
446     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
447     + \frac{2\alpha}{\pi^{1/2}}
448     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
449     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
450     \quad r\leqslant R_\textrm{c}.
451     \label{eq:DSFPot}
452     \end{split}
453     \end{equation}
454     The derivative of the above potential will lead to the following forces,
455     \begin{equation}
456     \begin{split}
457     F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
458     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
459     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
460     &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
461     {R_{\textrm{c}}^2}
462     + \frac{2\alpha}{\pi^{1/2}}
463     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
464     \right)\Biggr{]}
465     \quad r\leqslant R_\textrm{c}.
466     \label{eq:DSFForces}
467     \end{split}
468     \end{equation}
469     If the damping parameter $(\alpha)$ is set to zero, the undamped case,
470     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
471     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
472    
473     This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
474     derived by Zahn \textit{et al.}; however, there are two important
475     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
476     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
477     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
478     in the Zahn potential, resulting in a potential discontinuity as
479     particles cross $R_\textrm{c}$. Second, the sign of the derivative
480     portion is different. The missing $v_\textrm{c}$ term would not
481     affect molecular dynamics simulations (although the computed energy
482     would be expected to have sudden jumps as particle distances crossed
483     $R_c$). The sign problem is a potential source of errors, however.
484     In fact, it introduces a discontinuity in the forces at the cutoff,
485     because the force function is shifted in the wrong direction and
486     doesn't cross zero at $R_\textrm{c}$.
487    
488     Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
489     electrostatic summation method in which the potential and forces are
490     continuous at the cutoff radius and which incorporates the damping
491     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
492     this paper, we will evaluate exactly how good these methods ({\sc sp},
493     {\sc sf}, damping) are at reproducing the correct electrostatic
494     summation performed by the Ewald sum.
495    
496    
497     \section{Evaluating Pairwise Summation Techniques}
498    
499     In classical molecular mechanics simulations, there are two primary
500     techniques utilized to obtain information about the system of
501     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
502     techniques utilize pairwise summations of interactions between
503     particle sites, but they use these summations in different ways.
504    
505     In MC, the potential energy difference between configurations dictates
506     the progression of MC sampling. Going back to the origins of this
507     method, the acceptance criterion for the canonical ensemble laid out
508     by Metropolis \textit{et al.} states that a subsequent configuration
509     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
510     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
511     Maintaining the correct $\Delta E$ when using an alternate method for
512     handling the long-range electrostatics will ensure proper sampling
513     from the ensemble.
514    
515     In MD, the derivative of the potential governs how the system will
516     progress in time. Consequently, the force and torque vectors on each
517     body in the system dictate how the system evolves. If the magnitude
518     and direction of these vectors are similar when using alternate
519     electrostatic summation techniques, the dynamics in the short term
520     will be indistinguishable. Because error in MD calculations is
521     cumulative, one should expect greater deviation at longer times,
522     although methods which have large differences in the force and torque
523     vectors will diverge from each other more rapidly.
524    
525     \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
526    
527     The pairwise summation techniques (outlined in section
528     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
529     studying the energy differences between conformations. We took the
530     {\sc spme}-computed energy difference between two conformations to be the
531     correct behavior. An ideal performance by an alternative method would
532     reproduce these energy differences exactly (even if the absolute
533     energies calculated by the methods are different). Since none of the
534     methods provide exact energy differences, we used linear least squares
535     regressions of energy gap data to evaluate how closely the methods
536     mimicked the Ewald energy gaps. Unitary results for both the
537     correlation (slope) and correlation coefficient for these regressions
538     indicate perfect agreement between the alternative method and {\sc spme}.
539     Sample correlation plots for two alternate methods are shown in
540     Fig. \ref{fig:linearFit}.
541    
542     \begin{figure}
543     \centering
544     \includegraphics[width = \linewidth]{./figures/dualLinear.pdf}
545     \caption{Example least squares regressions of the configuration energy
546     differences for SPC/E water systems. The upper plot shows a data set
547     with a poor correlation coefficient ($R^2$), while the lower plot
548     shows a data set with a good correlation coefficient.}
549     \label{fig:linearFit}
550     \end{figure}
551    
552     Each of the seven system types (detailed in section \ref{sec:RepSims})
553     were represented using 500 independent configurations. Thus, each of
554     the alternative (non-Ewald) electrostatic summation methods was
555     evaluated using an accumulated 873,250 configurational energy
556     differences.
557    
558     Results and discussion for the individual analysis of each of the
559 chrisfen 2927 system types appear in sections \ref{sec:IndividualResults}, while the
560 chrisfen 2918 cumulative results over all the investigated systems appear below in
561     sections \ref{sec:EnergyResults}.
562    
563     \subsection{Molecular Dynamics and the Force and Torque
564     Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
565     (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
566     comparing the force and torque vectors with those obtained using the
567     reference Ewald summation ({\sc spme}). Both the magnitude and the
568     direction of these vectors on each of the bodies in the system were
569     analyzed. For the magnitude of these vectors, linear least squares
570     regression analyses were performed as described previously for
571     comparing $\Delta E$ values. Instead of a single energy difference
572     between two system configurations, we compared the magnitudes of the
573     forces (and torques) on each molecule in each configuration. For a
574     system of 1000 water molecules and 40 ions, there are 1040 force
575     vectors and 1000 torque vectors. With 500 configurations, this
576     results in 520,000 force and 500,000 torque vector comparisons.
577     Additionally, data from seven different system types was aggregated
578     before the comparison was made.
579    
580     The {\it directionality} of the force and torque vectors was
581     investigated through measurement of the angle ($\theta$) formed
582     between those computed from the particular method and those from {\sc spme},
583     \begin{equation}
584     \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
585     \cdot \hat{F}_\textrm{M}\right),
586     \end{equation}
587     where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
588     vector computed using method M. Each of these $\theta$ values was
589     accumulated in a distribution function and weighted by the area on the
590     unit sphere. Since this distribution is a measure of angular error
591     between two different electrostatic summation methods, there is no
592     {\it a priori} reason for the profile to adhere to any specific
593     shape. Thus, gaussian fits were used to measure the width of the
594     resulting distributions. The variance ($\sigma^2$) was extracted from
595     each of these fits and was used to compare distribution widths.
596     Values of $\sigma^2$ near zero indicate vector directions
597     indistinguishable from those calculated when using the reference
598     method ({\sc spme}).
599    
600     \subsection{Short-time Dynamics}
601    
602     The effects of the alternative electrostatic summation methods on the
603     short-time dynamics of charged systems were evaluated by considering a
604     NaCl crystal at a temperature of 1000 K. A subset of the best
605     performing pairwise methods was used in this comparison. The NaCl
606     crystal was chosen to avoid possible complications from the treatment
607     of orientational motion in molecular systems. All systems were
608     started with the same initial positions and velocities. Simulations
609     were performed under the microcanonical ensemble, and velocity
610     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
611     of the trajectories,
612     \begin{equation}
613     C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
614     \label{eq:vCorr}
615     \end{equation}
616     Velocity autocorrelation functions require detailed short time data,
617     thus velocity information was saved every 2 fs over 10 ps
618     trajectories. Because the NaCl crystal is composed of two different
619     atom types, the average of the two resulting velocity autocorrelation
620     functions was used for comparisons.
621    
622     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
623    
624     The effects of the same subset of alternative electrostatic methods on
625     the {\it long-time} dynamics of charged systems were evaluated using
626     the same model system (NaCl crystals at 1000K). The power spectrum
627     ($I(\omega)$) was obtained via Fourier transform of the velocity
628     autocorrelation function, \begin{equation} I(\omega) =
629     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
630     \label{eq:powerSpec}
631     \end{equation}
632     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
633     NaCl crystal is composed of two different atom types, the average of
634     the two resulting power spectra was used for comparisons. Simulations
635     were performed under the microcanonical ensemble, and velocity
636     information was saved every 5~fs over 100~ps trajectories.
637    
638     \subsection{Representative Simulations}\label{sec:RepSims}
639     A variety of representative molecular simulations were analyzed to
640     determine the relative effectiveness of the pairwise summation
641     techniques in reproducing the energetics and dynamics exhibited by
642     {\sc spme}. We wanted to span the space of typical molecular
643     simulations (i.e. from liquids of neutral molecules to ionic
644     crystals), so the systems studied were:
645    
646     \begin{enumerate}
647     \item liquid water (SPC/E),\cite{Berendsen87}
648     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
649     \item NaCl crystals,
650     \item NaCl melts,
651     \item a low ionic strength solution of NaCl in water (0.11 M),
652     \item a high ionic strength solution of NaCl in water (1.1 M), and
653     \item a 6\AA\ radius sphere of Argon in water.
654     \end{enumerate}
655    
656     By utilizing the pairwise techniques (outlined in section
657     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
658     charged particles, and mixtures of the two, we hope to discern under
659     which conditions it will be possible to use one of the alternative
660     summation methodologies instead of the Ewald sum.
661    
662     For the solid and liquid water configurations, configurations were
663     taken at regular intervals from high temperature trajectories of 1000
664     SPC/E water molecules. Each configuration was equilibrated
665     independently at a lower temperature (300K for the liquid, 200K for
666     the crystal). The solid and liquid NaCl systems consisted of 500
667     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
668     these systems were selected and equilibrated in the same manner as the
669     water systems. In order to introduce measurable fluctuations in the
670     configuration energy differences, the crystalline simulations were
671     equilibrated at 1000K, near the $T_\textrm{m}$ for NaCl. The liquid
672     NaCl configurations needed to represent a fully disordered array of
673     point charges, so the high temperature of 7000K was selected for
674     equilibration. The ionic solutions were made by solvating 4 (or 40)
675     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
676     water positions were then randomly swapped, and the resulting
677     configurations were again equilibrated individually. Finally, for the
678     Argon / Water ``charge void'' systems, the identities of all the SPC/E
679     waters within 6\AA\ of the center of the equilibrated water
680     configurations were converted to argon.
681    
682     These procedures guaranteed us a set of representative configurations
683     from chemically-relevant systems sampled from appropriate
684     ensembles. Force field parameters for the ions and Argon were taken
685     from the force field utilized by {\sc oopse}.\cite{Meineke05}
686    
687     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
688     We compared the following alternative summation methods with results
689     from the reference method ({\sc spme}):
690    
691     \begin{enumerate}
692     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
693     and 0.3\AA$^{-1}$,
694     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
695     and 0.3\AA$^{-1}$,
696     \item reaction field with an infinite dielectric constant, and
697     \item an unmodified cutoff.
698     \end{enumerate}
699    
700     Group-based cutoffs with a fifth-order polynomial switching function
701     were utilized for the reaction field simulations. Additionally, we
702     investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
703     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
704     implementation of {\sc spme},\cite{Ponder87} while all other calculations
705     were performed using the {\sc oopse} molecular mechanics
706     package.\cite{Meineke05} All other portions of the energy calculation
707     (i.e. Lennard-Jones interactions) were handled in exactly the same
708     manner across all systems and configurations.
709    
710     The alternative methods were also evaluated with three different
711     cutoff radii (9, 12, and 15\AA). As noted previously, the
712     convergence parameter ($\alpha$) plays a role in the balance of the
713     real-space and reciprocal-space portions of the Ewald calculation.
714     Typical molecular mechanics packages set this to a value dependent on
715     the cutoff radius and a tolerance (typically less than $1 \times
716     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
717     increasing accuracy at the expense of computational time spent on the
718     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
719     The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
720     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
721     0.3119, and 0.2476\AA$^{-1}$ for cutoff radii of 9, 12, and 15\AA\
722     respectively.
723    
724 chrisfen 2927 \section{Configuration Energy Difference Results}\label{sec:EnergyResults}
725 chrisfen 2918 In order to evaluate the performance of the pairwise electrostatic
726 chrisfen 2920 summation methods for Monte Carlo (MC) simulations, the energy
727     differences between configurations were compared to the values
728     obtained when using {\sc spme}. The results for the combined
729     regression analysis of all of the systems are shown in figure
730     \ref{fig:delE}.
731 chrisfen 2918
732     \begin{figure}
733     \centering
734     \includegraphics[width=4.75in]{./figures/delEplot.pdf}
735     \caption{Statistical analysis of the quality of configurational energy
736     differences for a given electrostatic method compared with the
737     reference Ewald sum. Results with a value equal to 1 (dashed line)
738     indicate $\Delta E$ values indistinguishable from those obtained using
739     {\sc spme}. Different values of the cutoff radius are indicated with
740     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
741     inverted triangles).}
742     \label{fig:delE}
743     \end{figure}
744    
745     The most striking feature of this plot is how well the Shifted Force
746     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
747     differences. For the undamped {\sc sf} method, and the
748     moderately-damped {\sc sp} methods, the results are nearly
749     indistinguishable from the Ewald results. The other common methods do
750     significantly less well.
751    
752     The unmodified cutoff method is essentially unusable. This is not
753     surprising since hard cutoffs give large energy fluctuations as atoms
754     or molecules move in and out of the cutoff
755     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
756     some degree by using group based cutoffs with a switching
757     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
758     significant improvement using the group-switched cutoff because the
759     salt and salt solution systems contain non-neutral groups. Section
760 chrisfen 2927 \ref{sec:IndividualResults} includes results for systems comprised entirely
761 chrisfen 2918 of neutral groups.
762    
763     For the {\sc sp} method, inclusion of electrostatic damping improves
764     the agreement with Ewald, and using an $\alpha$ of 0.2\AA $^{-1}$
765     shows an excellent correlation and quality of fit with the {\sc spme}
766     results, particularly with a cutoff radius greater than 12
767     \AA . Use of a larger damping parameter is more helpful for the
768     shortest cutoff shown, but it has a detrimental effect on simulations
769     with larger cutoffs.
770    
771     In the {\sc sf} sets, increasing damping results in progressively {\it
772     worse} correlation with Ewald. Overall, the undamped case is the best
773     performing set, as the correlation and quality of fits are
774     consistently superior regardless of the cutoff distance. The undamped
775     case is also less computationally demanding (because no evaluation of
776     the complementary error function is required).
777    
778     The reaction field results illustrates some of that method's
779     limitations, primarily that it was developed for use in homogenous
780     systems; although it does provide results that are an improvement over
781     those from an unmodified cutoff.
782    
783 chrisfen 2927 \section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults}
784 chrisfen 2918
785     Evaluation of pairwise methods for use in Molecular Dynamics
786     simulations requires consideration of effects on the forces and
787     torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
788     regression results for the force and torque vector magnitudes,
789     respectively. The data in these figures was generated from an
790     accumulation of the statistics from all of the system types.
791    
792     \begin{figure}
793     \centering
794     \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
795     \caption{Statistical analysis of the quality of the force vector
796     magnitudes for a given electrostatic method compared with the
797     reference Ewald sum. Results with a value equal to 1 (dashed line)
798     indicate force magnitude values indistinguishable from those obtained
799     using {\sc spme}. Different values of the cutoff radius are indicated with
800     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
801     inverted triangles).}
802     \label{fig:frcMag}
803     \end{figure}
804    
805     Again, it is striking how well the Shifted Potential and Shifted Force
806     methods are doing at reproducing the {\sc spme} forces. The undamped and
807     weakly-damped {\sc sf} method gives the best agreement with Ewald.
808     This is perhaps expected because this method explicitly incorporates a
809     smooth transition in the forces at the cutoff radius as well as the
810     neutralizing image charges.
811    
812     Figure \ref{fig:frcMag}, for the most part, parallels the results seen
813     in the previous $\Delta E$ section. The unmodified cutoff results are
814     poor, but using group based cutoffs and a switching function provides
815     an improvement much more significant than what was seen with $\Delta
816     E$.
817    
818     With moderate damping and a large enough cutoff radius, the {\sc sp}
819     method is generating usable forces. Further increases in damping,
820     while beneficial for simulations with a cutoff radius of 9\AA\ , is
821     detrimental to simulations with larger cutoff radii.
822    
823     The reaction field results are surprisingly good, considering the poor
824     quality of the fits for the $\Delta E$ results. There is still a
825     considerable degree of scatter in the data, but the forces correlate
826     well with the Ewald forces in general. We note that the reaction
827     field calculations do not include the pure NaCl systems, so these
828     results are partly biased towards conditions in which the method
829     performs more favorably.
830    
831     \begin{figure}
832     \centering
833     \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
834     \caption{Statistical analysis of the quality of the torque vector
835     magnitudes for a given electrostatic method compared with the
836     reference Ewald sum. Results with a value equal to 1 (dashed line)
837     indicate torque magnitude values indistinguishable from those obtained
838     using {\sc spme}. Different values of the cutoff radius are indicated with
839     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
840     inverted triangles).}
841     \label{fig:trqMag}
842     \end{figure}
843    
844     Molecular torques were only available from the systems which contained
845     rigid molecules (i.e. the systems containing water). The data in
846     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
847    
848     Torques appear to be much more sensitive to charges at a longer
849     distance. The striking feature in comparing the new electrostatic
850     methods with {\sc spme} is how much the agreement improves with increasing
851     cutoff radius. Again, the weakly damped and undamped {\sc sf} method
852     appears to be reproducing the {\sc spme} torques most accurately.
853    
854     Water molecules are dipolar, and the reaction field method reproduces
855     the effect of the surrounding polarized medium on each of the
856     molecular bodies. Therefore it is not surprising that reaction field
857     performs best of all of the methods on molecular torques.
858    
859 chrisfen 2927 \section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults}
860 chrisfen 2918
861     It is clearly important that a new electrostatic method can reproduce
862     the magnitudes of the force and torque vectors obtained via the Ewald
863     sum. However, the {\it directionality} of these vectors will also be
864     vital in calculating dynamical quantities accurately. Force and
865     torque directionalities were investigated by measuring the angles
866     formed between these vectors and the same vectors calculated using
867     {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
868     variance ($\sigma^2$) of the Gaussian fits of the angle error
869     distributions of the combined set over all system types.
870    
871     \begin{figure}
872     \centering
873     \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
874     \caption{Statistical analysis of the width of the angular distribution
875     that the force and torque vectors from a given electrostatic method
876     make with their counterparts obtained using the reference Ewald sum.
877     Results with a variance ($\sigma^2$) equal to zero (dashed line)
878     indicate force and torque directions indistinguishable from those
879     obtained using {\sc spme}. Different values of the cutoff radius are
880     indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
881     and 15\AA\ = inverted triangles).}
882     \label{fig:frcTrqAng}
883     \end{figure}
884    
885     Both the force and torque $\sigma^2$ results from the analysis of the
886     total accumulated system data are tabulated in figure
887     \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
888     sp}) method would be essentially unusable for molecular dynamics
889     unless the damping function is added. The Shifted Force ({\sc sf})
890     method, however, is generating force and torque vectors which are
891     within a few degrees of the Ewald results even with weak (or no)
892     damping.
893    
894     All of the sets (aside from the over-damped case) show the improvement
895     afforded by choosing a larger cutoff radius. Increasing the cutoff
896     from 9 to 12\AA\ typically results in a halving of the width of the
897     distribution, with a similar improvement when going from 12 to 15
898     \AA .
899    
900     The undamped {\sc sf}, group-based cutoff, and reaction field methods
901     all do equivalently well at capturing the direction of both the force
902     and torque vectors. Using the electrostatic damping improves the
903     angular behavior significantly for the {\sc sp} and moderately for the
904     {\sc sf} methods. Overdamping is detrimental to both methods. Again
905     it is important to recognize that the force vectors cover all
906     particles in all seven systems, while torque vectors are only
907     available for neutral molecular groups. Damping is more beneficial to
908     charged bodies, and this observation is investigated further in
909 chrisfen 2927 section \ref{IndividualResults}.
910 chrisfen 2918
911     Although not discussed previously, group based cutoffs can be applied
912     to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
913     will reintroduce small discontinuities at the cutoff radius, but the
914     effects of these can be minimized by utilizing a switching function.
915     Though there are no significant benefits or drawbacks observed in
916     $\Delta E$ and the force and torque magnitudes when doing this, there
917     is a measurable improvement in the directionality of the forces and
918     torques. Table \ref{tab:groupAngle} shows the angular variances
919     obtained both without (N) and with (Y) group based cutoffs and a
920     switching function. Note that the $\alpha$ values have units of
921     \AA$^{-1}$ and the variance values have units of degrees$^2$. The
922     {\sc sp} (with an $\alpha$ of 0.2\AA$^{-1}$ or smaller) shows much
923     narrower angular distributions when using group-based cutoffs. The
924     {\sc sf} method likewise shows improvement in the undamped and lightly
925     damped cases.
926    
927     \begin{table}
928     \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
929     THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
930     GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
931     THE REFERENCE EWALD SUMMATION}
932    
933     \footnotesize
934     \begin{center}
935     \begin{tabular}{@{} ccrrrrrrrr @{}} \\
936     \toprule
937     \toprule
938    
939     & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
940     Force} \\
941     \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
942     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
943     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
944    
945     \midrule
946    
947     9\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
948     & \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} \\
949     12\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
950     & \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} \\
951     15\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
952     & \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} \\
953    
954     \midrule
955    
956     9\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
957     & \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} \\
958     12\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
959     & \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} \\
960     15\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
961     & \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} \\
962    
963     \bottomrule
964     \end{tabular}
965     \end{center}
966     \label{tab:groupAngle}
967     \end{table}
968    
969     One additional trend in table \ref{tab:groupAngle} is that the
970     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
971     increases, something that is more obvious with group-based cutoffs.
972     The complimentary error function inserted into the potential weakens
973     the electrostatic interaction as the value of $\alpha$ is increased.
974     However, at larger values of $\alpha$, it is possible to overdamp the
975     electrostatic interaction and to remove it completely. Kast
976     \textit{et al.} developed a method for choosing appropriate $\alpha$
977     values for these types of electrostatic summation methods by fitting
978     to $g(r)$ data, and their methods indicate optimal values of 0.34,
979     0.25, and 0.16\AA$^{-1}$ for cutoff values of 9, 12, and 15\AA\
980     respectively.\cite{Kast03} These appear to be reasonable choices to
981     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
982     these findings, choices this high would introduce error in the
983     molecular torques, particularly for the shorter cutoffs. Based on our
984     observations, empirical damping up to 0.2\AA$^{-1}$ is beneficial,
985     but damping may be unnecessary when using the {\sc sf} method.
986    
987 chrisfen 2927 \section{Individual System Analysis Results}\label{sec:IndividualResults}
988 chrisfen 2918
989 chrisfen 2920 The combined results of the previous sections show how the pairwise
990     methods compare to the Ewald summation in the general sense over all
991     of the system types. It is also useful to consider each of the
992     studied systems in an individual fashion, so that we can identify
993     conditions that are particularly difficult for a selected pairwise
994     method to address. This allows us to further establish the limitations
995     of these pairwise techniques. Below, the energy difference, force
996     vector, and torque vector analyses are presented on an individual
997     system basis.
998    
999 chrisfen 2927 \subsection{SPC/E Water Results}\label{sec:WaterResults}
1000 chrisfen 2920
1001 chrisfen 2927 The first system considered was liquid water at 300K using the SPC/E
1002     model of water.\cite{Berendsen87} The results for the energy gap
1003     comparisons and the force and torque vector magnitude comparisons are
1004     shown in table \ref{tab:spce}. The force and torque vector
1005     directionality results are displayed separately in table
1006     \ref{tab:spceAng}, where the effect of group-based cutoffs and
1007     switching functions on the {\sc sp} and {\sc sf} potentials are also
1008     investigated. In all of the individual results table, the method
1009     abbreviations are as follows:
1010 chrisfen 2920
1011 chrisfen 2927 \begin{itemize}
1012     \item PC = Pure Cutoff,
1013     \item SP = Shifted Potential,
1014     \item SF = Shifted Force,
1015     \item GSC = Group Switched Cutoff,
1016     \item RF = Reaction Field (where $\varepsilon \approx\infty$),
1017     \item GSSP = Group Switched Shifted Potential, and
1018     \item GSSF = Group Switched Shifted Force.
1019     \end{itemize}
1020 chrisfen 2920
1021 chrisfen 2927 \begin{table}[htbp]
1022     \centering
1023     \caption{REGRESSION RESULTS OF THE LIQUID WATER SYSTEM FOR THE
1024     $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it middle})
1025     AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1026 chrisfen 2920
1027 chrisfen 2927 \footnotesize
1028     \begin{tabular}{@{} ccrrrrrr @{}}
1029     \\
1030     \toprule
1031     \toprule
1032     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1033     \cmidrule(lr){3-4}
1034     \cmidrule(lr){5-6}
1035     \cmidrule(l){7-8}
1036     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1037     \midrule
1038     PC & & 3.046 & 0.002 & -3.018 & 0.002 & 4.719 & 0.005 \\
1039     SP & 0.0 & 1.035 & 0.218 & 0.908 & 0.313 & 1.037 & 0.470 \\
1040     & 0.1 & 1.021 & 0.387 & 0.965 & 0.752 & 1.006 & 0.947 \\
1041     & 0.2 & 0.997 & 0.962 & 1.001 & 0.994 & 0.994 & 0.996 \\
1042     & 0.3 & 0.984 & 0.980 & 0.997 & 0.985 & 0.982 & 0.987 \\
1043     SF & 0.0 & 0.977 & 0.974 & 0.996 & 0.992 & 0.991 & 0.997 \\
1044     & 0.1 & 0.983 & 0.974 & 1.001 & 0.994 & 0.996 & 0.998 \\
1045     & 0.2 & 0.992 & 0.989 & 1.001 & 0.995 & 0.994 & 0.996 \\
1046     & 0.3 & 0.984 & 0.980 & 0.996 & 0.985 & 0.982 & 0.987 \\
1047     GSC & & 0.918 & 0.862 & 0.852 & 0.756 & 0.801 & 0.700 \\
1048     RF & & 0.971 & 0.958 & 0.975 & 0.987 & 0.959 & 0.983 \\
1049     \midrule
1050     PC & & -1.647 & 0.000 & -0.127 & 0.000 & -0.979 & 0.000 \\
1051     SP & 0.0 & 0.735 & 0.368 & 0.813 & 0.537 & 0.865 & 0.659 \\
1052     & 0.1 & 0.850 & 0.612 & 0.956 & 0.887 & 0.992 & 0.979 \\
1053     & 0.2 & 0.996 & 0.989 & 1.000 & 1.000 & 1.000 & 1.000 \\
1054     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1055     SF & 0.0 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 0.999 \\
1056     & 0.1 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1057     & 0.2 & 0.999 & 0.998 & 1.000 & 1.000 & 1.000 & 1.000 \\
1058     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1059     GSC & & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1060     RF & & 0.999 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1061     \midrule
1062     PC & & 2.387 & 0.000 & 0.183 & 0.000 & 1.282 & 0.000 \\
1063     SP & 0.0 & 0.847 & 0.543 & 0.904 & 0.694 & 0.935 & 0.786 \\
1064     & 0.1 & 0.922 & 0.749 & 0.980 & 0.934 & 0.996 & 0.988 \\
1065     & 0.2 & 0.987 & 0.985 & 0.989 & 0.992 & 0.990 & 0.993 \\
1066     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1067     SF & 0.0 & 0.978 & 0.990 & 0.988 & 0.997 & 0.993 & 0.999 \\
1068     & 0.1 & 0.983 & 0.991 & 0.993 & 0.997 & 0.997 & 0.999 \\
1069     & 0.2 & 0.986 & 0.989 & 0.989 & 0.992 & 0.990 & 0.993 \\
1070     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1071     GSC & & 0.995 & 0.981 & 0.999 & 0.991 & 1.001 & 0.994 \\
1072     RF & & 0.993 & 0.989 & 0.998 & 0.996 & 1.000 & 0.999 \\
1073     \bottomrule
1074     \end{tabular}
1075     \label{tab:spce}
1076     \end{table}
1077 chrisfen 2920
1078 chrisfen 2927 \begin{table}[htbp]
1079     \centering
1080     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1081     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE LIQUID WATER
1082     SYSTEM}
1083 chrisfen 2920
1084 chrisfen 2927 \footnotesize
1085     \begin{tabular}{@{} ccrrrrrr @{}}
1086     \\
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. Overdamping 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 overdamping 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     \\
1184     \toprule
1185     \toprule
1186     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1187     \cmidrule(lr){3-4}
1188     \cmidrule(lr){5-6}
1189     \cmidrule(l){7-8}
1190     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1191     \midrule
1192     PC & & 19.897 & 0.047 & -29.214 & 0.048 & -3.771 & 0.001 \\
1193     SP & 0.0 & -0.014 & 0.000 & 2.135 & 0.347 & 0.457 & 0.045 \\
1194     & 0.1 & 0.321 & 0.017 & 1.490 & 0.584 & 0.886 & 0.796 \\
1195     & 0.2 & 0.896 & 0.872 & 1.011 & 0.998 & 0.997 & 0.999 \\
1196     & 0.3 & 0.983 & 0.997 & 0.992 & 0.997 & 0.991 & 0.997 \\
1197     SF & 0.0 & 0.943 & 0.979 & 1.048 & 0.978 & 0.995 & 0.999 \\
1198     & 0.1 & 0.948 & 0.979 & 1.044 & 0.983 & 1.000 & 0.999 \\
1199     & 0.2 & 0.982 & 0.997 & 0.969 & 0.960 & 0.997 & 0.999 \\
1200     & 0.3 & 0.985 & 0.997 & 0.961 & 0.961 & 0.991 & 0.997 \\
1201     GSC & & 0.983 & 0.985 & 0.966 & 0.994 & 1.003 & 0.999 \\
1202     RF & & 0.924 & 0.944 & 0.990 & 0.996 & 0.991 & 0.998 \\
1203     \midrule
1204     PC & & -4.375 & 0.000 & 6.781 & 0.000 & -3.369 & 0.000 \\
1205     SP & 0.0 & 0.515 & 0.164 & 0.856 & 0.426 & 0.743 & 0.478 \\
1206     & 0.1 & 0.696 & 0.405 & 0.977 & 0.817 & 0.974 & 0.964 \\
1207     & 0.2 & 0.981 & 0.980 & 1.001 & 1.000 & 1.000 & 1.000 \\
1208     & 0.3 & 0.996 & 0.998 & 0.997 & 0.999 & 0.997 & 0.999 \\
1209     SF & 0.0 & 0.991 & 0.995 & 1.003 & 0.998 & 0.999 & 1.000 \\
1210     & 0.1 & 0.992 & 0.995 & 1.003 & 0.998 & 1.000 & 1.000 \\
1211     & 0.2 & 0.998 & 0.998 & 0.981 & 0.962 & 1.000 & 1.000 \\
1212     & 0.3 & 0.996 & 0.998 & 0.976 & 0.957 & 0.997 & 0.999 \\
1213     GSC & & 0.997 & 0.996 & 0.998 & 0.999 & 1.000 & 1.000 \\
1214     RF & & 0.988 & 0.989 & 1.000 & 0.999 & 1.000 & 1.000 \\
1215     \midrule
1216     PC & & -6.367 & 0.000 & -3.552 & 0.000 & -3.447 & 0.000 \\
1217     SP & 0.0 & 0.643 & 0.409 & 0.833 & 0.607 & 0.961 & 0.805 \\
1218     & 0.1 & 0.791 & 0.683 & 0.957 & 0.914 & 1.000 & 0.989 \\
1219     & 0.2 & 0.974 & 0.991 & 0.993 & 0.998 & 0.993 & 0.998 \\
1220     & 0.3 & 0.976 & 0.992 & 0.977 & 0.992 & 0.977 & 0.992 \\
1221     SF & 0.0 & 0.979 & 0.997 & 0.992 & 0.999 & 0.994 & 1.000 \\
1222     & 0.1 & 0.984 & 0.997 & 0.996 & 0.999 & 0.998 & 1.000 \\
1223     & 0.2 & 0.991 & 0.997 & 0.974 & 0.958 & 0.993 & 0.998 \\
1224     & 0.3 & 0.977 & 0.992 & 0.956 & 0.948 & 0.977 & 0.992 \\
1225     GSC & & 0.999 & 0.997 & 0.996 & 0.999 & 1.002 & 1.000 \\
1226     RF & & 0.994 & 0.997 & 0.997 & 0.999 & 1.000 & 1.000 \\
1227     \bottomrule
1228     \end{tabular}
1229     \label{tab:ice}
1230     \end{table}
1231    
1232     \begin{table}[htbp]
1233     \centering
1234     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1235     OF THE FORCE AND TORQUE VECTORS IN THE ICE I$_\textrm{c}$ SYSTEM}
1236    
1237     \footnotesize
1238     \begin{tabular}{@{} ccrrrrrr @{}}
1239     \\
1240     \toprule
1241     \toprule
1242     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque
1243     $\sigma^2$} \\
1244     \cmidrule(lr){3-5}
1245     \cmidrule(l){6-8}
1246     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1247     \midrule
1248     PC & & 2128.921 & 603.197 & 715.579 & 329.056 & 221.397 & 81.042 \\
1249     SP & 0.0 & 1429.341 & 470.320 & 447.557 & 301.678 & 197.437 & 73.840 \\
1250     & 0.1 & 590.008 & 107.510 & 18.883 & 118.201 & 32.472 & 3.599 \\
1251     & 0.2 & 10.057 & 0.105 & 0.038 & 2.875 & 0.572 & 0.518 \\
1252     & 0.3 & 0.245 & 0.260 & 0.262 & 2.365 & 2.396 & 2.327 \\
1253     SF & 0.0 & 1.745 & 1.161 & 0.212 & 1.135 & 0.426 & 0.155 \\
1254     & 0.1 & 1.721 & 0.868 & 0.082 & 1.118 & 0.358 & 0.118 \\
1255     & 0.2 & 0.201 & 0.040 & 0.038 & 0.786 & 0.555 & 0.518 \\
1256     & 0.3 & 0.241 & 0.260 & 0.262 & 2.368 & 2.400 & 2.327 \\
1257     GSC & & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1258     RF & & 2.887 & 0.217 & 0.107 & 1.006 & 0.281 & 0.085 \\
1259     \midrule
1260     GSSP & 0.0 & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1261     & 0.1 & 1.341 & 0.123 & 0.037 & 0.835 & 0.234 & 0.085 \\
1262     & 0.2 & 0.558 & 0.040 & 0.037 & 0.823 & 0.557 & 0.519 \\
1263     & 0.3 & 0.250 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1264     GSSF & 0.0 & 2.124 & 0.132 & 0.069 & 0.919 & 0.263 & 0.099 \\
1265     & 0.1 & 2.165 & 0.101 & 0.035 & 0.895 & 0.244 & 0.096 \\
1266     & 0.2 & 0.706 & 0.040 & 0.037 & 0.870 & 0.559 & 0.519 \\
1267     & 0.3 & 0.251 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1268     \bottomrule
1269     \end{tabular}
1270     \label{tab:iceAng}
1271     \end{table}
1272    
1273     Highly ordered systems are a difficult test for the pairwise methods
1274     in that they lack the implicit periodicity of the Ewald summation. As
1275     expected, the energy gap agreement with {\sc spme} is reduced for the
1276     {\sc sp} and {\sc sf} methods with parameters that were ideal for the
1277     disordered liquid system. Moving to higher $R_\textrm{c}$ helps
1278     improve the agreement, though at an increase in computational cost.
1279     The dynamics of this crystalline system (both in magnitude and
1280     direction) are little affected. Both methods still reproduce the Ewald
1281     behavior with the same parameter recommendations from the previous
1282     section.
1283    
1284     It is also worth noting that {\sc rf} exhibits improved energy gap
1285     results over the liquid water system. One possible explanation is
1286     that the ice I$_\textrm{c}$ crystal is ordered such that the net
1287     dipole moment of the crystal is zero. With $\epsilon_\textrm{S} =
1288     \infty$, the reaction field incorporates this structural organization
1289     by actively enforcing a zeroed dipole moment within each cutoff
1290     sphere.
1291    
1292     \subsection{NaCl Melt Results}\label{sec:SaltMeltResults}
1293    
1294     A high temperature NaCl melt was tested to gauge the accuracy of the
1295     pairwise summation methods in a disordered system of charges. The
1296     results for the energy gap comparisons and the force vector magnitude
1297     comparisons are shown in table \ref{tab:melt}. The force vector
1298     directionality results are displayed separately in table
1299     \ref{tab:meltAng}.
1300    
1301     \begin{table}[htbp]
1302     \centering
1303     \caption{REGRESSION RESULTS OF THE MOLTEN SODIUM CHLORIDE SYSTEM FOR
1304     $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES ({\it
1305     lower})}
1306    
1307     \footnotesize
1308     \begin{tabular}{@{} ccrrrrrr @{}}
1309     \\
1310     \toprule
1311     \toprule
1312     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1313     \cmidrule(lr){3-4}
1314     \cmidrule(lr){5-6}
1315     \cmidrule(l){7-8}
1316     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1317     \midrule
1318     PC & & -0.008 & 0.000 & -0.049 & 0.005 & -0.136 & 0.020 \\
1319     SP & 0.0 & 0.928 & 0.996 & 0.931 & 0.998 & 0.950 & 0.999 \\
1320     & 0.1 & 0.977 & 0.998 & 0.998 & 1.000 & 0.997 & 1.000 \\
1321     & 0.2 & 0.960 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1322     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1323     SF & 0.0 & 0.996 & 1.000 & 0.995 & 1.000 & 0.997 & 1.000 \\
1324     & 0.1 & 1.021 & 1.000 & 1.024 & 1.000 & 1.007 & 1.000 \\
1325     & 0.2 & 0.966 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1326     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1327     \midrule
1328     PC & & 1.103 & 0.000 & 0.989 & 0.000 & 0.802 & 0.000 \\
1329     SP & 0.0 & 0.973 & 0.981 & 0.975 & 0.988 & 0.979 & 0.992 \\
1330     & 0.1 & 0.987 & 0.992 & 0.993 & 0.998 & 0.997 & 0.999 \\
1331     & 0.2 & 0.993 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1332     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1333     SF & 0.0 & 0.996 & 0.997 & 0.997 & 0.999 & 0.998 & 1.000 \\
1334     & 0.1 & 1.000 & 0.997 & 1.001 & 0.999 & 1.000 & 1.000 \\
1335     & 0.2 & 0.994 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1336     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1337     \bottomrule
1338     \end{tabular}
1339     \label{tab:melt}
1340     \end{table}
1341    
1342     \begin{table}[htbp]
1343     \centering
1344     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1345     OF THE FORCE VECTORS IN THE MOLTEN SODIUM CHLORIDE SYSTEM}
1346    
1347     \footnotesize
1348     \begin{tabular}{@{} ccrrrrrr @{}}
1349     \\
1350     \toprule
1351     \toprule
1352     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1353     \cmidrule(lr){3-5}
1354     \cmidrule(l){6-8}
1355     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1356     \midrule
1357     PC & & 13.294 & 8.035 & 5.366 \\
1358     SP & 0.0 & 13.316 & 8.037 & 5.385 \\
1359     & 0.1 & 5.705 & 1.391 & 0.360 \\
1360     & 0.2 & 2.415 & 7.534 & 13.927 \\
1361     & 0.3 & 23.769 & 67.306 & 57.252 \\
1362     SF & 0.0 & 1.693 & 0.603 & 0.256 \\
1363     & 0.1 & 1.687 & 0.653 & 0.272 \\
1364     & 0.2 & 2.598 & 7.523 & 13.930 \\
1365     & 0.3 & 23.734 & 67.305 & 57.252 \\
1366     \bottomrule
1367     \end{tabular}
1368     \label{tab:meltAng}
1369     \end{table}
1370    
1371     The molten NaCl system shows more sensitivity to the electrostatic
1372     damping than the water systems. The most noticeable point is that the
1373     undamped {\sc sf} method does very well at replicating the {\sc spme}
1374     configurational energy differences and forces. Light damping appears
1375     to minimally improve the dynamics, but this comes with a deterioration
1376     of the energy gap results. In contrast, this light damping improves
1377     the {\sc sp} energy gaps and forces. Moderate and heavy electrostatic
1378     damping reduce the agreement with {\sc spme} for both methods. From
1379     these observations, the undamped {\sc sf} method is the best choice
1380     for disordered systems of charges.
1381    
1382     \subsection{NaCl Crystal Results}\label{sec:SaltCrystalResults}
1383    
1384     Similar to the use of ice I$_\textrm{c}$ to investigate the role of
1385     order in molecular systems on the effectiveness of the pairwise
1386     methods, the 1000K NaCl crystal system was used to investigate the
1387     accuracy of the pairwise summation methods in an ordered system of
1388     charged particles. The results for the energy gap comparisons and the
1389     force vector magnitude comparisons are shown in table \ref{tab:salt}.
1390     The force vector directionality results are displayed separately in
1391     table \ref{tab:saltAng}.
1392    
1393     \begin{table}[htbp]
1394     \centering
1395     \caption{REGRESSION RESULTS OF THE CRYSTALLINE SODIUM CHLORIDE
1396     SYSTEM FOR $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES
1397     ({\it lower})}
1398    
1399     \footnotesize
1400     \begin{tabular}{@{} ccrrrrrr @{}}
1401     \\
1402     \toprule
1403     \toprule
1404     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1405     \cmidrule(lr){3-4}
1406     \cmidrule(lr){5-6}
1407     \cmidrule(l){7-8}
1408     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1409     \midrule
1410     PC & & -20.241 & 0.228 & -20.248 & 0.229 & -20.239 & 0.228 \\
1411     SP & 0.0 & 1.039 & 0.733 & 2.037 & 0.565 & 1.225 & 0.743 \\
1412     & 0.1 & 1.049 & 0.865 & 1.424 & 0.784 & 1.029 & 0.980 \\
1413     & 0.2 & 0.982 & 0.976 & 0.969 & 0.980 & 0.960 & 0.980 \\
1414     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.945 \\
1415     SF & 0.0 & 1.041 & 0.967 & 0.994 & 0.989 & 0.957 & 0.993 \\
1416     & 0.1 & 1.050 & 0.968 & 0.996 & 0.991 & 0.972 & 0.995 \\
1417     & 0.2 & 0.982 & 0.975 & 0.959 & 0.980 & 0.960 & 0.980 \\
1418     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.944 \\
1419     \midrule
1420     PC & & 0.795 & 0.000 & 0.792 & 0.000 & 0.793 & 0.000 \\
1421     SP & 0.0 & 0.916 & 0.829 & 1.086 & 0.791 & 1.010 & 0.936 \\
1422     & 0.1 & 0.958 & 0.917 & 1.049 & 0.943 & 1.001 & 0.995 \\
1423     & 0.2 & 0.981 & 0.981 & 0.982 & 0.984 & 0.981 & 0.984 \\
1424     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1425     SF & 0.0 & 1.002 & 0.983 & 0.997 & 0.994 & 0.991 & 0.997 \\
1426     & 0.1 & 1.003 & 0.984 & 0.996 & 0.995 & 0.993 & 0.997 \\
1427     & 0.2 & 0.983 & 0.980 & 0.981 & 0.984 & 0.981 & 0.984 \\
1428     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1429     \bottomrule
1430     \end{tabular}
1431     \label{tab:salt}
1432     \end{table}
1433    
1434     \begin{table}[htbp]
1435     \centering
1436     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1437     DISTRIBUTIONS OF THE FORCE VECTORS IN THE CRYSTALLINE SODIUM CHLORIDE
1438     SYSTEM}
1439    
1440     \footnotesize
1441     \begin{tabular}{@{} ccrrrrrr @{}}
1442     \\
1443     \toprule
1444     \toprule
1445     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1446     \cmidrule(lr){3-5}
1447     \cmidrule(l){6-8}
1448     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1449     \midrule
1450     PC & & 111.945 & 111.824 & 111.866 \\
1451     SP & 0.0 & 112.414 & 152.215 & 38.087 \\
1452     & 0.1 & 52.361 & 42.574 & 2.819 \\
1453     & 0.2 & 10.847 & 9.709 & 9.686 \\
1454     & 0.3 & 31.128 & 31.104 & 31.029 \\
1455     SF & 0.0 & 10.025 & 3.555 & 1.648 \\
1456     & 0.1 & 9.462 & 3.303 & 1.721 \\
1457     & 0.2 & 11.454 & 9.813 & 9.701 \\
1458     & 0.3 & 31.120 & 31.105 & 31.029 \\
1459     \bottomrule
1460     \end{tabular}
1461     \label{tab:saltAng}
1462     \end{table}
1463    
1464     The crystalline NaCl system is the most challenging test case for the
1465     pairwise summation methods, as evidenced by the results in tables
1466     \ref{tab:salt} and \ref{tab:saltAng}. The undamped and weakly damped
1467     {\sc sf} methods seem to be the best choices. These methods match well
1468     with {\sc spme} across the energy gap, force magnitude, and force
1469     directionality tests. The {\sc sp} method struggles in all cases,
1470     with the exception of good dynamics reproduction when using weak
1471     electrostatic damping with a large cutoff radius.
1472    
1473     The moderate electrostatic damping case is not as good as we would
1474     expect given the long-time dynamics results observed for this system
1475     (see section \ref{sec:LongTimeDynamics}). Since the data tabulated in
1476     tables \ref{tab:salt} and \ref{tab:saltAng} are a test of
1477     instantaneous dynamics, this indicates that good long-time dynamics
1478     comes in part at the expense of short-time dynamics.
1479    
1480     \subsection{0.11M NaCl Solution Results}
1481    
1482     In an effort to bridge the charged atomic and neutral molecular
1483     systems, Na$^+$ and Cl$^-$ ion charge defects were incorporated into
1484     the liquid water system. This low ionic strength system consists of 4
1485     ions in the 1000 SPC/E water solvent ($\approx$0.11 M). The results
1486     for the energy gap comparisons and the force and torque vector
1487     magnitude comparisons are shown in table \ref{tab:solnWeak}. The
1488     force and torque vector directionality results are displayed
1489     separately in table \ref{tab:solnWeakAng}, where the effect of
1490     group-based cutoffs and switching functions on the {\sc sp} and {\sc
1491     sf} potentials are investigated.
1492    
1493     \begin{table}[htbp]
1494     \centering
1495     \caption{REGRESSION RESULTS OF THE WEAK SODIUM CHLORIDE SOLUTION
1496     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1497     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1498    
1499     \footnotesize
1500     \begin{tabular}{@{} ccrrrrrr @{}}
1501     \\
1502     \toprule
1503     \toprule
1504     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1505     \cmidrule(lr){3-4}
1506     \cmidrule(lr){5-6}
1507     \cmidrule(l){7-8}
1508     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1509     \midrule
1510     PC & & 0.247 & 0.000 & -1.103 & 0.001 & 5.480 & 0.015 \\
1511     SP & 0.0 & 0.935 & 0.388 & 0.984 & 0.541 & 1.010 & 0.685 \\
1512     & 0.1 & 0.951 & 0.603 & 0.993 & 0.875 & 1.001 & 0.979 \\
1513     & 0.2 & 0.969 & 0.968 & 0.996 & 0.997 & 0.994 & 0.997 \\
1514     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1515     SF & 0.0 & 0.963 & 0.971 & 0.989 & 0.996 & 0.991 & 0.998 \\
1516     & 0.1 & 0.970 & 0.971 & 0.995 & 0.997 & 0.997 & 0.999 \\
1517     & 0.2 & 0.972 & 0.975 & 0.996 & 0.997 & 0.994 & 0.997 \\
1518     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1519     GSC & & 0.964 & 0.731 & 0.984 & 0.704 & 1.005 & 0.770 \\
1520     RF & & 0.968 & 0.605 & 0.974 & 0.541 & 1.014 & 0.614 \\
1521     \midrule
1522     PC & & 1.354 & 0.000 & -1.190 & 0.000 & -0.314 & 0.000 \\
1523     SP & 0.0 & 0.720 & 0.338 & 0.808 & 0.523 & 0.860 & 0.643 \\
1524     & 0.1 & 0.839 & 0.583 & 0.955 & 0.882 & 0.992 & 0.978 \\
1525     & 0.2 & 0.995 & 0.987 & 0.999 & 1.000 & 0.999 & 1.000 \\
1526     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1527     SF & 0.0 & 0.998 & 0.994 & 1.000 & 0.998 & 1.000 & 0.999 \\
1528     & 0.1 & 0.997 & 0.994 & 1.000 & 0.999 & 1.000 & 1.000 \\
1529     & 0.2 & 0.999 & 0.998 & 0.999 & 1.000 & 0.999 & 1.000 \\
1530     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1531     GSC & & 0.995 & 0.990 & 0.998 & 0.997 & 0.998 & 0.996 \\
1532     RF & & 0.998 & 0.993 & 0.999 & 0.998 & 0.999 & 0.996 \\
1533     \midrule
1534     PC & & 2.437 & 0.000 & -1.872 & 0.000 & 2.138 & 0.000 \\
1535     SP & 0.0 & 0.838 & 0.525 & 0.901 & 0.686 & 0.932 & 0.779 \\
1536     & 0.1 & 0.914 & 0.733 & 0.979 & 0.932 & 0.995 & 0.987 \\
1537     & 0.2 & 0.977 & 0.969 & 0.988 & 0.990 & 0.989 & 0.990 \\
1538     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1539     SF & 0.0 & 0.969 & 0.977 & 0.987 & 0.996 & 0.993 & 0.998 \\
1540     & 0.1 & 0.975 & 0.978 & 0.993 & 0.996 & 0.997 & 0.998 \\
1541     & 0.2 & 0.976 & 0.973 & 0.988 & 0.990 & 0.989 & 0.990 \\
1542     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1543     GSC & & 0.980 & 0.959 & 0.990 & 0.983 & 0.992 & 0.989 \\
1544     RF & & 0.984 & 0.975 & 0.996 & 0.995 & 0.998 & 0.998 \\
1545     \bottomrule
1546     \end{tabular}
1547     \label{tab:solnWeak}
1548     \end{table}
1549    
1550     \begin{table}[htbp]
1551     \centering
1552     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1553     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE WEAK SODIUM
1554     CHLORIDE SOLUTION SYSTEM}
1555    
1556     \footnotesize
1557     \begin{tabular}{@{} ccrrrrrr @{}}
1558     \\
1559     \toprule
1560     \toprule
1561     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1562     \cmidrule(lr){3-5}
1563     \cmidrule(l){6-8}
1564     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1565     \midrule
1566     PC & & 882.863 & 510.435 & 344.201 & 277.691 & 154.231 & 100.131 \\
1567     SP & 0.0 & 732.569 & 405.704 & 257.756 & 261.445 & 142.245 & 91.497 \\
1568     & 0.1 & 329.031 & 70.746 & 12.014 & 118.496 & 25.218 & 4.711 \\
1569     & 0.2 & 6.772 & 0.153 & 0.118 & 9.780 & 2.101 & 2.102 \\
1570     & 0.3 & 0.951 & 0.774 & 0.784 & 12.108 & 7.673 & 7.851 \\
1571     SF & 0.0 & 2.555 & 0.762 & 0.313 & 6.590 & 1.328 & 0.558 \\
1572     & 0.1 & 2.561 & 0.560 & 0.123 & 6.464 & 1.162 & 0.457 \\
1573     & 0.2 & 0.501 & 0.118 & 0.118 & 5.698 & 2.074 & 2.099 \\
1574     & 0.3 & 0.943 & 0.774 & 0.784 & 12.118 & 7.674 & 7.851 \\
1575     GSC & & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1576     RF & & 2.415 & 0.452 & 0.130 & 6.915 & 1.423 & 0.507 \\
1577     \midrule
1578     GSSP & 0.0 & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1579     & 0.1 & 2.251 & 0.324 & 0.064 & 7.628 & 1.639 & 0.497 \\
1580     & 0.2 & 0.590 & 0.118 & 0.116 & 6.080 & 2.096 & 2.103 \\
1581     & 0.3 & 0.953 & 0.759 & 0.780 & 12.347 & 7.683 & 7.849 \\
1582     GSSF & 0.0 & 1.541 & 0.301 & 0.096 & 6.407 & 1.316 & 0.496 \\
1583     & 0.1 & 1.541 & 0.237 & 0.050 & 6.356 & 1.202 & 0.457 \\
1584     & 0.2 & 0.568 & 0.118 & 0.116 & 6.166 & 2.105 & 2.105 \\
1585     & 0.3 & 0.954 & 0.759 & 0.780 & 12.337 & 7.684 & 7.849 \\
1586     \bottomrule
1587     \end{tabular}
1588     \label{tab:solnWeakAng}
1589     \end{table}
1590    
1591     Because this system is a perturbation of the pure liquid water system,
1592     comparisons are best drawn between these two sets. The {\sc sp} and
1593     {\sc sf} methods are not significantly affected by the inclusion of a
1594     few ions. The aspect of cutoff sphere neutralization aids in the
1595     smooth incorporation of these ions; thus, all of the observations
1596     regarding these methods carry over from section
1597     \ref{sec:WaterResults}. The differences between these systems are more
1598     visible for the {\sc rf} method. Though good force agreement is still
1599     maintained, the energy gaps show a significant increase in the scatter
1600     of the data.
1601    
1602     \subsection{1.1M NaCl Solution Results}
1603    
1604     The bridging of the charged atomic and neutral molecular systems was
1605     further developed by considering a high ionic strength system
1606     consisting of 40 ions in the 1000 SPC/E water solvent ($\approx$1.1
1607     M). The results for the energy gap comparisons and the force and
1608     torque vector magnitude comparisons are shown in table
1609     \ref{tab:solnStr}. The force and torque vector directionality
1610     results are displayed separately in table \ref{tab:solnStrAng}, where
1611     the effect of group-based cutoffs and switching functions on the {\sc
1612     sp} and {\sc sf} potentials are investigated.
1613    
1614     \begin{table}[htbp]
1615     \centering
1616     \caption{REGRESSION RESULTS OF THE STRONG SODIUM CHLORIDE SOLUTION
1617     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1618     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1619    
1620     \footnotesize
1621     \begin{tabular}{@{} ccrrrrrr @{}}
1622     \\
1623     \toprule
1624     \toprule
1625     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1626     \cmidrule(lr){3-4}
1627     \cmidrule(lr){5-6}
1628     \cmidrule(l){7-8}
1629     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1630     \midrule
1631     PC & & -0.081 & 0.000 & 0.945 & 0.001 & 0.073 & 0.000 \\
1632     SP & 0.0 & 0.978 & 0.469 & 0.996 & 0.672 & 0.975 & 0.668 \\
1633     & 0.1 & 0.944 & 0.645 & 0.997 & 0.886 & 0.991 & 0.978 \\
1634     & 0.2 & 0.873 & 0.896 & 0.985 & 0.993 & 0.980 & 0.993 \\
1635     & 0.3 & 0.831 & 0.860 & 0.960 & 0.979 & 0.955 & 0.977 \\
1636     SF & 0.0 & 0.858 & 0.905 & 0.985 & 0.970 & 0.990 & 0.998 \\
1637     & 0.1 & 0.865 & 0.907 & 0.992 & 0.974 & 0.994 & 0.999 \\
1638     & 0.2 & 0.862 & 0.894 & 0.985 & 0.993 & 0.980 & 0.993 \\
1639     & 0.3 & 0.831 & 0.859 & 0.960 & 0.979 & 0.955 & 0.977 \\
1640     GSC & & 1.985 & 0.152 & 0.760 & 0.031 & 1.106 & 0.062 \\
1641     RF & & 2.414 & 0.116 & 0.813 & 0.017 & 1.434 & 0.047 \\
1642     \midrule
1643     PC & & -7.028 & 0.000 & -9.364 & 0.000 & 0.925 & 0.865 \\
1644     SP & 0.0 & 0.701 & 0.319 & 0.909 & 0.773 & 0.861 & 0.665 \\
1645     & 0.1 & 0.824 & 0.565 & 0.970 & 0.930 & 0.990 & 0.979 \\
1646     & 0.2 & 0.988 & 0.981 & 0.995 & 0.998 & 0.991 & 0.998 \\
1647     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1648     SF & 0.0 & 0.993 & 0.988 & 0.992 & 0.984 & 0.998 & 0.999 \\
1649     & 0.1 & 0.993 & 0.989 & 0.993 & 0.986 & 0.998 & 1.000 \\
1650     & 0.2 & 0.993 & 0.992 & 0.995 & 0.998 & 0.991 & 0.998 \\
1651     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1652     GSC & & 0.964 & 0.897 & 0.970 & 0.917 & 0.925 & 0.865 \\
1653     RF & & 0.994 & 0.864 & 0.988 & 0.865 & 0.980 & 0.784 \\
1654     \midrule
1655     PC & & -2.212 & 0.000 & -0.588 & 0.000 & 0.953 & 0.925 \\
1656     SP & 0.0 & 0.800 & 0.479 & 0.930 & 0.804 & 0.924 & 0.759 \\
1657     & 0.1 & 0.883 & 0.694 & 0.976 & 0.942 & 0.993 & 0.986 \\
1658     & 0.2 & 0.952 & 0.943 & 0.980 & 0.984 & 0.980 & 0.983 \\
1659     & 0.3 & 0.914 & 0.909 & 0.943 & 0.948 & 0.944 & 0.946 \\
1660     SF & 0.0 & 0.945 & 0.953 & 0.980 & 0.984 & 0.991 & 0.998 \\
1661     & 0.1 & 0.951 & 0.954 & 0.987 & 0.986 & 0.995 & 0.998 \\
1662     & 0.2 & 0.951 & 0.946 & 0.980 & 0.984 & 0.980 & 0.983 \\
1663     & 0.3 & 0.914 & 0.908 & 0.943 & 0.948 & 0.944 & 0.946 \\
1664     GSC & & 0.882 & 0.818 & 0.939 & 0.902 & 0.953 & 0.925 \\
1665     RF & & 0.949 & 0.939 & 0.988 & 0.988 & 0.992 & 0.993 \\
1666     \bottomrule
1667     \end{tabular}
1668     \label{tab:solnStr}
1669     \end{table}
1670    
1671     \begin{table}[htbp]
1672     \centering
1673     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1674     OF THE FORCE AND TORQUE VECTORS IN THE STRONG SODIUM CHLORIDE SOLUTION
1675     SYSTEM}
1676    
1677     \footnotesize
1678     \begin{tabular}{@{} ccrrrrrr @{}}
1679     \\
1680     \toprule
1681     \toprule
1682     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1683     \cmidrule(lr){3-5}
1684     \cmidrule(l){6-8}
1685     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1686     \midrule
1687     PC & & 957.784 & 513.373 & 2.260 & 340.043 & 179.443 & 13.079 \\
1688     SP & 0.0 & 786.244 & 139.985 & 259.289 & 311.519 & 90.280 & 105.187 \\
1689     & 0.1 & 354.697 & 38.614 & 12.274 & 144.531 & 23.787 & 5.401 \\
1690     & 0.2 & 7.674 & 0.363 & 0.215 & 16.655 & 3.601 & 3.634 \\
1691     & 0.3 & 1.745 & 1.456 & 1.449 & 23.669 & 14.376 & 14.240 \\
1692     SF & 0.0 & 3.282 & 8.567 & 0.369 & 11.904 & 6.589 & 0.717 \\
1693     & 0.1 & 3.263 & 7.479 & 0.142 & 11.634 & 5.750 & 0.591 \\
1694     & 0.2 & 0.686 & 0.324 & 0.215 & 10.809 & 3.580 & 3.635 \\
1695     & 0.3 & 1.749 & 1.456 & 1.449 & 23.635 & 14.375 & 14.240 \\
1696     GSC & & 6.181 & 2.904 & 2.263 & 44.349 & 19.442 & 12.873 \\
1697     RF & & 3.891 & 0.847 & 0.323 & 18.628 & 3.995 & 2.072 \\
1698     \midrule
1699     GSSP & 0.0 & 6.197 & 2.929 & 2.290 & 44.441 & 19.442 & 12.873 \\
1700     & 0.1 & 4.688 & 1.064 & 0.260 & 31.208 & 6.967 & 2.303 \\
1701     & 0.2 & 1.021 & 0.218 & 0.213 & 14.425 & 3.629 & 3.649 \\
1702     & 0.3 & 1.752 & 1.454 & 1.451 & 23.540 & 14.390 & 14.245 \\
1703     GSSF & 0.0 & 2.494 & 0.546 & 0.217 & 16.391 & 3.230 & 1.613 \\
1704     & 0.1 & 2.448 & 0.429 & 0.106 & 16.390 & 2.827 & 1.159 \\
1705     & 0.2 & 0.899 & 0.214 & 0.213 & 13.542 & 3.583 & 3.645 \\
1706     & 0.3 & 1.752 & 1.454 & 1.451 & 23.587 & 14.390 & 14.245 \\
1707     \bottomrule
1708     \end{tabular}
1709     \label{tab:solnStrAng}
1710     \end{table}
1711    
1712     The {\sc rf} method struggles with the jump in ionic strength. The
1713     configuration energy differences degrade to unusable levels while the
1714     forces and torques show a more modest reduction in the agreement with
1715     {\sc spme}. The {\sc rf} method was designed for homogeneous systems,
1716     and this attribute is apparent in these results.
1717    
1718     The {\sc sp} and {\sc sf} methods require larger cutoffs to maintain
1719     their agreement with {\sc spme}. With these results, we still
1720     recommend undamped to moderate damping for the {\sc sf} method and
1721     moderate damping for the {\sc sp} method, both with cutoffs greater
1722     than 12\AA.
1723    
1724 chrisfen 2920 \subsection{6\AA\ Argon Sphere in SPC/E Water Results}
1725    
1726 chrisfen 2927 The final model system studied was a 6\AA\ sphere of Argon solvated
1727     by SPC/E water. This serves as a test case of a specifically sized
1728     electrostatic defect in a disordered molecular system. The results for
1729     the energy gap comparisons and the force and torque vector magnitude
1730     comparisons are shown in table \ref{tab:argon}. The force and torque
1731     vector directionality results are displayed separately in table
1732     \ref{tab:argonAng}, where the effect of group-based cutoffs and
1733     switching functions on the {\sc sp} and {\sc sf} potentials are
1734     investigated.
1735 chrisfen 2920
1736 chrisfen 2927 \begin{table}[htbp]
1737     \centering
1738     \caption{REGRESSION RESULTS OF THE 6\AA\ ARGON SPHERE IN LIQUID
1739     WATER SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR
1740     MAGNITUDES ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1741    
1742     \footnotesize
1743     \begin{tabular}{@{} ccrrrrrr @{}}
1744     \\
1745     \toprule
1746     \toprule
1747     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1748     \cmidrule(lr){3-4}
1749     \cmidrule(lr){5-6}
1750     \cmidrule(l){7-8}
1751     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1752     \midrule
1753     PC & & 2.320 & 0.008 & -0.650 & 0.001 & 3.848 & 0.029 \\
1754     SP & 0.0 & 1.053 & 0.711 & 0.977 & 0.820 & 0.974 & 0.882 \\
1755     & 0.1 & 1.032 & 0.846 & 0.989 & 0.965 & 0.992 & 0.994 \\
1756     & 0.2 & 0.993 & 0.995 & 0.982 & 0.998 & 0.986 & 0.998 \\
1757     & 0.3 & 0.968 & 0.995 & 0.954 & 0.992 & 0.961 & 0.994 \\
1758     SF & 0.0 & 0.982 & 0.996 & 0.992 & 0.999 & 0.993 & 1.000 \\
1759     & 0.1 & 0.987 & 0.996 & 0.996 & 0.999 & 0.997 & 1.000 \\
1760     & 0.2 & 0.989 & 0.998 & 0.984 & 0.998 & 0.989 & 0.998 \\
1761     & 0.3 & 0.971 & 0.995 & 0.957 & 0.992 & 0.965 & 0.994 \\
1762     GSC & & 1.002 & 0.983 & 0.992 & 0.973 & 0.996 & 0.971 \\
1763     RF & & 0.998 & 0.995 & 0.999 & 0.998 & 0.998 & 0.998 \\
1764     \midrule
1765     PC & & -36.559 & 0.002 & -44.917 & 0.004 & -52.945 & 0.006 \\
1766     SP & 0.0 & 0.890 & 0.786 & 0.927 & 0.867 & 0.949 & 0.909 \\
1767     & 0.1 & 0.942 & 0.895 & 0.984 & 0.974 & 0.997 & 0.995 \\
1768     & 0.2 & 0.999 & 0.997 & 1.000 & 1.000 & 1.000 & 1.000 \\
1769     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1770     SF & 0.0 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1771     & 0.1 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1772     & 0.2 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 \\
1773     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1774     GSC & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1775     RF & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1776     \midrule
1777     PC & & 1.984 & 0.000 & 0.012 & 0.000 & 1.357 & 0.000 \\
1778     SP & 0.0 & 0.850 & 0.552 & 0.907 & 0.703 & 0.938 & 0.793 \\
1779     & 0.1 & 0.924 & 0.755 & 0.980 & 0.936 & 0.995 & 0.988 \\
1780     & 0.2 & 0.985 & 0.983 & 0.986 & 0.988 & 0.987 & 0.988 \\
1781     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1782     SF & 0.0 & 0.977 & 0.989 & 0.987 & 0.995 & 0.992 & 0.998 \\
1783     & 0.1 & 0.982 & 0.989 & 0.992 & 0.996 & 0.997 & 0.998 \\
1784     & 0.2 & 0.984 & 0.987 & 0.986 & 0.987 & 0.987 & 0.988 \\
1785     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1786     GSC & & 0.995 & 0.981 & 0.999 & 0.990 & 1.000 & 0.993 \\
1787     RF & & 0.993 & 0.988 & 0.997 & 0.995 & 0.999 & 0.998 \\
1788     \bottomrule
1789     \end{tabular}
1790     \label{tab:argon}
1791     \end{table}
1792    
1793     \begin{table}[htbp]
1794     \centering
1795     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1796     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE 6\AA\ SPHERE OF
1797     ARGON IN LIQUID WATER SYSTEM}
1798    
1799     \footnotesize
1800     \begin{tabular}{@{} ccrrrrrr @{}}
1801     \\
1802     \toprule
1803     \toprule
1804     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1805     \cmidrule(lr){3-5}
1806     \cmidrule(l){6-8}
1807     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1808     \midrule
1809     PC & & 568.025 & 265.993 & 195.099 & 246.626 & 138.600 & 91.654 \\
1810     SP & 0.0 & 504.578 & 251.694 & 179.932 & 231.568 & 131.444 & 85.119 \\
1811     & 0.1 & 224.886 & 49.746 & 9.346 & 104.482 & 23.683 & 4.480 \\
1812     & 0.2 & 4.889 & 0.197 & 0.155 & 6.029 & 2.507 & 2.269 \\
1813     & 0.3 & 0.817 & 0.833 & 0.812 & 8.286 & 8.436 & 8.135 \\
1814     SF & 0.0 & 1.924 & 0.675 & 0.304 & 3.658 & 1.448 & 0.600 \\
1815     & 0.1 & 1.937 & 0.515 & 0.143 & 3.565 & 1.308 & 0.546 \\
1816     & 0.2 & 0.407 & 0.166 & 0.156 & 3.086 & 2.501 & 2.274 \\
1817     & 0.3 & 0.815 & 0.833 & 0.812 & 8.330 & 8.437 & 8.135 \\
1818     GSC & & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1819     RF & & 1.822 & 0.408 & 0.142 & 3.799 & 1.362 & 0.550 \\
1820     \midrule
1821     GSSP & 0.0 & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1822     & 0.1 & 1.652 & 0.309 & 0.087 & 4.197 & 1.401 & 0.590 \\
1823     & 0.2 & 0.465 & 0.165 & 0.153 & 3.323 & 2.529 & 2.273 \\
1824     & 0.3 & 0.813 & 0.825 & 0.816 & 8.316 & 8.447 & 8.132 \\
1825     GSSF & 0.0 & 1.173 & 0.292 & 0.113 & 3.452 & 1.347 & 0.583 \\
1826     & 0.1 & 1.166 & 0.240 & 0.076 & 3.381 & 1.281 & 0.575 \\
1827     & 0.2 & 0.459 & 0.165 & 0.153 & 3.430 & 2.542 & 2.273 \\
1828     & 0.3 & 0.814 & 0.825 & 0.816 & 8.325 & 8.447 & 8.132 \\
1829     \bottomrule
1830     \end{tabular}
1831     \label{tab:argonAng}
1832     \end{table}
1833    
1834     This system does not appear to show any significant deviations from
1835     the previously observed results. The {\sc sp} and {\sc sf} methods
1836     have aggrements similar to those observed in section
1837     \ref{sec:WaterResults}. The only significant difference is the
1838     improvement in the configuration energy differences for the {\sc rf}
1839     method. This is surprising in that we are introducing an inhomogeneity
1840     to the system; however, this inhomogeneity is charge-neutral and does
1841     not result in charged cutoff spheres. The charge-neutrality of the
1842     cutoff spheres, which the {\sc sp} and {\sc sf} methods explicitly
1843     enforce, seems to play a greater role in the stability of the {\sc rf}
1844     method than the required homogeneity of the environment.
1845    
1846    
1847     \section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics}
1848    
1849 chrisfen 2918 Zahn {\it et al.} investigated the structure and dynamics of water
1850     using eqs. (\ref{eq:ZahnPot}) and
1851     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
1852     that a method similar (but not identical with) the damped {\sc sf}
1853     method resulted in properties very similar to those obtained when
1854     using the Ewald summation. The properties they studied (pair
1855     distribution functions, diffusion constants, and velocity and
1856     orientational correlation functions) may not be particularly sensitive
1857     to the long-range and collective behavior that governs the
1858     low-frequency behavior in crystalline systems. Additionally, the
1859     ionic crystals are the worst case scenario for the pairwise methods
1860     because they lack the reciprocal space contribution contained in the
1861     Ewald summation.
1862    
1863     We are using two separate measures to probe the effects of these
1864     alternative electrostatic methods on the dynamics in crystalline
1865     materials. For short- and intermediate-time dynamics, we are
1866     computing the velocity autocorrelation function, and for long-time
1867     and large length-scale collective motions, we are looking at the
1868     low-frequency portion of the power spectrum.
1869    
1870     \begin{figure}
1871     \centering
1872     \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
1873     \caption{Velocity autocorrelation functions of NaCl crystals at
1874 chrisfen 2927 1000K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \&
1875     0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2\AA$^{-1}$). The inset is
1876     a magnification of the area around the first minimum. The times to
1877     first collision are nearly identical, but differences can be seen in
1878     the peaks and troughs, where the undamped and weakly damped methods
1879     are stiffer than the moderately damped and {\sc spme} methods.}
1880 chrisfen 2918 \label{fig:vCorrPlot}
1881     \end{figure}
1882    
1883     The short-time decay of the velocity autocorrelation function through
1884     the first collision are nearly identical in figure
1885     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1886     how the methods differ. The undamped {\sc sf} method has deeper
1887     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1888     any of the other methods. As the damping parameter ($\alpha$) is
1889     increased, these peaks are smoothed out, and the {\sc sf} method
1890     approaches the {\sc spme} results. With $\alpha$ values of 0.2\AA$^{-1}$,
1891     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1892     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1893     and {\sc sp} potentials become nearly identical with increased
1894     damping. However, this appears to indicate that once damping is
1895     utilized, the details of the form of the potential (and forces)
1896     constructed out of the damped electrostatic interaction are less
1897     important.
1898    
1899 chrisfen 2927 \section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics}
1900 chrisfen 2918
1901     To evaluate how the differences between the methods affect the
1902     collective long-time motion, we computed power spectra from long-time
1903     traces of the velocity autocorrelation function. The power spectra for
1904     the best-performing alternative methods are shown in
1905     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1906     a cubic switching function between 40 and 50 ps was used to reduce the
1907     ringing resulting from data truncation. This procedure had no
1908     noticeable effect on peak location or magnitude.
1909    
1910     \begin{figure}
1911     \centering
1912     \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
1913     \caption{Power spectra obtained from the velocity auto-correlation
1914     functions of NaCl crystals at 1000K while using {\sc spme}, {\sc sf}
1915 chrisfen 2927 ($\alpha$ = 0, 0.1, \& 0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ =
1916     0.2\AA$^{-1}$). The inset shows the frequency region below 100
1917     cm$^{-1}$ to highlight where the spectra differ.}
1918 chrisfen 2918 \label{fig:methodPS}
1919     \end{figure}
1920    
1921     While the high frequency regions of the power spectra for the
1922     alternative methods are quantitatively identical with Ewald spectrum,
1923     the low frequency region shows how the summation methods differ.
1924     Considering the low-frequency inset (expanded in the upper frame of
1925     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1926     correlated motions are blue-shifted when using undamped or weakly
1927     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1928     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1929     correlated motion to the Ewald method (which has a convergence
1930     parameter of 0.3119\AA$^{-1}$). This weakening of the electrostatic
1931     interaction with increased damping explains why the long-ranged
1932     correlated motions are at lower frequencies for the moderately damped
1933     methods than for undamped or weakly damped methods.
1934    
1935     To isolate the role of the damping constant, we have computed the
1936     spectra for a single method ({\sc sf}) with a range of damping
1937     constants and compared this with the {\sc spme} spectrum.
1938     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1939     electrostatic damping red-shifts the lowest frequency phonon modes.
1940     However, even without any electrostatic damping, the {\sc sf} method
1941     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1942     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1943     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1944     Most} of the collective behavior in the crystal is accurately captured
1945     using the {\sc sf} method. Quantitative agreement with Ewald can be
1946     obtained using moderate damping in addition to the shifting at the
1947     cutoff distance.
1948    
1949     \begin{figure}
1950     \centering
1951     \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1952     \caption{Effect of damping on the two lowest-frequency phonon modes in
1953     the NaCl crystal at 1000K. The undamped shifted force ({\sc sf})
1954     method is off by less than 10 cm$^{-1}$, and increasing the
1955     electrostatic damping to 0.25\AA$^{-1}$ gives quantitative agreement
1956     with the power spectrum obtained using the Ewald sum. Overdamping can
1957     result in underestimates of frequencies of the long-wavelength
1958     motions.}
1959     \label{fig:dampInc}
1960     \end{figure}
1961    
1962 chrisfen 2927 \section{Synopsis of the Pairwise Method Evaluation}\label{sec:PairwiseSynopsis}
1963 chrisfen 2918
1964 chrisfen 2927 The above investigation of pairwise electrostatic summation techniques
1965     shows that there are viable and computationally efficient alternatives
1966     to the Ewald summation. These methods are derived from the damped and
1967     cutoff-neutralized Coulombic sum originally proposed by Wolf
1968     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1969     method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1970     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1971     energetic and dynamic characteristics exhibited by simulations
1972     employing lattice summation techniques. The cumulative energy
1973     difference results showed the undamped {\sc sf} and moderately damped
1974     {\sc sp} methods produced results nearly identical to {\sc spme}.
1975     Similarly for the dynamic features, the undamped or moderately damped
1976     {\sc sf} and moderately damped {\sc sp} methods produce force and
1977     torque vector magnitude and directions very similar to the expected
1978     values. These results translate into long-time dynamic behavior
1979     equivalent to that produced in simulations using {\sc spme}.
1980    
1981     As in all purely-pairwise cutoff methods, these methods are expected
1982     to scale approximately {\it linearly} with system size, and they are
1983     easily parallelizable. This should result in substantial reductions
1984     in the computational cost of performing large simulations.
1985    
1986     Aside from the computational cost benefit, these techniques have
1987     applicability in situations where the use of the Ewald sum can prove
1988     problematic. Of greatest interest is their potential use in
1989     interfacial systems, where the unmodified lattice sum techniques
1990     artificially accentuate the periodicity of the system in an
1991     undesirable manner. There have been alterations to the standard Ewald
1992     techniques, via corrections and reformulations, to compensate for
1993     these systems; but the pairwise techniques discussed here require no
1994     modifications, making them natural tools to tackle these problems.
1995     Additionally, this transferability gives them benefits over other
1996     pairwise methods, like reaction field, because estimations of physical
1997     properties (e.g. the dielectric constant) are unnecessary.
1998    
1999     If a researcher is using Monte Carlo simulations of large chemical
2000     systems containing point charges, most structural features will be
2001     accurately captured using the undamped {\sc sf} method or the {\sc sp}
2002     method with an electrostatic damping of 0.2\AA$^{-1}$. These methods
2003     would also be appropriate for molecular dynamics simulations where the
2004     data of interest is either structural or short-time dynamical
2005     quantities. For long-time dynamics and collective motions, the safest
2006     pairwise method we have evaluated is the {\sc sf} method with an
2007     electrostatic damping between 0.2 and 0.25\AA$^{-1}$.
2008    
2009     We are not suggesting that there is any flaw with the Ewald sum; in
2010     fact, it is the standard by which these simple pairwise sums have been
2011     judged. However, these results do suggest that in the typical
2012     simulations performed today, the Ewald summation may no longer be
2013     required to obtain the level of accuracy most researchers have come to
2014     expect.
2015    
2016     \section{An Application: TIP5P-E Water}
2017    
2018    
2019 chrisfen 2918 \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
2020    
2021     \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}
2022    
2023     \chapter{\label{chap:shapes}SPHERICAL HARMONIC APPROXIMATIONS FOR MOLECULAR
2024     SIMULATIONS}
2025    
2026     \chapter{\label{chap:conclusion}CONCLUSION}
2027    
2028     \backmatter
2029    
2030     \bibliographystyle{ndthesis}
2031 chrisfen 2927 \bibliography{dissertation}
2032 chrisfen 2918
2033     \end{document}
2034    
2035    
2036     \endinput