ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/dissertation.tex
Revision: 2968
Committed: Sat Jul 29 13:10:15 2006 UTC (18 years, 1 month ago) by chrisfen
Content type: application/x-tex
File size: 106582 byte(s)
Log Message:
some text adjustments

File Contents

# User Rev Content
1 chrisfen 2957 \documentclass[11pt]{ndthesis}
2 chrisfen 2918
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 chrisfen 2957 \usepackage{cite}
11 chrisfen 2918
12     \begin{document}
13    
14     \frontmatter
15    
16     \title{APPLICATION AND DEVELOPMENT OF MOLECULAR DYNAMICS TECHNIQUES FOR THE
17     STUDY OF WATER}
18     \author{Christopher Joseph Fennell}
19     \work{Dissertation}
20     \degprior{B.Sc.}
21     \degaward{Doctor of Philosophy}
22     \advisor{J. Daniel Gezelter}
23     \department{Chemistry and Biochemistry}
24    
25     \maketitle
26    
27     \begin{abstract}
28     \end{abstract}
29    
30     \begin{dedication}
31     \end{dedication}
32    
33     \tableofcontents
34     \listoffigures
35     \listoftables
36    
37     \begin{acknowledge}
38     \end{acknowledge}
39    
40     \mainmatter
41    
42     \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
43    
44     \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION
45     TECHNIQUES}
46    
47     In molecular simulations, proper accumulation of the electrostatic
48     interactions is essential and is one of the most
49     computationally-demanding tasks. The common molecular mechanics force
50     fields represent atomic sites with full or partial charges protected
51     by Lennard-Jones (short range) interactions. This means that nearly
52     every pair interaction involves a calculation of charge-charge forces.
53     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
54     interactions quickly become the most expensive part of molecular
55     simulations. Historically, the electrostatic pair interaction would
56     not have decayed appreciably within the typical box lengths that could
57     be feasibly simulated. In the larger systems that are more typical of
58     modern simulations, large cutoffs should be used to incorporate
59     electrostatics correctly.
60    
61     There have been many efforts to address the proper and practical
62     handling of electrostatic interactions, and these have resulted in a
63     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
64     typically classified as implicit methods (i.e., continuum dielectrics,
65     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
66     (i.e., Ewald summations, interaction shifting or
67     truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
68     reaction field type methods, fast multipole
69     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
70     often preferred because they physically incorporate solvent molecules
71     in the system of interest, but these methods are sometimes difficult
72     to utilize because of their high computational cost.\cite{Roux99} In
73     addition to the computational cost, there have been some questions
74     regarding possible artifacts caused by the inherent periodicity of the
75     explicit Ewald summation.\cite{Tobias01}
76    
77     In this chapter, we focus on a new set of pairwise methods devised by
78     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
79     methods along with a few other mixed methods (i.e. reaction field) are
80     compared with the smooth particle mesh Ewald
81     sum,\cite{Onsager36,Essmann99} which is our reference method for
82     handling long-range electrostatic interactions. The new methods for
83     handling electrostatics have the potential to scale linearly with
84     increasing system size since they involve only a simple modification
85     to the direct pairwise sum. They also lack the added periodicity of
86     the Ewald sum, so they can be used for systems which are non-periodic
87     or which have one- or two-dimensional periodicity. Below, these
88     methods are evaluated using a variety of model systems to
89     establish their usability in molecular simulations.
90    
91     \section{The Ewald Sum}
92    
93     The complete accumulation of the electrostatic interactions in a system with
94     periodic boundary conditions (PBC) requires the consideration of the
95     effect of all charges within a (cubic) simulation box as well as those
96     in the periodic replicas,
97     \begin{equation}
98     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
99     \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
100     \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
101     \right],
102     \label{eq:PBCSum}
103     \end{equation}
104     where the sum over $\mathbf{n}$ is a sum over all periodic box
105     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
106     prime indicates $i = j$ are neglected for $\mathbf{n} =
107     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
108     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
109     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
110     $j$, and $\phi$ is the solution to Poisson's equation
111     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
112     charge-charge interactions). In the case of monopole electrostatics,
113     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
114     non-neutral systems.
115    
116     The electrostatic summation problem was originally studied by Ewald
117     for the case of an infinite crystal.\cite{Ewald21}. The approach he
118     took was to convert this conditionally convergent sum into two
119     absolutely convergent summations: a short-ranged real-space summation
120     and a long-ranged reciprocal-space summation,
121     \begin{equation}
122     \begin{split}
123     V_\textrm{elec} = \frac{1}{2}&
124     \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
125     \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
126     {|\mathbf{r}_{ij}+\mathbf{n}|} \\
127     &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
128     \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
129     \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
130     &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
131     + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
132     \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
133     \end{split}
134     \label{eq:EwaldSum}
135     \end{equation}
136     where $\alpha$ is the damping or convergence parameter with units of
137     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
138     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
139     constant of the surrounding medium. The final two terms of
140     eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
141     for interacting with a surrounding dielectric.\cite{Allen87} This
142     dipolar term was neglected in early applications in molecular
143     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
144     Leeuw {\it et al.} to address situations where the unit cell has a
145     dipole moment which is magnified through replication of the periodic
146     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
147     system is said to be using conducting (or ``tin-foil'') boundary
148     conditions, $\epsilon_{\rm S} = \infty$. Figure
149     \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
150     time. Initially, due to the small system sizes that could be
151     simulated feasibly, the entire simulation box was replicated to
152     convergence. In more modern simulations, the systems have grown large
153     enough that a real-space cutoff could potentially give convergent
154     behavior. Indeed, it has been observed that with the choice of a
155     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
156     rapidly convergent and small relative to the real-space
157     portion.\cite{Karasawa89,Kolafa92}
158    
159     \begin{figure}
160     \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
161     \caption{The change in the need for the Ewald sum with
162     increasing computational power. A:~Initially, only small systems
163     could be studied, and the Ewald sum replicated the simulation box to
164     convergence. B:~Now, radial cutoff methods should be able to reach
165     convergence for the larger systems of charges that are common today.}
166     \label{fig:ewaldTime}
167     \end{figure}
168    
169     The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
170     convergence parameter $(\alpha)$ plays an important role in balancing
171     the computational cost between the direct and reciprocal-space
172     portions of the summation. The choice of this value allows one to
173     select whether the real-space or reciprocal space portion of the
174     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
175     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
176     $\alpha$ and thoughtful algorithm development, this cost can be
177     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
178     taken to reduce the cost of the Ewald summation even further is to set
179     $\alpha$ such that the real-space interactions decay rapidly, allowing
180     for a short spherical cutoff. Then the reciprocal space summation is
181     optimized. These optimizations usually involve utilization of the
182     fast Fourier transform (FFT),\cite{Hockney81} leading to the
183     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
184     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
185     methods, the cost of the reciprocal-space portion of the Ewald
186     summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
187     \log N)$.
188    
189     These developments and optimizations have made the use of the Ewald
190     summation routine in simulations with periodic boundary
191     conditions. However, in certain systems, such as vapor-liquid
192     interfaces and membranes, the intrinsic three-dimensional periodicity
193     can prove problematic. The Ewald sum has been reformulated to handle
194     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
195     methods are computationally expensive.\cite{Spohr97,Yeh99} More
196     recently, there have been several successful efforts toward reducing
197     the computational cost of 2-D lattice
198     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
199     bringing them more in line with the cost of the full 3-D summation.
200    
201     Several studies have recognized that the inherent periodicity in the
202     Ewald sum can also have an effect on three-dimensional
203     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
204     Solvated proteins are essentially kept at high concentration due to
205     the periodicity of the electrostatic summation method. In these
206     systems, the more compact folded states of a protein can be
207     artificially stabilized by the periodic replicas introduced by the
208     Ewald summation.\cite{Weber00} Thus, care must be taken when
209     considering the use of the Ewald summation where the assumed
210     periodicity would introduce spurious effects in the system dynamics.
211    
212    
213     \section{The Wolf and Zahn Methods}
214    
215     In a recent paper by Wolf \textit{et al.}, a procedure was outlined
216     for the accurate accumulation of electrostatic interactions in an
217     efficient pairwise fashion. This procedure lacks the inherent
218     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
219     observed that the electrostatic interaction is effectively
220     short-ranged in condensed phase systems and that neutralization of the
221     charge contained within the cutoff radius is crucial for potential
222     stability. They devised a pairwise summation method that ensures
223     charge neutrality and gives results similar to those obtained with the
224     Ewald summation. The resulting shifted Coulomb potential includes
225     image-charges subtracted out through placement on the cutoff sphere
226     and a distance-dependent damping function (identical to that seen in
227     the real-space portion of the Ewald sum) to aid convergence
228     \begin{equation}
229     V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
230     - \lim_{r_{ij}\rightarrow R_\textrm{c}}
231     \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
232     \label{eq:WolfPot}
233     \end{equation}
234     Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
235     potential. However, neutralizing the charge contained within each
236     cutoff sphere requires the placement of a self-image charge on the
237     surface of the cutoff sphere. This additional self-term in the total
238     potential enabled Wolf {\it et al.} to obtain excellent estimates of
239     Madelung energies for many crystals.
240    
241     In order to use their charge-neutralized potential in molecular
242     dynamics simulations, Wolf \textit{et al.} suggested taking the
243     derivative of this potential prior to evaluation of the limit. This
244     procedure gives an expression for the forces,
245     \begin{equation}
246     \begin{split}
247     F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
248     \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
249     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
250     \Biggr]\\
251     &-\Biggr[
252     \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
253     + \frac{2\alpha}{\pi^{1/2}}
254     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
255     \Biggr]\Biggr\},
256     \end{split}
257     \label{eq:WolfForces}
258     \end{equation}
259     that incorporates both image charges and damping of the electrostatic
260     interaction.
261    
262     More recently, Zahn \textit{et al.} investigated these potential and
263     force expressions for use in simulations involving water.\cite{Zahn02}
264     In their work, they pointed out that the forces and derivative of
265     the potential are not commensurate. Attempts to use both
266     eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
267     to poor energy conservation. They correctly observed that taking the
268     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
269     derivatives gives forces for a different potential energy function
270     than the one shown in eq. (\ref{eq:WolfPot}).
271    
272     Zahn \textit{et al.} introduced a modified form of this summation
273     method as a way to use the technique in Molecular Dynamics
274     simulations. They proposed a new damped Coulomb potential,
275     \begin{equation}
276     \begin{split}
277     V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
278     \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
279     &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
280     + \frac{2\alpha}{\pi^{1/2}}
281     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
282     \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
283     \end{split}
284     \label{eq:ZahnPot}
285     \end{equation}
286     and showed that this potential does fairly well at capturing the
287     structural and dynamic properties of water compared the same
288     properties obtained using the Ewald sum.
289    
290 chrisfen 2957 \section{Simple Forms for Pairwise Electrostatics}\label{sec:PairwiseDerivation}
291 chrisfen 2918
292     The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
293     al.} are constructed using two different (and separable) computational
294     tricks:
295    
296     \begin{enumerate}
297     \item shifting through the use of image charges, and
298     \item damping the electrostatic interaction.
299     \end{enumerate}
300     Wolf \textit{et al.} treated the
301     development of their summation method as a progressive application of
302     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
303     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
304     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
305     both techniques. It is possible, however, to separate these
306     tricks and study their effects independently.
307    
308     Starting with the original observation that the effective range of the
309     electrostatic interaction in condensed phases is considerably less
310     than $r^{-1}$, either the cutoff sphere neutralization or the
311     distance-dependent damping technique could be used as a foundation for
312     a new pairwise summation method. Wolf \textit{et al.} made the
313     observation that charge neutralization within the cutoff sphere plays
314     a significant role in energy convergence; therefore we will begin our
315     analysis with the various shifted forms that maintain this charge
316     neutralization. We can evaluate the methods of Wolf
317     \textit{et al.} and Zahn \textit{et al.} by considering the standard
318     shifted potential,
319     \begin{equation}
320     V_\textrm{SP}(r) = \begin{cases}
321     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
322     R_\textrm{c}
323     \end{cases},
324     \label{eq:shiftingPotForm}
325     \end{equation}
326     and shifted force,
327     \begin{equation}
328     V_\textrm{SF}(r) = \begin{cases}
329     v(r) - v_\textrm{c}
330     - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
331     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
332     \end{cases},
333     \label{eq:shiftingForm}
334     \end{equation}
335     functions where $v(r)$ is the unshifted form of the potential, and
336     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
337     that both the potential and the forces goes to zero at the cutoff
338     radius, while the Shifted Potential ({\sc sp}) form only ensures the
339     potential is smooth at the cutoff radius
340     ($R_\textrm{c}$).\cite{Allen87}
341    
342     The forces associated with the shifted potential are simply the forces
343     of the unshifted potential itself (when inside the cutoff sphere),
344     \begin{equation}
345     F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
346     \end{equation}
347     and are zero outside. Inside the cutoff sphere, the forces associated
348     with the shifted force form can be written,
349     \begin{equation}
350     F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
351     v(r)}{dr} \right)_{r=R_\textrm{c}}.
352     \end{equation}
353    
354     If the potential, $v(r)$, is taken to be the normal Coulomb potential,
355     \begin{equation}
356     v(r) = \frac{q_i q_j}{r},
357     \label{eq:Coulomb}
358     \end{equation}
359     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
360     al.}'s undamped prescription:
361     \begin{equation}
362     V_\textrm{SP}(r) =
363     q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
364     r\leqslant R_\textrm{c},
365     \label{eq:SPPot}
366     \end{equation}
367     with associated forces,
368     \begin{equation}
369     F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
370     \quad r\leqslant R_\textrm{c}.
371     \label{eq:SPForces}
372     \end{equation}
373     These forces are identical to the forces of the standard Coulomb
374     interaction, and cutting these off at $R_c$ was addressed by Wolf
375     \textit{et al.} as undesirable. They pointed out that the effect of
376     the image charges is neglected in the forces when this form is
377     used,\cite{Wolf99} thereby eliminating any benefit from the method in
378     molecular dynamics. Additionally, there is a discontinuity in the
379     forces at the cutoff radius which results in energy drift during MD
380     simulations.
381    
382     The shifted force ({\sc sf}) form using the normal Coulomb potential
383     will give,
384     \begin{equation}
385     V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
386     + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
387     \quad r\leqslant R_\textrm{c}.
388     \label{eq:SFPot}
389     \end{equation}
390     with associated forces,
391     \begin{equation}
392     F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
393     \quad r\leqslant R_\textrm{c}.
394     \label{eq:SFForces}
395     \end{equation}
396     This formulation has the benefits that there are no discontinuities at
397     the cutoff radius, while the neutralizing image charges are present in
398     both the energy and force expressions. It would be simple to add the
399     self-neutralizing term back when computing the total energy of the
400     system, thereby maintaining the agreement with the Madelung energies.
401     A side effect of this treatment is the alteration in the shape of the
402     potential that comes from the derivative term. Thus, a degree of
403     clarity about agreement with the empirical potential is lost in order
404     to gain functionality in dynamics simulations.
405    
406     Wolf \textit{et al.} originally discussed the energetics of the
407     shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
408     insufficient for accurate determination of the energy with reasonable
409     cutoff distances. The calculated Madelung energies fluctuated around
410     the expected value as the cutoff radius was increased, but the
411     oscillations converged toward the correct value.\cite{Wolf99} A
412     damping function was incorporated to accelerate the convergence; and
413     though alternative forms for the damping function could be
414     used,\cite{Jones56,Heyes81} the complimentary error function was
415     chosen to mirror the effective screening used in the Ewald summation.
416     Incorporating this error function damping into the simple Coulomb
417     potential,
418     \begin{equation}
419     v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
420     \label{eq:dampCoulomb}
421     \end{equation}
422     the shifted potential (eq. (\ref{eq:SPPot})) becomes
423     \begin{equation}
424     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
425     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
426     \quad r\leqslant R_\textrm{c},
427     \label{eq:DSPPot}
428     \end{equation}
429     with associated forces,
430     \begin{equation}
431     F_{\textrm{DSP}}(r) = q_iq_j
432     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
433     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
434     \quad r\leqslant R_\textrm{c}.
435     \label{eq:DSPForces}
436     \end{equation}
437     Again, this damped shifted potential suffers from a
438     force-discontinuity at the cutoff radius, and the image charges play
439     no role in the forces. To remedy these concerns, one may derive a
440     {\sc sf} variant by including the derivative term in
441     eq. (\ref{eq:shiftingForm}),
442     \begin{equation}
443     \begin{split}
444     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
445     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
446     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
447     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
448     + \frac{2\alpha}{\pi^{1/2}}
449     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
450     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
451     \quad r\leqslant R_\textrm{c}.
452     \label{eq:DSFPot}
453     \end{split}
454     \end{equation}
455     The derivative of the above potential will lead to the following forces,
456     \begin{equation}
457     \begin{split}
458     F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
459     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
460     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
461     &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
462     {R_{\textrm{c}}^2}
463     + \frac{2\alpha}{\pi^{1/2}}
464     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
465     \right)\Biggr{]}
466     \quad r\leqslant R_\textrm{c}.
467     \label{eq:DSFForces}
468     \end{split}
469     \end{equation}
470     If the damping parameter $(\alpha)$ is set to zero, the undamped case,
471     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
472     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
473    
474     This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
475     derived by Zahn \textit{et al.}; however, there are two important
476     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
477     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
478     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
479     in the Zahn potential, resulting in a potential discontinuity as
480     particles cross $R_\textrm{c}$. Second, the sign of the derivative
481     portion is different. The missing $v_\textrm{c}$ term would not
482     affect molecular dynamics simulations (although the computed energy
483     would be expected to have sudden jumps as particle distances crossed
484     $R_c$). The sign problem is a potential source of errors, however.
485     In fact, it introduces a discontinuity in the forces at the cutoff,
486     because the force function is shifted in the wrong direction and
487     doesn't cross zero at $R_\textrm{c}$.
488    
489     Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
490     electrostatic summation method in which the potential and forces are
491     continuous at the cutoff radius and which incorporates the damping
492     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
493     this paper, we will evaluate exactly how good these methods ({\sc sp},
494     {\sc sf}, damping) are at reproducing the correct electrostatic
495     summation performed by the Ewald sum.
496    
497    
498     \section{Evaluating Pairwise Summation Techniques}
499    
500     In classical molecular mechanics simulations, there are two primary
501     techniques utilized to obtain information about the system of
502     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
503     techniques utilize pairwise summations of interactions between
504     particle sites, but they use these summations in different ways.
505    
506     In MC, the potential energy difference between configurations dictates
507     the progression of MC sampling. Going back to the origins of this
508     method, the acceptance criterion for the canonical ensemble laid out
509     by Metropolis \textit{et al.} states that a subsequent configuration
510     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
511     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
512     Maintaining the correct $\Delta E$ when using an alternate method for
513     handling the long-range electrostatics will ensure proper sampling
514     from the ensemble.
515    
516     In MD, the derivative of the potential governs how the system will
517     progress in time. Consequently, the force and torque vectors on each
518     body in the system dictate how the system evolves. If the magnitude
519     and direction of these vectors are similar when using alternate
520     electrostatic summation techniques, the dynamics in the short term
521     will be indistinguishable. Because error in MD calculations is
522     cumulative, one should expect greater deviation at longer times,
523     although methods which have large differences in the force and torque
524     vectors will diverge from each other more rapidly.
525    
526     \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
527    
528     The pairwise summation techniques (outlined in section
529     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
530     studying the energy differences between conformations. We took the
531     {\sc spme}-computed energy difference between two conformations to be the
532     correct behavior. An ideal performance by an alternative method would
533     reproduce these energy differences exactly (even if the absolute
534     energies calculated by the methods are different). Since none of the
535     methods provide exact energy differences, we used linear least squares
536     regressions of energy gap data to evaluate how closely the methods
537     mimicked the Ewald energy gaps. Unitary results for both the
538     correlation (slope) and correlation coefficient for these regressions
539     indicate perfect agreement between the alternative method and {\sc spme}.
540     Sample correlation plots for two alternate methods are shown in
541     Fig. \ref{fig:linearFit}.
542    
543     \begin{figure}
544     \centering
545     \includegraphics[width = \linewidth]{./figures/dualLinear.pdf}
546     \caption{Example least squares regressions of the configuration energy
547     differences for SPC/E water systems. The upper plot shows a data set
548     with a poor correlation coefficient ($R^2$), while the lower plot
549     shows a data set with a good correlation coefficient.}
550     \label{fig:linearFit}
551     \end{figure}
552    
553     Each of the seven system types (detailed in section \ref{sec:RepSims})
554     were represented using 500 independent configurations. Thus, each of
555     the alternative (non-Ewald) electrostatic summation methods was
556     evaluated using an accumulated 873,250 configurational energy
557     differences.
558    
559     Results and discussion for the individual analysis of each of the
560 chrisfen 2927 system types appear in sections \ref{sec:IndividualResults}, while the
561 chrisfen 2918 cumulative results over all the investigated systems appear below in
562     sections \ref{sec:EnergyResults}.
563    
564     \subsection{Molecular Dynamics and the Force and Torque
565     Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
566     (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
567     comparing the force and torque vectors with those obtained using the
568     reference Ewald summation ({\sc spme}). Both the magnitude and the
569     direction of these vectors on each of the bodies in the system were
570     analyzed. For the magnitude of these vectors, linear least squares
571     regression analyses were performed as described previously for
572     comparing $\Delta E$ values. Instead of a single energy difference
573     between two system configurations, we compared the magnitudes of the
574     forces (and torques) on each molecule in each configuration. For a
575     system of 1000 water molecules and 40 ions, there are 1040 force
576     vectors and 1000 torque vectors. With 500 configurations, this
577     results in 520,000 force and 500,000 torque vector comparisons.
578     Additionally, data from seven different system types was aggregated
579     before the comparison was made.
580    
581     The {\it directionality} of the force and torque vectors was
582     investigated through measurement of the angle ($\theta$) formed
583     between those computed from the particular method and those from {\sc spme},
584     \begin{equation}
585     \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
586     \cdot \hat{F}_\textrm{M}\right),
587     \end{equation}
588     where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
589     vector computed using method M. Each of these $\theta$ values was
590     accumulated in a distribution function and weighted by the area on the
591     unit sphere. Since this distribution is a measure of angular error
592     between two different electrostatic summation methods, there is no
593     {\it a priori} reason for the profile to adhere to any specific
594     shape. Thus, gaussian fits were used to measure the width of the
595     resulting distributions. The variance ($\sigma^2$) was extracted from
596     each of these fits and was used to compare distribution widths.
597     Values of $\sigma^2$ near zero indicate vector directions
598     indistinguishable from those calculated when using the reference
599     method ({\sc spme}).
600    
601     \subsection{Short-time Dynamics}
602    
603     The effects of the alternative electrostatic summation methods on the
604     short-time dynamics of charged systems were evaluated by considering a
605     NaCl crystal at a temperature of 1000 K. A subset of the best
606     performing pairwise methods was used in this comparison. The NaCl
607     crystal was chosen to avoid possible complications from the treatment
608     of orientational motion in molecular systems. All systems were
609     started with the same initial positions and velocities. Simulations
610     were performed under the microcanonical ensemble, and velocity
611     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
612     of the trajectories,
613     \begin{equation}
614     C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
615     \label{eq:vCorr}
616     \end{equation}
617     Velocity autocorrelation functions require detailed short time data,
618     thus velocity information was saved every 2 fs over 10 ps
619     trajectories. Because the NaCl crystal is composed of two different
620     atom types, the average of the two resulting velocity autocorrelation
621     functions was used for comparisons.
622    
623     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
624    
625     The effects of the same subset of alternative electrostatic methods on
626     the {\it long-time} dynamics of charged systems were evaluated using
627     the same model system (NaCl crystals at 1000K). The power spectrum
628     ($I(\omega)$) was obtained via Fourier transform of the velocity
629     autocorrelation function, \begin{equation} I(\omega) =
630     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
631     \label{eq:powerSpec}
632     \end{equation}
633     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
634     NaCl crystal is composed of two different atom types, the average of
635     the two resulting power spectra was used for comparisons. Simulations
636     were performed under the microcanonical ensemble, and velocity
637     information was saved every 5~fs over 100~ps trajectories.
638    
639     \subsection{Representative Simulations}\label{sec:RepSims}
640     A variety of representative molecular simulations were analyzed to
641     determine the relative effectiveness of the pairwise summation
642     techniques in reproducing the energetics and dynamics exhibited by
643     {\sc spme}. We wanted to span the space of typical molecular
644     simulations (i.e. from liquids of neutral molecules to ionic
645     crystals), so the systems studied were:
646    
647     \begin{enumerate}
648     \item liquid water (SPC/E),\cite{Berendsen87}
649     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
650     \item NaCl crystals,
651     \item NaCl melts,
652     \item a low ionic strength solution of NaCl in water (0.11 M),
653     \item a high ionic strength solution of NaCl in water (1.1 M), and
654     \item a 6\AA\ radius sphere of Argon in water.
655     \end{enumerate}
656    
657     By utilizing the pairwise techniques (outlined in section
658     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
659     charged particles, and mixtures of the two, we hope to discern under
660     which conditions it will be possible to use one of the alternative
661     summation methodologies instead of the Ewald sum.
662    
663     For the solid and liquid water configurations, configurations were
664     taken at regular intervals from high temperature trajectories of 1000
665     SPC/E water molecules. Each configuration was equilibrated
666     independently at a lower temperature (300K for the liquid, 200K for
667     the crystal). The solid and liquid NaCl systems consisted of 500
668     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
669     these systems were selected and equilibrated in the same manner as the
670     water systems. In order to introduce measurable fluctuations in the
671     configuration energy differences, the crystalline simulations were
672     equilibrated at 1000K, near the $T_\textrm{m}$ for NaCl. The liquid
673     NaCl configurations needed to represent a fully disordered array of
674     point charges, so the high temperature of 7000K was selected for
675     equilibration. The ionic solutions were made by solvating 4 (or 40)
676     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
677     water positions were then randomly swapped, and the resulting
678     configurations were again equilibrated individually. Finally, for the
679     Argon / Water ``charge void'' systems, the identities of all the SPC/E
680     waters within 6\AA\ of the center of the equilibrated water
681     configurations were converted to argon.
682    
683     These procedures guaranteed us a set of representative configurations
684     from chemically-relevant systems sampled from appropriate
685     ensembles. Force field parameters for the ions and Argon were taken
686     from the force field utilized by {\sc oopse}.\cite{Meineke05}
687    
688     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
689     We compared the following alternative summation methods with results
690     from the reference method ({\sc spme}):
691    
692     \begin{enumerate}
693     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
694     and 0.3\AA$^{-1}$,
695     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
696     and 0.3\AA$^{-1}$,
697     \item reaction field with an infinite dielectric constant, and
698     \item an unmodified cutoff.
699     \end{enumerate}
700    
701     Group-based cutoffs with a fifth-order polynomial switching function
702     were utilized for the reaction field simulations. Additionally, we
703     investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
704     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
705     implementation of {\sc spme},\cite{Ponder87} while all other calculations
706     were performed using the {\sc oopse} molecular mechanics
707     package.\cite{Meineke05} All other portions of the energy calculation
708     (i.e. Lennard-Jones interactions) were handled in exactly the same
709     manner across all systems and configurations.
710    
711     The alternative methods were also evaluated with three different
712     cutoff radii (9, 12, and 15\AA). As noted previously, the
713     convergence parameter ($\alpha$) plays a role in the balance of the
714     real-space and reciprocal-space portions of the Ewald calculation.
715     Typical molecular mechanics packages set this to a value dependent on
716     the cutoff radius and a tolerance (typically less than $1 \times
717     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
718     increasing accuracy at the expense of computational time spent on the
719     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
720     The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
721     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
722     0.3119, and 0.2476\AA$^{-1}$ for cutoff radii of 9, 12, and 15\AA\
723     respectively.
724    
725 chrisfen 2927 \section{Configuration Energy Difference Results}\label{sec:EnergyResults}
726 chrisfen 2918 In order to evaluate the performance of the pairwise electrostatic
727 chrisfen 2920 summation methods for Monte Carlo (MC) simulations, the energy
728     differences between configurations were compared to the values
729     obtained when using {\sc spme}. The results for the combined
730     regression analysis of all of the systems are shown in figure
731     \ref{fig:delE}.
732 chrisfen 2918
733     \begin{figure}
734     \centering
735     \includegraphics[width=4.75in]{./figures/delEplot.pdf}
736     \caption{Statistical analysis of the quality of configurational energy
737     differences for a given electrostatic method compared with the
738     reference Ewald sum. Results with a value equal to 1 (dashed line)
739     indicate $\Delta E$ values indistinguishable from those obtained using
740     {\sc spme}. Different values of the cutoff radius are indicated with
741     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
742     inverted triangles).}
743     \label{fig:delE}
744     \end{figure}
745    
746     The most striking feature of this plot is how well the Shifted Force
747     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
748     differences. For the undamped {\sc sf} method, and the
749     moderately-damped {\sc sp} methods, the results are nearly
750     indistinguishable from the Ewald results. The other common methods do
751     significantly less well.
752    
753     The unmodified cutoff method is essentially unusable. This is not
754     surprising since hard cutoffs give large energy fluctuations as atoms
755     or molecules move in and out of the cutoff
756     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
757     some degree by using group based cutoffs with a switching
758     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
759     significant improvement using the group-switched cutoff because the
760     salt and salt solution systems contain non-neutral groups. Section
761 chrisfen 2927 \ref{sec:IndividualResults} includes results for systems comprised entirely
762 chrisfen 2918 of neutral groups.
763    
764     For the {\sc sp} method, inclusion of electrostatic damping improves
765     the agreement with Ewald, and using an $\alpha$ of 0.2\AA $^{-1}$
766     shows an excellent correlation and quality of fit with the {\sc spme}
767     results, particularly with a cutoff radius greater than 12
768     \AA . Use of a larger damping parameter is more helpful for the
769     shortest cutoff shown, but it has a detrimental effect on simulations
770     with larger cutoffs.
771    
772     In the {\sc sf} sets, increasing damping results in progressively {\it
773     worse} correlation with Ewald. Overall, the undamped case is the best
774     performing set, as the correlation and quality of fits are
775     consistently superior regardless of the cutoff distance. The undamped
776     case is also less computationally demanding (because no evaluation of
777     the complementary error function is required).
778    
779     The reaction field results illustrates some of that method's
780 chrisfen 2957 limitations, primarily that it was developed for use in homogeneous
781 chrisfen 2918 systems; although it does provide results that are an improvement over
782     those from an unmodified cutoff.
783    
784 chrisfen 2927 \section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults}
785 chrisfen 2918
786     Evaluation of pairwise methods for use in Molecular Dynamics
787     simulations requires consideration of effects on the forces and
788     torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
789     regression results for the force and torque vector magnitudes,
790     respectively. The data in these figures was generated from an
791     accumulation of the statistics from all of the system types.
792    
793     \begin{figure}
794     \centering
795     \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
796     \caption{Statistical analysis of the quality of the force vector
797     magnitudes for a given electrostatic method compared with the
798     reference Ewald sum. Results with a value equal to 1 (dashed line)
799     indicate force magnitude values indistinguishable from those obtained
800     using {\sc spme}. Different values of the cutoff radius are indicated with
801     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
802     inverted triangles).}
803     \label{fig:frcMag}
804     \end{figure}
805    
806     Again, it is striking how well the Shifted Potential and Shifted Force
807     methods are doing at reproducing the {\sc spme} forces. The undamped and
808     weakly-damped {\sc sf} method gives the best agreement with Ewald.
809     This is perhaps expected because this method explicitly incorporates a
810     smooth transition in the forces at the cutoff radius as well as the
811     neutralizing image charges.
812    
813     Figure \ref{fig:frcMag}, for the most part, parallels the results seen
814     in the previous $\Delta E$ section. The unmodified cutoff results are
815     poor, but using group based cutoffs and a switching function provides
816     an improvement much more significant than what was seen with $\Delta
817     E$.
818    
819     With moderate damping and a large enough cutoff radius, the {\sc sp}
820     method is generating usable forces. Further increases in damping,
821     while beneficial for simulations with a cutoff radius of 9\AA\ , is
822     detrimental to simulations with larger cutoff radii.
823    
824     The reaction field results are surprisingly good, considering the poor
825     quality of the fits for the $\Delta E$ results. There is still a
826     considerable degree of scatter in the data, but the forces correlate
827     well with the Ewald forces in general. We note that the reaction
828     field calculations do not include the pure NaCl systems, so these
829     results are partly biased towards conditions in which the method
830     performs more favorably.
831    
832     \begin{figure}
833     \centering
834     \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
835     \caption{Statistical analysis of the quality of the torque vector
836     magnitudes for a given electrostatic method compared with the
837     reference Ewald sum. Results with a value equal to 1 (dashed line)
838     indicate torque magnitude values indistinguishable from those obtained
839     using {\sc spme}. Different values of the cutoff radius are indicated with
840     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
841     inverted triangles).}
842     \label{fig:trqMag}
843     \end{figure}
844    
845     Molecular torques were only available from the systems which contained
846     rigid molecules (i.e. the systems containing water). The data in
847     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
848    
849     Torques appear to be much more sensitive to charges at a longer
850     distance. The striking feature in comparing the new electrostatic
851     methods with {\sc spme} is how much the agreement improves with increasing
852     cutoff radius. Again, the weakly damped and undamped {\sc sf} method
853     appears to be reproducing the {\sc spme} torques most accurately.
854    
855     Water molecules are dipolar, and the reaction field method reproduces
856     the effect of the surrounding polarized medium on each of the
857     molecular bodies. Therefore it is not surprising that reaction field
858     performs best of all of the methods on molecular torques.
859    
860 chrisfen 2927 \section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults}
861 chrisfen 2918
862     It is clearly important that a new electrostatic method can reproduce
863     the magnitudes of the force and torque vectors obtained via the Ewald
864     sum. However, the {\it directionality} of these vectors will also be
865     vital in calculating dynamical quantities accurately. Force and
866     torque directionalities were investigated by measuring the angles
867     formed between these vectors and the same vectors calculated using
868     {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
869     variance ($\sigma^2$) of the Gaussian fits of the angle error
870     distributions of the combined set over all system types.
871    
872     \begin{figure}
873     \centering
874     \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
875     \caption{Statistical analysis of the width of the angular distribution
876     that the force and torque vectors from a given electrostatic method
877     make with their counterparts obtained using the reference Ewald sum.
878     Results with a variance ($\sigma^2$) equal to zero (dashed line)
879     indicate force and torque directions indistinguishable from those
880     obtained using {\sc spme}. Different values of the cutoff radius are
881     indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
882     and 15\AA\ = inverted triangles).}
883     \label{fig:frcTrqAng}
884     \end{figure}
885    
886     Both the force and torque $\sigma^2$ results from the analysis of the
887     total accumulated system data are tabulated in figure
888     \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
889     sp}) method would be essentially unusable for molecular dynamics
890     unless the damping function is added. The Shifted Force ({\sc sf})
891     method, however, is generating force and torque vectors which are
892     within a few degrees of the Ewald results even with weak (or no)
893     damping.
894    
895     All of the sets (aside from the over-damped case) show the improvement
896     afforded by choosing a larger cutoff radius. Increasing the cutoff
897     from 9 to 12\AA\ typically results in a halving of the width of the
898     distribution, with a similar improvement when going from 12 to 15
899     \AA .
900    
901     The undamped {\sc sf}, group-based cutoff, and reaction field methods
902     all do equivalently well at capturing the direction of both the force
903     and torque vectors. Using the electrostatic damping improves the
904     angular behavior significantly for the {\sc sp} and moderately for the
905 chrisfen 2957 {\sc sf} methods. Over-damping is detrimental to both methods. Again
906 chrisfen 2918 it is important to recognize that the force vectors cover all
907     particles in all seven systems, while torque vectors are only
908     available for neutral molecular groups. Damping is more beneficial to
909     charged bodies, and this observation is investigated further in
910 chrisfen 2957 section \ref{sec:IndividualResults}.
911 chrisfen 2918
912     Although not discussed previously, group based cutoffs can be applied
913     to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
914     will reintroduce small discontinuities at the cutoff radius, but the
915     effects of these can be minimized by utilizing a switching function.
916     Though there are no significant benefits or drawbacks observed in
917     $\Delta E$ and the force and torque magnitudes when doing this, there
918     is a measurable improvement in the directionality of the forces and
919     torques. Table \ref{tab:groupAngle} shows the angular variances
920     obtained both without (N) and with (Y) group based cutoffs and a
921     switching function. Note that the $\alpha$ values have units of
922     \AA$^{-1}$ and the variance values have units of degrees$^2$. The
923     {\sc sp} (with an $\alpha$ of 0.2\AA$^{-1}$ or smaller) shows much
924     narrower angular distributions when using group-based cutoffs. The
925     {\sc sf} method likewise shows improvement in the undamped and lightly
926     damped cases.
927    
928     \begin{table}
929     \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
930     THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
931     GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
932     THE REFERENCE EWALD SUMMATION}
933    
934     \footnotesize
935     \begin{center}
936     \begin{tabular}{@{} ccrrrrrrrr @{}} \\
937     \toprule
938     \toprule
939    
940     & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
941     Force} \\
942     \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
943     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
944     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
945    
946     \midrule
947    
948     9\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
949     & \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} \\
950     12\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
951     & \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} \\
952     15\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
953     & \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} \\
954    
955     \midrule
956    
957     9\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
958     & \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} \\
959     12\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
960     & \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} \\
961     15\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
962     & \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} \\
963    
964     \bottomrule
965     \end{tabular}
966     \end{center}
967     \label{tab:groupAngle}
968     \end{table}
969    
970     One additional trend in table \ref{tab:groupAngle} is that the
971     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
972     increases, something that is more obvious with group-based cutoffs.
973     The complimentary error function inserted into the potential weakens
974     the electrostatic interaction as the value of $\alpha$ is increased.
975 chrisfen 2957 However, at larger values of $\alpha$, it is possible to over-damp the
976 chrisfen 2918 electrostatic interaction and to remove it completely. Kast
977     \textit{et al.} developed a method for choosing appropriate $\alpha$
978     values for these types of electrostatic summation methods by fitting
979     to $g(r)$ data, and their methods indicate optimal values of 0.34,
980     0.25, and 0.16\AA$^{-1}$ for cutoff values of 9, 12, and 15\AA\
981     respectively.\cite{Kast03} These appear to be reasonable choices to
982     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
983     these findings, choices this high would introduce error in the
984     molecular torques, particularly for the shorter cutoffs. Based on our
985     observations, empirical damping up to 0.2\AA$^{-1}$ is beneficial,
986     but damping may be unnecessary when using the {\sc sf} method.
987    
988 chrisfen 2927 \section{Individual System Analysis Results}\label{sec:IndividualResults}
989 chrisfen 2918
990 chrisfen 2920 The combined results of the previous sections show how the pairwise
991     methods compare to the Ewald summation in the general sense over all
992     of the system types. It is also useful to consider each of the
993     studied systems in an individual fashion, so that we can identify
994     conditions that are particularly difficult for a selected pairwise
995     method to address. This allows us to further establish the limitations
996     of these pairwise techniques. Below, the energy difference, force
997     vector, and torque vector analyses are presented on an individual
998     system basis.
999    
1000 chrisfen 2927 \subsection{SPC/E Water Results}\label{sec:WaterResults}
1001 chrisfen 2920
1002 chrisfen 2927 The first system considered was liquid water at 300K using the SPC/E
1003     model of water.\cite{Berendsen87} The results for the energy gap
1004     comparisons and the force and torque vector magnitude comparisons are
1005     shown in table \ref{tab:spce}. The force and torque vector
1006     directionality results are displayed separately in table
1007     \ref{tab:spceAng}, where the effect of group-based cutoffs and
1008     switching functions on the {\sc sp} and {\sc sf} potentials are also
1009     investigated. In all of the individual results table, the method
1010     abbreviations are as follows:
1011 chrisfen 2920
1012 chrisfen 2927 \begin{itemize}
1013     \item PC = Pure Cutoff,
1014     \item SP = Shifted Potential,
1015     \item SF = Shifted Force,
1016     \item GSC = Group Switched Cutoff,
1017     \item RF = Reaction Field (where $\varepsilon \approx\infty$),
1018     \item GSSP = Group Switched Shifted Potential, and
1019     \item GSSF = Group Switched Shifted Force.
1020     \end{itemize}
1021 chrisfen 2920
1022 chrisfen 2927 \begin{table}[htbp]
1023     \centering
1024     \caption{REGRESSION RESULTS OF THE LIQUID WATER SYSTEM FOR THE
1025     $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it middle})
1026     AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1027 chrisfen 2920
1028 chrisfen 2927 \footnotesize
1029     \begin{tabular}{@{} ccrrrrrr @{}}
1030     \\
1031     \toprule
1032     \toprule
1033     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1034     \cmidrule(lr){3-4}
1035     \cmidrule(lr){5-6}
1036     \cmidrule(l){7-8}
1037     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1038     \midrule
1039     PC & & 3.046 & 0.002 & -3.018 & 0.002 & 4.719 & 0.005 \\
1040     SP & 0.0 & 1.035 & 0.218 & 0.908 & 0.313 & 1.037 & 0.470 \\
1041     & 0.1 & 1.021 & 0.387 & 0.965 & 0.752 & 1.006 & 0.947 \\
1042     & 0.2 & 0.997 & 0.962 & 1.001 & 0.994 & 0.994 & 0.996 \\
1043     & 0.3 & 0.984 & 0.980 & 0.997 & 0.985 & 0.982 & 0.987 \\
1044     SF & 0.0 & 0.977 & 0.974 & 0.996 & 0.992 & 0.991 & 0.997 \\
1045     & 0.1 & 0.983 & 0.974 & 1.001 & 0.994 & 0.996 & 0.998 \\
1046     & 0.2 & 0.992 & 0.989 & 1.001 & 0.995 & 0.994 & 0.996 \\
1047     & 0.3 & 0.984 & 0.980 & 0.996 & 0.985 & 0.982 & 0.987 \\
1048     GSC & & 0.918 & 0.862 & 0.852 & 0.756 & 0.801 & 0.700 \\
1049     RF & & 0.971 & 0.958 & 0.975 & 0.987 & 0.959 & 0.983 \\
1050     \midrule
1051     PC & & -1.647 & 0.000 & -0.127 & 0.000 & -0.979 & 0.000 \\
1052     SP & 0.0 & 0.735 & 0.368 & 0.813 & 0.537 & 0.865 & 0.659 \\
1053     & 0.1 & 0.850 & 0.612 & 0.956 & 0.887 & 0.992 & 0.979 \\
1054     & 0.2 & 0.996 & 0.989 & 1.000 & 1.000 & 1.000 & 1.000 \\
1055     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1056     SF & 0.0 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 0.999 \\
1057     & 0.1 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1058     & 0.2 & 0.999 & 0.998 & 1.000 & 1.000 & 1.000 & 1.000 \\
1059     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1060     GSC & & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1061     RF & & 0.999 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1062     \midrule
1063     PC & & 2.387 & 0.000 & 0.183 & 0.000 & 1.282 & 0.000 \\
1064     SP & 0.0 & 0.847 & 0.543 & 0.904 & 0.694 & 0.935 & 0.786 \\
1065     & 0.1 & 0.922 & 0.749 & 0.980 & 0.934 & 0.996 & 0.988 \\
1066     & 0.2 & 0.987 & 0.985 & 0.989 & 0.992 & 0.990 & 0.993 \\
1067     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1068     SF & 0.0 & 0.978 & 0.990 & 0.988 & 0.997 & 0.993 & 0.999 \\
1069     & 0.1 & 0.983 & 0.991 & 0.993 & 0.997 & 0.997 & 0.999 \\
1070     & 0.2 & 0.986 & 0.989 & 0.989 & 0.992 & 0.990 & 0.993 \\
1071     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1072     GSC & & 0.995 & 0.981 & 0.999 & 0.991 & 1.001 & 0.994 \\
1073     RF & & 0.993 & 0.989 & 0.998 & 0.996 & 1.000 & 0.999 \\
1074     \bottomrule
1075     \end{tabular}
1076     \label{tab:spce}
1077     \end{table}
1078 chrisfen 2920
1079 chrisfen 2927 \begin{table}[htbp]
1080     \centering
1081     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1082     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE LIQUID WATER
1083     SYSTEM}
1084 chrisfen 2920
1085 chrisfen 2927 \footnotesize
1086     \begin{tabular}{@{} ccrrrrrr @{}}
1087     \\
1088     \toprule
1089     \toprule
1090     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1091     \cmidrule(lr){3-5}
1092     \cmidrule(l){6-8}
1093     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1094     \midrule
1095     PC & & 783.759 & 481.353 & 332.677 & 248.674 & 144.382 & 98.535 \\
1096     SP & 0.0 & 659.440 & 380.699 & 250.002 & 235.151 & 134.661 & 88.135 \\
1097     & 0.1 & 293.849 & 67.772 & 11.609 & 105.090 & 23.813 & 4.369 \\
1098     & 0.2 & 5.975 & 0.136 & 0.094 & 5.553 & 1.784 & 1.536 \\
1099     & 0.3 & 0.725 & 0.707 & 0.693 & 7.293 & 6.933 & 6.748 \\
1100     SF & 0.0 & 2.238 & 0.713 & 0.292 & 3.290 & 1.090 & 0.416 \\
1101     & 0.1 & 2.238 & 0.524 & 0.115 & 3.184 & 0.945 & 0.326 \\
1102     & 0.2 & 0.374 & 0.102 & 0.094 & 2.598 & 1.755 & 1.537 \\
1103     & 0.3 & 0.721 & 0.707 & 0.693 & 7.322 & 6.933 & 6.748 \\
1104     GSC & & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1105     RF & & 2.091 & 0.403 & 0.113 & 3.583 & 1.071 & 0.399 \\
1106     \midrule
1107     GSSP & 0.0 & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1108     & 0.1 & 1.879 & 0.291 & 0.057 & 3.983 & 1.117 & 0.370 \\
1109     & 0.2 & 0.443 & 0.103 & 0.093 & 2.821 & 1.794 & 1.532 \\
1110     & 0.3 & 0.728 & 0.694 & 0.692 & 7.387 & 6.942 & 6.748 \\
1111     GSSF & 0.0 & 1.298 & 0.270 & 0.083 & 3.098 & 0.992 & 0.375 \\
1112     & 0.1 & 1.296 & 0.210 & 0.044 & 3.055 & 0.922 & 0.330 \\
1113     & 0.2 & 0.433 & 0.104 & 0.093 & 2.895 & 1.797 & 1.532 \\
1114     & 0.3 & 0.728 & 0.694 & 0.692 & 7.410 & 6.942 & 6.748 \\
1115     \bottomrule
1116     \end{tabular}
1117     \label{tab:spceAng}
1118     \end{table}
1119    
1120     The water results parallel the combined results seen in sections
1121     \ref{sec:EnergyResults} through \ref{sec:FTDirResults}. There is good
1122     agreement with {\sc spme} in both energetic and dynamic behavior when
1123     using the {\sc sf} method with and without damping. The {\sc sp}
1124     method does well with an $\alpha$ around 0.2\AA$^{-1}$, particularly
1125 chrisfen 2957 with cutoff radii greater than 12\AA. Over-damping the electrostatics
1126 chrisfen 2927 reduces the agreement between both these methods and {\sc spme}.
1127    
1128     The pure cutoff ({\sc pc}) method performs poorly, again mirroring the
1129     observations from the combined results. In contrast to these results, however, the use of a switching function and group
1130     based cutoffs greatly improves the results for these neutral water
1131     molecules. The group switched cutoff ({\sc gsc}) does not mimic the
1132     energetics of {\sc spme} as well as the {\sc sp} (with moderate
1133     damping) and {\sc sf} methods, but the dynamics are quite good. The
1134     switching functions correct discontinuities in the potential and
1135     forces, leading to these improved results. Such improvements with the
1136     use of a switching function have been recognized in previous
1137     studies,\cite{Andrea83,Steinbach94} and this proves to be a useful
1138     tactic for stably incorporating local area electrostatic effects.
1139    
1140     The reaction field ({\sc rf}) method simply extends upon the results
1141     observed in the {\sc gsc} case. Both methods are similar in form
1142     (i.e. neutral groups, switching function), but {\sc rf} incorporates
1143     an added effect from the external dielectric. This similarity
1144     translates into the same good dynamic results and improved energetic
1145     agreement with {\sc spme}. Though this agreement is not to the level
1146     of the moderately damped {\sc sp} and {\sc sf} methods, these results
1147     show how incorporating some implicit properties of the surroundings
1148     (i.e. $\epsilon_\textrm{S}$) can improve the solvent depiction.
1149    
1150     As a final note for the liquid water system, use of group cutoffs and a
1151     switching function leads to noticeable improvements in the {\sc sp}
1152     and {\sc sf} methods, primarily in directionality of the force and
1153     torque vectors (table \ref{tab:spceAng}). The {\sc sp} method shows
1154     significant narrowing of the angle distribution when using little to
1155     no damping and only modest improvement for the recommended conditions
1156     ($\alpha = 0.2$\AA$^{-1}$ and $R_\textrm{c}~\geqslant~12$\AA). The
1157     {\sc sf} method shows modest narrowing across all damping and cutoff
1158 chrisfen 2957 ranges of interest. When over-damping these methods, group cutoffs and
1159 chrisfen 2927 the switching function do not improve the force and torque
1160     directionalities.
1161    
1162     \subsection{SPC/E Ice I$_\textrm{c}$ Results}\label{sec:IceResults}
1163    
1164     In addition to the disordered molecular system above, the ordered
1165     molecular system of ice I$_\textrm{c}$ was also considered. Ice
1166     polymorph could have been used to fit this role; however, ice
1167     I$_\textrm{c}$ was chosen because it can form an ideal periodic
1168     lattice with the same number of water molecules used in the disordered
1169     liquid state case. The results for the energy gap comparisons and the
1170     force and torque vector magnitude comparisons are shown in table
1171     \ref{tab:ice}. The force and torque vector directionality results are
1172     displayed separately in table \ref{tab:iceAng}, where the effect of
1173     group-based cutoffs and switching functions on the {\sc sp} and {\sc
1174     sf} potentials are also displayed.
1175    
1176     \begin{table}[htbp]
1177     \centering
1178     \caption{REGRESSION RESULTS OF THE ICE I$_\textrm{c}$ SYSTEM FOR
1179     $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it
1180     middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1181    
1182     \footnotesize
1183     \begin{tabular}{@{} ccrrrrrr @{}}
1184     \\
1185     \toprule
1186     \toprule
1187     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1188     \cmidrule(lr){3-4}
1189     \cmidrule(lr){5-6}
1190     \cmidrule(l){7-8}
1191     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1192     \midrule
1193     PC & & 19.897 & 0.047 & -29.214 & 0.048 & -3.771 & 0.001 \\
1194     SP & 0.0 & -0.014 & 0.000 & 2.135 & 0.347 & 0.457 & 0.045 \\
1195     & 0.1 & 0.321 & 0.017 & 1.490 & 0.584 & 0.886 & 0.796 \\
1196     & 0.2 & 0.896 & 0.872 & 1.011 & 0.998 & 0.997 & 0.999 \\
1197     & 0.3 & 0.983 & 0.997 & 0.992 & 0.997 & 0.991 & 0.997 \\
1198     SF & 0.0 & 0.943 & 0.979 & 1.048 & 0.978 & 0.995 & 0.999 \\
1199     & 0.1 & 0.948 & 0.979 & 1.044 & 0.983 & 1.000 & 0.999 \\
1200     & 0.2 & 0.982 & 0.997 & 0.969 & 0.960 & 0.997 & 0.999 \\
1201     & 0.3 & 0.985 & 0.997 & 0.961 & 0.961 & 0.991 & 0.997 \\
1202     GSC & & 0.983 & 0.985 & 0.966 & 0.994 & 1.003 & 0.999 \\
1203     RF & & 0.924 & 0.944 & 0.990 & 0.996 & 0.991 & 0.998 \\
1204     \midrule
1205     PC & & -4.375 & 0.000 & 6.781 & 0.000 & -3.369 & 0.000 \\
1206     SP & 0.0 & 0.515 & 0.164 & 0.856 & 0.426 & 0.743 & 0.478 \\
1207     & 0.1 & 0.696 & 0.405 & 0.977 & 0.817 & 0.974 & 0.964 \\
1208     & 0.2 & 0.981 & 0.980 & 1.001 & 1.000 & 1.000 & 1.000 \\
1209     & 0.3 & 0.996 & 0.998 & 0.997 & 0.999 & 0.997 & 0.999 \\
1210     SF & 0.0 & 0.991 & 0.995 & 1.003 & 0.998 & 0.999 & 1.000 \\
1211     & 0.1 & 0.992 & 0.995 & 1.003 & 0.998 & 1.000 & 1.000 \\
1212     & 0.2 & 0.998 & 0.998 & 0.981 & 0.962 & 1.000 & 1.000 \\
1213     & 0.3 & 0.996 & 0.998 & 0.976 & 0.957 & 0.997 & 0.999 \\
1214     GSC & & 0.997 & 0.996 & 0.998 & 0.999 & 1.000 & 1.000 \\
1215     RF & & 0.988 & 0.989 & 1.000 & 0.999 & 1.000 & 1.000 \\
1216     \midrule
1217     PC & & -6.367 & 0.000 & -3.552 & 0.000 & -3.447 & 0.000 \\
1218     SP & 0.0 & 0.643 & 0.409 & 0.833 & 0.607 & 0.961 & 0.805 \\
1219     & 0.1 & 0.791 & 0.683 & 0.957 & 0.914 & 1.000 & 0.989 \\
1220     & 0.2 & 0.974 & 0.991 & 0.993 & 0.998 & 0.993 & 0.998 \\
1221     & 0.3 & 0.976 & 0.992 & 0.977 & 0.992 & 0.977 & 0.992 \\
1222     SF & 0.0 & 0.979 & 0.997 & 0.992 & 0.999 & 0.994 & 1.000 \\
1223     & 0.1 & 0.984 & 0.997 & 0.996 & 0.999 & 0.998 & 1.000 \\
1224     & 0.2 & 0.991 & 0.997 & 0.974 & 0.958 & 0.993 & 0.998 \\
1225     & 0.3 & 0.977 & 0.992 & 0.956 & 0.948 & 0.977 & 0.992 \\
1226     GSC & & 0.999 & 0.997 & 0.996 & 0.999 & 1.002 & 1.000 \\
1227     RF & & 0.994 & 0.997 & 0.997 & 0.999 & 1.000 & 1.000 \\
1228     \bottomrule
1229     \end{tabular}
1230     \label{tab:ice}
1231     \end{table}
1232    
1233     \begin{table}[htbp]
1234     \centering
1235     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1236     OF THE FORCE AND TORQUE VECTORS IN THE ICE I$_\textrm{c}$ SYSTEM}
1237    
1238     \footnotesize
1239     \begin{tabular}{@{} ccrrrrrr @{}}
1240     \\
1241     \toprule
1242     \toprule
1243     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque
1244     $\sigma^2$} \\
1245     \cmidrule(lr){3-5}
1246     \cmidrule(l){6-8}
1247     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1248     \midrule
1249     PC & & 2128.921 & 603.197 & 715.579 & 329.056 & 221.397 & 81.042 \\
1250     SP & 0.0 & 1429.341 & 470.320 & 447.557 & 301.678 & 197.437 & 73.840 \\
1251     & 0.1 & 590.008 & 107.510 & 18.883 & 118.201 & 32.472 & 3.599 \\
1252     & 0.2 & 10.057 & 0.105 & 0.038 & 2.875 & 0.572 & 0.518 \\
1253     & 0.3 & 0.245 & 0.260 & 0.262 & 2.365 & 2.396 & 2.327 \\
1254     SF & 0.0 & 1.745 & 1.161 & 0.212 & 1.135 & 0.426 & 0.155 \\
1255     & 0.1 & 1.721 & 0.868 & 0.082 & 1.118 & 0.358 & 0.118 \\
1256     & 0.2 & 0.201 & 0.040 & 0.038 & 0.786 & 0.555 & 0.518 \\
1257     & 0.3 & 0.241 & 0.260 & 0.262 & 2.368 & 2.400 & 2.327 \\
1258     GSC & & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1259     RF & & 2.887 & 0.217 & 0.107 & 1.006 & 0.281 & 0.085 \\
1260     \midrule
1261     GSSP & 0.0 & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1262     & 0.1 & 1.341 & 0.123 & 0.037 & 0.835 & 0.234 & 0.085 \\
1263     & 0.2 & 0.558 & 0.040 & 0.037 & 0.823 & 0.557 & 0.519 \\
1264     & 0.3 & 0.250 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1265     GSSF & 0.0 & 2.124 & 0.132 & 0.069 & 0.919 & 0.263 & 0.099 \\
1266     & 0.1 & 2.165 & 0.101 & 0.035 & 0.895 & 0.244 & 0.096 \\
1267     & 0.2 & 0.706 & 0.040 & 0.037 & 0.870 & 0.559 & 0.519 \\
1268     & 0.3 & 0.251 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1269     \bottomrule
1270     \end{tabular}
1271     \label{tab:iceAng}
1272     \end{table}
1273    
1274     Highly ordered systems are a difficult test for the pairwise methods
1275     in that they lack the implicit periodicity of the Ewald summation. As
1276     expected, the energy gap agreement with {\sc spme} is reduced for the
1277     {\sc sp} and {\sc sf} methods with parameters that were ideal for the
1278     disordered liquid system. Moving to higher $R_\textrm{c}$ helps
1279     improve the agreement, though at an increase in computational cost.
1280     The dynamics of this crystalline system (both in magnitude and
1281     direction) are little affected. Both methods still reproduce the Ewald
1282     behavior with the same parameter recommendations from the previous
1283     section.
1284    
1285     It is also worth noting that {\sc rf} exhibits improved energy gap
1286     results over the liquid water system. One possible explanation is
1287     that the ice I$_\textrm{c}$ crystal is ordered such that the net
1288     dipole moment of the crystal is zero. With $\epsilon_\textrm{S} =
1289     \infty$, the reaction field incorporates this structural organization
1290     by actively enforcing a zeroed dipole moment within each cutoff
1291     sphere.
1292    
1293     \subsection{NaCl Melt Results}\label{sec:SaltMeltResults}
1294    
1295     A high temperature NaCl melt was tested to gauge the accuracy of the
1296     pairwise summation methods in a disordered system of charges. The
1297     results for the energy gap comparisons and the force vector magnitude
1298     comparisons are shown in table \ref{tab:melt}. The force vector
1299     directionality results are displayed separately in table
1300     \ref{tab:meltAng}.
1301    
1302     \begin{table}[htbp]
1303     \centering
1304     \caption{REGRESSION RESULTS OF THE MOLTEN SODIUM CHLORIDE SYSTEM FOR
1305     $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES ({\it
1306     lower})}
1307    
1308     \footnotesize
1309     \begin{tabular}{@{} ccrrrrrr @{}}
1310     \\
1311     \toprule
1312     \toprule
1313     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1314     \cmidrule(lr){3-4}
1315     \cmidrule(lr){5-6}
1316     \cmidrule(l){7-8}
1317     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1318     \midrule
1319     PC & & -0.008 & 0.000 & -0.049 & 0.005 & -0.136 & 0.020 \\
1320     SP & 0.0 & 0.928 & 0.996 & 0.931 & 0.998 & 0.950 & 0.999 \\
1321     & 0.1 & 0.977 & 0.998 & 0.998 & 1.000 & 0.997 & 1.000 \\
1322     & 0.2 & 0.960 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1323     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1324     SF & 0.0 & 0.996 & 1.000 & 0.995 & 1.000 & 0.997 & 1.000 \\
1325     & 0.1 & 1.021 & 1.000 & 1.024 & 1.000 & 1.007 & 1.000 \\
1326     & 0.2 & 0.966 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1327     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1328     \midrule
1329     PC & & 1.103 & 0.000 & 0.989 & 0.000 & 0.802 & 0.000 \\
1330     SP & 0.0 & 0.973 & 0.981 & 0.975 & 0.988 & 0.979 & 0.992 \\
1331     & 0.1 & 0.987 & 0.992 & 0.993 & 0.998 & 0.997 & 0.999 \\
1332     & 0.2 & 0.993 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1333     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1334     SF & 0.0 & 0.996 & 0.997 & 0.997 & 0.999 & 0.998 & 1.000 \\
1335     & 0.1 & 1.000 & 0.997 & 1.001 & 0.999 & 1.000 & 1.000 \\
1336     & 0.2 & 0.994 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1337     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1338     \bottomrule
1339     \end{tabular}
1340     \label{tab:melt}
1341     \end{table}
1342    
1343     \begin{table}[htbp]
1344     \centering
1345     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1346     OF THE FORCE VECTORS IN THE MOLTEN SODIUM CHLORIDE SYSTEM}
1347    
1348     \footnotesize
1349     \begin{tabular}{@{} ccrrrrrr @{}}
1350     \\
1351     \toprule
1352     \toprule
1353     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1354     \cmidrule(lr){3-5}
1355     \cmidrule(l){6-8}
1356     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1357     \midrule
1358     PC & & 13.294 & 8.035 & 5.366 \\
1359     SP & 0.0 & 13.316 & 8.037 & 5.385 \\
1360     & 0.1 & 5.705 & 1.391 & 0.360 \\
1361     & 0.2 & 2.415 & 7.534 & 13.927 \\
1362     & 0.3 & 23.769 & 67.306 & 57.252 \\
1363     SF & 0.0 & 1.693 & 0.603 & 0.256 \\
1364     & 0.1 & 1.687 & 0.653 & 0.272 \\
1365     & 0.2 & 2.598 & 7.523 & 13.930 \\
1366     & 0.3 & 23.734 & 67.305 & 57.252 \\
1367     \bottomrule
1368     \end{tabular}
1369     \label{tab:meltAng}
1370     \end{table}
1371    
1372     The molten NaCl system shows more sensitivity to the electrostatic
1373     damping than the water systems. The most noticeable point is that the
1374     undamped {\sc sf} method does very well at replicating the {\sc spme}
1375     configurational energy differences and forces. Light damping appears
1376     to minimally improve the dynamics, but this comes with a deterioration
1377     of the energy gap results. In contrast, this light damping improves
1378     the {\sc sp} energy gaps and forces. Moderate and heavy electrostatic
1379     damping reduce the agreement with {\sc spme} for both methods. From
1380     these observations, the undamped {\sc sf} method is the best choice
1381     for disordered systems of charges.
1382    
1383     \subsection{NaCl Crystal Results}\label{sec:SaltCrystalResults}
1384    
1385     Similar to the use of ice I$_\textrm{c}$ to investigate the role of
1386     order in molecular systems on the effectiveness of the pairwise
1387     methods, the 1000K NaCl crystal system was used to investigate the
1388     accuracy of the pairwise summation methods in an ordered system of
1389     charged particles. The results for the energy gap comparisons and the
1390     force vector magnitude comparisons are shown in table \ref{tab:salt}.
1391     The force vector directionality results are displayed separately in
1392     table \ref{tab:saltAng}.
1393    
1394     \begin{table}[htbp]
1395     \centering
1396     \caption{REGRESSION RESULTS OF THE CRYSTALLINE SODIUM CHLORIDE
1397     SYSTEM FOR $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES
1398     ({\it lower})}
1399    
1400     \footnotesize
1401     \begin{tabular}{@{} ccrrrrrr @{}}
1402     \\
1403     \toprule
1404     \toprule
1405     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1406     \cmidrule(lr){3-4}
1407     \cmidrule(lr){5-6}
1408     \cmidrule(l){7-8}
1409     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1410     \midrule
1411     PC & & -20.241 & 0.228 & -20.248 & 0.229 & -20.239 & 0.228 \\
1412     SP & 0.0 & 1.039 & 0.733 & 2.037 & 0.565 & 1.225 & 0.743 \\
1413     & 0.1 & 1.049 & 0.865 & 1.424 & 0.784 & 1.029 & 0.980 \\
1414     & 0.2 & 0.982 & 0.976 & 0.969 & 0.980 & 0.960 & 0.980 \\
1415     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.945 \\
1416     SF & 0.0 & 1.041 & 0.967 & 0.994 & 0.989 & 0.957 & 0.993 \\
1417     & 0.1 & 1.050 & 0.968 & 0.996 & 0.991 & 0.972 & 0.995 \\
1418     & 0.2 & 0.982 & 0.975 & 0.959 & 0.980 & 0.960 & 0.980 \\
1419     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.944 \\
1420     \midrule
1421     PC & & 0.795 & 0.000 & 0.792 & 0.000 & 0.793 & 0.000 \\
1422     SP & 0.0 & 0.916 & 0.829 & 1.086 & 0.791 & 1.010 & 0.936 \\
1423     & 0.1 & 0.958 & 0.917 & 1.049 & 0.943 & 1.001 & 0.995 \\
1424     & 0.2 & 0.981 & 0.981 & 0.982 & 0.984 & 0.981 & 0.984 \\
1425     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1426     SF & 0.0 & 1.002 & 0.983 & 0.997 & 0.994 & 0.991 & 0.997 \\
1427     & 0.1 & 1.003 & 0.984 & 0.996 & 0.995 & 0.993 & 0.997 \\
1428     & 0.2 & 0.983 & 0.980 & 0.981 & 0.984 & 0.981 & 0.984 \\
1429     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1430     \bottomrule
1431     \end{tabular}
1432     \label{tab:salt}
1433     \end{table}
1434    
1435     \begin{table}[htbp]
1436     \centering
1437     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1438     DISTRIBUTIONS OF THE FORCE VECTORS IN THE CRYSTALLINE SODIUM CHLORIDE
1439     SYSTEM}
1440    
1441     \footnotesize
1442     \begin{tabular}{@{} ccrrrrrr @{}}
1443     \\
1444     \toprule
1445     \toprule
1446     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1447     \cmidrule(lr){3-5}
1448     \cmidrule(l){6-8}
1449     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1450     \midrule
1451     PC & & 111.945 & 111.824 & 111.866 \\
1452     SP & 0.0 & 112.414 & 152.215 & 38.087 \\
1453     & 0.1 & 52.361 & 42.574 & 2.819 \\
1454     & 0.2 & 10.847 & 9.709 & 9.686 \\
1455     & 0.3 & 31.128 & 31.104 & 31.029 \\
1456     SF & 0.0 & 10.025 & 3.555 & 1.648 \\
1457     & 0.1 & 9.462 & 3.303 & 1.721 \\
1458     & 0.2 & 11.454 & 9.813 & 9.701 \\
1459     & 0.3 & 31.120 & 31.105 & 31.029 \\
1460     \bottomrule
1461     \end{tabular}
1462     \label{tab:saltAng}
1463     \end{table}
1464    
1465     The crystalline NaCl system is the most challenging test case for the
1466     pairwise summation methods, as evidenced by the results in tables
1467     \ref{tab:salt} and \ref{tab:saltAng}. The undamped and weakly damped
1468     {\sc sf} methods seem to be the best choices. These methods match well
1469     with {\sc spme} across the energy gap, force magnitude, and force
1470     directionality tests. The {\sc sp} method struggles in all cases,
1471     with the exception of good dynamics reproduction when using weak
1472     electrostatic damping with a large cutoff radius.
1473    
1474     The moderate electrostatic damping case is not as good as we would
1475     expect given the long-time dynamics results observed for this system
1476     (see section \ref{sec:LongTimeDynamics}). Since the data tabulated in
1477     tables \ref{tab:salt} and \ref{tab:saltAng} are a test of
1478     instantaneous dynamics, this indicates that good long-time dynamics
1479     comes in part at the expense of short-time dynamics.
1480    
1481     \subsection{0.11M NaCl Solution Results}
1482    
1483     In an effort to bridge the charged atomic and neutral molecular
1484     systems, Na$^+$ and Cl$^-$ ion charge defects were incorporated into
1485     the liquid water system. This low ionic strength system consists of 4
1486     ions in the 1000 SPC/E water solvent ($\approx$0.11 M). The results
1487     for the energy gap comparisons and the force and torque vector
1488     magnitude comparisons are shown in table \ref{tab:solnWeak}. The
1489     force and torque vector directionality results are displayed
1490     separately in table \ref{tab:solnWeakAng}, where the effect of
1491     group-based cutoffs and switching functions on the {\sc sp} and {\sc
1492     sf} potentials are investigated.
1493    
1494     \begin{table}[htbp]
1495     \centering
1496     \caption{REGRESSION RESULTS OF THE WEAK SODIUM CHLORIDE SOLUTION
1497     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1498     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1499    
1500     \footnotesize
1501     \begin{tabular}{@{} ccrrrrrr @{}}
1502     \\
1503     \toprule
1504     \toprule
1505     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1506     \cmidrule(lr){3-4}
1507     \cmidrule(lr){5-6}
1508     \cmidrule(l){7-8}
1509     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1510     \midrule
1511     PC & & 0.247 & 0.000 & -1.103 & 0.001 & 5.480 & 0.015 \\
1512     SP & 0.0 & 0.935 & 0.388 & 0.984 & 0.541 & 1.010 & 0.685 \\
1513     & 0.1 & 0.951 & 0.603 & 0.993 & 0.875 & 1.001 & 0.979 \\
1514     & 0.2 & 0.969 & 0.968 & 0.996 & 0.997 & 0.994 & 0.997 \\
1515     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1516     SF & 0.0 & 0.963 & 0.971 & 0.989 & 0.996 & 0.991 & 0.998 \\
1517     & 0.1 & 0.970 & 0.971 & 0.995 & 0.997 & 0.997 & 0.999 \\
1518     & 0.2 & 0.972 & 0.975 & 0.996 & 0.997 & 0.994 & 0.997 \\
1519     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1520     GSC & & 0.964 & 0.731 & 0.984 & 0.704 & 1.005 & 0.770 \\
1521     RF & & 0.968 & 0.605 & 0.974 & 0.541 & 1.014 & 0.614 \\
1522     \midrule
1523     PC & & 1.354 & 0.000 & -1.190 & 0.000 & -0.314 & 0.000 \\
1524     SP & 0.0 & 0.720 & 0.338 & 0.808 & 0.523 & 0.860 & 0.643 \\
1525     & 0.1 & 0.839 & 0.583 & 0.955 & 0.882 & 0.992 & 0.978 \\
1526     & 0.2 & 0.995 & 0.987 & 0.999 & 1.000 & 0.999 & 1.000 \\
1527     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1528     SF & 0.0 & 0.998 & 0.994 & 1.000 & 0.998 & 1.000 & 0.999 \\
1529     & 0.1 & 0.997 & 0.994 & 1.000 & 0.999 & 1.000 & 1.000 \\
1530     & 0.2 & 0.999 & 0.998 & 0.999 & 1.000 & 0.999 & 1.000 \\
1531     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1532     GSC & & 0.995 & 0.990 & 0.998 & 0.997 & 0.998 & 0.996 \\
1533     RF & & 0.998 & 0.993 & 0.999 & 0.998 & 0.999 & 0.996 \\
1534     \midrule
1535     PC & & 2.437 & 0.000 & -1.872 & 0.000 & 2.138 & 0.000 \\
1536     SP & 0.0 & 0.838 & 0.525 & 0.901 & 0.686 & 0.932 & 0.779 \\
1537     & 0.1 & 0.914 & 0.733 & 0.979 & 0.932 & 0.995 & 0.987 \\
1538     & 0.2 & 0.977 & 0.969 & 0.988 & 0.990 & 0.989 & 0.990 \\
1539     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1540     SF & 0.0 & 0.969 & 0.977 & 0.987 & 0.996 & 0.993 & 0.998 \\
1541     & 0.1 & 0.975 & 0.978 & 0.993 & 0.996 & 0.997 & 0.998 \\
1542     & 0.2 & 0.976 & 0.973 & 0.988 & 0.990 & 0.989 & 0.990 \\
1543     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1544     GSC & & 0.980 & 0.959 & 0.990 & 0.983 & 0.992 & 0.989 \\
1545     RF & & 0.984 & 0.975 & 0.996 & 0.995 & 0.998 & 0.998 \\
1546     \bottomrule
1547     \end{tabular}
1548     \label{tab:solnWeak}
1549     \end{table}
1550    
1551     \begin{table}[htbp]
1552     \centering
1553     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1554     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE WEAK SODIUM
1555     CHLORIDE SOLUTION SYSTEM}
1556    
1557     \footnotesize
1558     \begin{tabular}{@{} ccrrrrrr @{}}
1559     \\
1560     \toprule
1561     \toprule
1562     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1563     \cmidrule(lr){3-5}
1564     \cmidrule(l){6-8}
1565     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1566     \midrule
1567     PC & & 882.863 & 510.435 & 344.201 & 277.691 & 154.231 & 100.131 \\
1568     SP & 0.0 & 732.569 & 405.704 & 257.756 & 261.445 & 142.245 & 91.497 \\
1569     & 0.1 & 329.031 & 70.746 & 12.014 & 118.496 & 25.218 & 4.711 \\
1570     & 0.2 & 6.772 & 0.153 & 0.118 & 9.780 & 2.101 & 2.102 \\
1571     & 0.3 & 0.951 & 0.774 & 0.784 & 12.108 & 7.673 & 7.851 \\
1572     SF & 0.0 & 2.555 & 0.762 & 0.313 & 6.590 & 1.328 & 0.558 \\
1573     & 0.1 & 2.561 & 0.560 & 0.123 & 6.464 & 1.162 & 0.457 \\
1574     & 0.2 & 0.501 & 0.118 & 0.118 & 5.698 & 2.074 & 2.099 \\
1575     & 0.3 & 0.943 & 0.774 & 0.784 & 12.118 & 7.674 & 7.851 \\
1576     GSC & & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1577     RF & & 2.415 & 0.452 & 0.130 & 6.915 & 1.423 & 0.507 \\
1578     \midrule
1579     GSSP & 0.0 & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1580     & 0.1 & 2.251 & 0.324 & 0.064 & 7.628 & 1.639 & 0.497 \\
1581     & 0.2 & 0.590 & 0.118 & 0.116 & 6.080 & 2.096 & 2.103 \\
1582     & 0.3 & 0.953 & 0.759 & 0.780 & 12.347 & 7.683 & 7.849 \\
1583     GSSF & 0.0 & 1.541 & 0.301 & 0.096 & 6.407 & 1.316 & 0.496 \\
1584     & 0.1 & 1.541 & 0.237 & 0.050 & 6.356 & 1.202 & 0.457 \\
1585     & 0.2 & 0.568 & 0.118 & 0.116 & 6.166 & 2.105 & 2.105 \\
1586     & 0.3 & 0.954 & 0.759 & 0.780 & 12.337 & 7.684 & 7.849 \\
1587     \bottomrule
1588     \end{tabular}
1589     \label{tab:solnWeakAng}
1590     \end{table}
1591    
1592     Because this system is a perturbation of the pure liquid water system,
1593     comparisons are best drawn between these two sets. The {\sc sp} and
1594     {\sc sf} methods are not significantly affected by the inclusion of a
1595     few ions. The aspect of cutoff sphere neutralization aids in the
1596     smooth incorporation of these ions; thus, all of the observations
1597     regarding these methods carry over from section
1598     \ref{sec:WaterResults}. The differences between these systems are more
1599     visible for the {\sc rf} method. Though good force agreement is still
1600     maintained, the energy gaps show a significant increase in the scatter
1601     of the data.
1602    
1603     \subsection{1.1M NaCl Solution Results}
1604    
1605     The bridging of the charged atomic and neutral molecular systems was
1606     further developed by considering a high ionic strength system
1607     consisting of 40 ions in the 1000 SPC/E water solvent ($\approx$1.1
1608     M). The results for the energy gap comparisons and the force and
1609     torque vector magnitude comparisons are shown in table
1610     \ref{tab:solnStr}. The force and torque vector directionality
1611     results are displayed separately in table \ref{tab:solnStrAng}, where
1612     the effect of group-based cutoffs and switching functions on the {\sc
1613     sp} and {\sc sf} potentials are investigated.
1614    
1615     \begin{table}[htbp]
1616     \centering
1617     \caption{REGRESSION RESULTS OF THE STRONG SODIUM CHLORIDE SOLUTION
1618     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1619     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1620    
1621     \footnotesize
1622     \begin{tabular}{@{} ccrrrrrr @{}}
1623     \\
1624     \toprule
1625     \toprule
1626     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1627     \cmidrule(lr){3-4}
1628     \cmidrule(lr){5-6}
1629     \cmidrule(l){7-8}
1630     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1631     \midrule
1632     PC & & -0.081 & 0.000 & 0.945 & 0.001 & 0.073 & 0.000 \\
1633     SP & 0.0 & 0.978 & 0.469 & 0.996 & 0.672 & 0.975 & 0.668 \\
1634     & 0.1 & 0.944 & 0.645 & 0.997 & 0.886 & 0.991 & 0.978 \\
1635     & 0.2 & 0.873 & 0.896 & 0.985 & 0.993 & 0.980 & 0.993 \\
1636     & 0.3 & 0.831 & 0.860 & 0.960 & 0.979 & 0.955 & 0.977 \\
1637     SF & 0.0 & 0.858 & 0.905 & 0.985 & 0.970 & 0.990 & 0.998 \\
1638     & 0.1 & 0.865 & 0.907 & 0.992 & 0.974 & 0.994 & 0.999 \\
1639     & 0.2 & 0.862 & 0.894 & 0.985 & 0.993 & 0.980 & 0.993 \\
1640     & 0.3 & 0.831 & 0.859 & 0.960 & 0.979 & 0.955 & 0.977 \\
1641     GSC & & 1.985 & 0.152 & 0.760 & 0.031 & 1.106 & 0.062 \\
1642     RF & & 2.414 & 0.116 & 0.813 & 0.017 & 1.434 & 0.047 \\
1643     \midrule
1644     PC & & -7.028 & 0.000 & -9.364 & 0.000 & 0.925 & 0.865 \\
1645     SP & 0.0 & 0.701 & 0.319 & 0.909 & 0.773 & 0.861 & 0.665 \\
1646     & 0.1 & 0.824 & 0.565 & 0.970 & 0.930 & 0.990 & 0.979 \\
1647     & 0.2 & 0.988 & 0.981 & 0.995 & 0.998 & 0.991 & 0.998 \\
1648     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1649     SF & 0.0 & 0.993 & 0.988 & 0.992 & 0.984 & 0.998 & 0.999 \\
1650     & 0.1 & 0.993 & 0.989 & 0.993 & 0.986 & 0.998 & 1.000 \\
1651     & 0.2 & 0.993 & 0.992 & 0.995 & 0.998 & 0.991 & 0.998 \\
1652     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1653     GSC & & 0.964 & 0.897 & 0.970 & 0.917 & 0.925 & 0.865 \\
1654     RF & & 0.994 & 0.864 & 0.988 & 0.865 & 0.980 & 0.784 \\
1655     \midrule
1656     PC & & -2.212 & 0.000 & -0.588 & 0.000 & 0.953 & 0.925 \\
1657     SP & 0.0 & 0.800 & 0.479 & 0.930 & 0.804 & 0.924 & 0.759 \\
1658     & 0.1 & 0.883 & 0.694 & 0.976 & 0.942 & 0.993 & 0.986 \\
1659     & 0.2 & 0.952 & 0.943 & 0.980 & 0.984 & 0.980 & 0.983 \\
1660     & 0.3 & 0.914 & 0.909 & 0.943 & 0.948 & 0.944 & 0.946 \\
1661     SF & 0.0 & 0.945 & 0.953 & 0.980 & 0.984 & 0.991 & 0.998 \\
1662     & 0.1 & 0.951 & 0.954 & 0.987 & 0.986 & 0.995 & 0.998 \\
1663     & 0.2 & 0.951 & 0.946 & 0.980 & 0.984 & 0.980 & 0.983 \\
1664     & 0.3 & 0.914 & 0.908 & 0.943 & 0.948 & 0.944 & 0.946 \\
1665     GSC & & 0.882 & 0.818 & 0.939 & 0.902 & 0.953 & 0.925 \\
1666     RF & & 0.949 & 0.939 & 0.988 & 0.988 & 0.992 & 0.993 \\
1667     \bottomrule
1668     \end{tabular}
1669     \label{tab:solnStr}
1670     \end{table}
1671    
1672     \begin{table}[htbp]
1673     \centering
1674     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1675     OF THE FORCE AND TORQUE VECTORS IN THE STRONG SODIUM CHLORIDE SOLUTION
1676     SYSTEM}
1677    
1678     \footnotesize
1679     \begin{tabular}{@{} ccrrrrrr @{}}
1680     \\
1681     \toprule
1682     \toprule
1683     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1684     \cmidrule(lr){3-5}
1685     \cmidrule(l){6-8}
1686     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1687     \midrule
1688     PC & & 957.784 & 513.373 & 2.260 & 340.043 & 179.443 & 13.079 \\
1689     SP & 0.0 & 786.244 & 139.985 & 259.289 & 311.519 & 90.280 & 105.187 \\
1690     & 0.1 & 354.697 & 38.614 & 12.274 & 144.531 & 23.787 & 5.401 \\
1691     & 0.2 & 7.674 & 0.363 & 0.215 & 16.655 & 3.601 & 3.634 \\
1692     & 0.3 & 1.745 & 1.456 & 1.449 & 23.669 & 14.376 & 14.240 \\
1693     SF & 0.0 & 3.282 & 8.567 & 0.369 & 11.904 & 6.589 & 0.717 \\
1694     & 0.1 & 3.263 & 7.479 & 0.142 & 11.634 & 5.750 & 0.591 \\
1695     & 0.2 & 0.686 & 0.324 & 0.215 & 10.809 & 3.580 & 3.635 \\
1696     & 0.3 & 1.749 & 1.456 & 1.449 & 23.635 & 14.375 & 14.240 \\
1697     GSC & & 6.181 & 2.904 & 2.263 & 44.349 & 19.442 & 12.873 \\
1698     RF & & 3.891 & 0.847 & 0.323 & 18.628 & 3.995 & 2.072 \\
1699     \midrule
1700     GSSP & 0.0 & 6.197 & 2.929 & 2.290 & 44.441 & 19.442 & 12.873 \\
1701     & 0.1 & 4.688 & 1.064 & 0.260 & 31.208 & 6.967 & 2.303 \\
1702     & 0.2 & 1.021 & 0.218 & 0.213 & 14.425 & 3.629 & 3.649 \\
1703     & 0.3 & 1.752 & 1.454 & 1.451 & 23.540 & 14.390 & 14.245 \\
1704     GSSF & 0.0 & 2.494 & 0.546 & 0.217 & 16.391 & 3.230 & 1.613 \\
1705     & 0.1 & 2.448 & 0.429 & 0.106 & 16.390 & 2.827 & 1.159 \\
1706     & 0.2 & 0.899 & 0.214 & 0.213 & 13.542 & 3.583 & 3.645 \\
1707     & 0.3 & 1.752 & 1.454 & 1.451 & 23.587 & 14.390 & 14.245 \\
1708     \bottomrule
1709     \end{tabular}
1710     \label{tab:solnStrAng}
1711     \end{table}
1712    
1713     The {\sc rf} method struggles with the jump in ionic strength. The
1714     configuration energy differences degrade to unusable levels while the
1715     forces and torques show a more modest reduction in the agreement with
1716     {\sc spme}. The {\sc rf} method was designed for homogeneous systems,
1717     and this attribute is apparent in these results.
1718    
1719     The {\sc sp} and {\sc sf} methods require larger cutoffs to maintain
1720     their agreement with {\sc spme}. With these results, we still
1721     recommend undamped to moderate damping for the {\sc sf} method and
1722     moderate damping for the {\sc sp} method, both with cutoffs greater
1723     than 12\AA.
1724    
1725 chrisfen 2920 \subsection{6\AA\ Argon Sphere in SPC/E Water Results}
1726    
1727 chrisfen 2927 The final model system studied was a 6\AA\ sphere of Argon solvated
1728     by SPC/E water. This serves as a test case of a specifically sized
1729     electrostatic defect in a disordered molecular system. The results for
1730     the energy gap comparisons and the force and torque vector magnitude
1731     comparisons are shown in table \ref{tab:argon}. The force and torque
1732     vector directionality results are displayed separately in table
1733     \ref{tab:argonAng}, where the effect of group-based cutoffs and
1734     switching functions on the {\sc sp} and {\sc sf} potentials are
1735     investigated.
1736 chrisfen 2920
1737 chrisfen 2927 \begin{table}[htbp]
1738     \centering
1739     \caption{REGRESSION RESULTS OF THE 6\AA\ ARGON SPHERE IN LIQUID
1740     WATER SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR
1741     MAGNITUDES ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1742    
1743     \footnotesize
1744     \begin{tabular}{@{} ccrrrrrr @{}}
1745     \\
1746     \toprule
1747     \toprule
1748     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1749     \cmidrule(lr){3-4}
1750     \cmidrule(lr){5-6}
1751     \cmidrule(l){7-8}
1752     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1753     \midrule
1754     PC & & 2.320 & 0.008 & -0.650 & 0.001 & 3.848 & 0.029 \\
1755     SP & 0.0 & 1.053 & 0.711 & 0.977 & 0.820 & 0.974 & 0.882 \\
1756     & 0.1 & 1.032 & 0.846 & 0.989 & 0.965 & 0.992 & 0.994 \\
1757     & 0.2 & 0.993 & 0.995 & 0.982 & 0.998 & 0.986 & 0.998 \\
1758     & 0.3 & 0.968 & 0.995 & 0.954 & 0.992 & 0.961 & 0.994 \\
1759     SF & 0.0 & 0.982 & 0.996 & 0.992 & 0.999 & 0.993 & 1.000 \\
1760     & 0.1 & 0.987 & 0.996 & 0.996 & 0.999 & 0.997 & 1.000 \\
1761     & 0.2 & 0.989 & 0.998 & 0.984 & 0.998 & 0.989 & 0.998 \\
1762     & 0.3 & 0.971 & 0.995 & 0.957 & 0.992 & 0.965 & 0.994 \\
1763     GSC & & 1.002 & 0.983 & 0.992 & 0.973 & 0.996 & 0.971 \\
1764     RF & & 0.998 & 0.995 & 0.999 & 0.998 & 0.998 & 0.998 \\
1765     \midrule
1766     PC & & -36.559 & 0.002 & -44.917 & 0.004 & -52.945 & 0.006 \\
1767     SP & 0.0 & 0.890 & 0.786 & 0.927 & 0.867 & 0.949 & 0.909 \\
1768     & 0.1 & 0.942 & 0.895 & 0.984 & 0.974 & 0.997 & 0.995 \\
1769     & 0.2 & 0.999 & 0.997 & 1.000 & 1.000 & 1.000 & 1.000 \\
1770     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1771     SF & 0.0 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1772     & 0.1 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1773     & 0.2 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 \\
1774     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1775     GSC & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1776     RF & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1777     \midrule
1778     PC & & 1.984 & 0.000 & 0.012 & 0.000 & 1.357 & 0.000 \\
1779     SP & 0.0 & 0.850 & 0.552 & 0.907 & 0.703 & 0.938 & 0.793 \\
1780     & 0.1 & 0.924 & 0.755 & 0.980 & 0.936 & 0.995 & 0.988 \\
1781     & 0.2 & 0.985 & 0.983 & 0.986 & 0.988 & 0.987 & 0.988 \\
1782     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1783     SF & 0.0 & 0.977 & 0.989 & 0.987 & 0.995 & 0.992 & 0.998 \\
1784     & 0.1 & 0.982 & 0.989 & 0.992 & 0.996 & 0.997 & 0.998 \\
1785     & 0.2 & 0.984 & 0.987 & 0.986 & 0.987 & 0.987 & 0.988 \\
1786     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1787     GSC & & 0.995 & 0.981 & 0.999 & 0.990 & 1.000 & 0.993 \\
1788     RF & & 0.993 & 0.988 & 0.997 & 0.995 & 0.999 & 0.998 \\
1789     \bottomrule
1790     \end{tabular}
1791     \label{tab:argon}
1792     \end{table}
1793    
1794     \begin{table}[htbp]
1795     \centering
1796     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1797     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE 6\AA\ SPHERE OF
1798     ARGON IN LIQUID WATER SYSTEM}
1799    
1800     \footnotesize
1801     \begin{tabular}{@{} ccrrrrrr @{}}
1802     \\
1803     \toprule
1804     \toprule
1805     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1806     \cmidrule(lr){3-5}
1807     \cmidrule(l){6-8}
1808     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1809     \midrule
1810     PC & & 568.025 & 265.993 & 195.099 & 246.626 & 138.600 & 91.654 \\
1811     SP & 0.0 & 504.578 & 251.694 & 179.932 & 231.568 & 131.444 & 85.119 \\
1812     & 0.1 & 224.886 & 49.746 & 9.346 & 104.482 & 23.683 & 4.480 \\
1813     & 0.2 & 4.889 & 0.197 & 0.155 & 6.029 & 2.507 & 2.269 \\
1814     & 0.3 & 0.817 & 0.833 & 0.812 & 8.286 & 8.436 & 8.135 \\
1815     SF & 0.0 & 1.924 & 0.675 & 0.304 & 3.658 & 1.448 & 0.600 \\
1816     & 0.1 & 1.937 & 0.515 & 0.143 & 3.565 & 1.308 & 0.546 \\
1817     & 0.2 & 0.407 & 0.166 & 0.156 & 3.086 & 2.501 & 2.274 \\
1818     & 0.3 & 0.815 & 0.833 & 0.812 & 8.330 & 8.437 & 8.135 \\
1819     GSC & & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1820     RF & & 1.822 & 0.408 & 0.142 & 3.799 & 1.362 & 0.550 \\
1821     \midrule
1822     GSSP & 0.0 & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1823     & 0.1 & 1.652 & 0.309 & 0.087 & 4.197 & 1.401 & 0.590 \\
1824     & 0.2 & 0.465 & 0.165 & 0.153 & 3.323 & 2.529 & 2.273 \\
1825     & 0.3 & 0.813 & 0.825 & 0.816 & 8.316 & 8.447 & 8.132 \\
1826     GSSF & 0.0 & 1.173 & 0.292 & 0.113 & 3.452 & 1.347 & 0.583 \\
1827     & 0.1 & 1.166 & 0.240 & 0.076 & 3.381 & 1.281 & 0.575 \\
1828     & 0.2 & 0.459 & 0.165 & 0.153 & 3.430 & 2.542 & 2.273 \\
1829     & 0.3 & 0.814 & 0.825 & 0.816 & 8.325 & 8.447 & 8.132 \\
1830     \bottomrule
1831     \end{tabular}
1832     \label{tab:argonAng}
1833     \end{table}
1834    
1835     This system does not appear to show any significant deviations from
1836     the previously observed results. The {\sc sp} and {\sc sf} methods
1837 chrisfen 2957 have agreements similar to those observed in section
1838 chrisfen 2927 \ref{sec:WaterResults}. The only significant difference is the
1839     improvement in the configuration energy differences for the {\sc rf}
1840     method. This is surprising in that we are introducing an inhomogeneity
1841     to the system; however, this inhomogeneity is charge-neutral and does
1842     not result in charged cutoff spheres. The charge-neutrality of the
1843     cutoff spheres, which the {\sc sp} and {\sc sf} methods explicitly
1844     enforce, seems to play a greater role in the stability of the {\sc rf}
1845     method than the required homogeneity of the environment.
1846    
1847    
1848     \section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics}
1849    
1850 chrisfen 2918 Zahn {\it et al.} investigated the structure and dynamics of water
1851     using eqs. (\ref{eq:ZahnPot}) and
1852     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
1853     that a method similar (but not identical with) the damped {\sc sf}
1854     method resulted in properties very similar to those obtained when
1855     using the Ewald summation. The properties they studied (pair
1856     distribution functions, diffusion constants, and velocity and
1857     orientational correlation functions) may not be particularly sensitive
1858     to the long-range and collective behavior that governs the
1859     low-frequency behavior in crystalline systems. Additionally, the
1860     ionic crystals are the worst case scenario for the pairwise methods
1861     because they lack the reciprocal space contribution contained in the
1862     Ewald summation.
1863    
1864     We are using two separate measures to probe the effects of these
1865     alternative electrostatic methods on the dynamics in crystalline
1866     materials. For short- and intermediate-time dynamics, we are
1867     computing the velocity autocorrelation function, and for long-time
1868     and large length-scale collective motions, we are looking at the
1869     low-frequency portion of the power spectrum.
1870    
1871     \begin{figure}
1872     \centering
1873     \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
1874     \caption{Velocity autocorrelation functions of NaCl crystals at
1875 chrisfen 2927 1000K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \&
1876     0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2\AA$^{-1}$). The inset is
1877     a magnification of the area around the first minimum. The times to
1878     first collision are nearly identical, but differences can be seen in
1879     the peaks and troughs, where the undamped and weakly damped methods
1880     are stiffer than the moderately damped and {\sc spme} methods.}
1881 chrisfen 2918 \label{fig:vCorrPlot}
1882     \end{figure}
1883    
1884     The short-time decay of the velocity autocorrelation function through
1885     the first collision are nearly identical in figure
1886     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1887     how the methods differ. The undamped {\sc sf} method has deeper
1888     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1889     any of the other methods. As the damping parameter ($\alpha$) is
1890     increased, these peaks are smoothed out, and the {\sc sf} method
1891     approaches the {\sc spme} results. With $\alpha$ values of 0.2\AA$^{-1}$,
1892     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1893     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1894     and {\sc sp} potentials become nearly identical with increased
1895     damping. However, this appears to indicate that once damping is
1896     utilized, the details of the form of the potential (and forces)
1897     constructed out of the damped electrostatic interaction are less
1898     important.
1899    
1900 chrisfen 2927 \section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics}
1901 chrisfen 2918
1902     To evaluate how the differences between the methods affect the
1903     collective long-time motion, we computed power spectra from long-time
1904     traces of the velocity autocorrelation function. The power spectra for
1905     the best-performing alternative methods are shown in
1906     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1907     a cubic switching function between 40 and 50 ps was used to reduce the
1908     ringing resulting from data truncation. This procedure had no
1909     noticeable effect on peak location or magnitude.
1910    
1911     \begin{figure}
1912     \centering
1913     \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
1914     \caption{Power spectra obtained from the velocity auto-correlation
1915     functions of NaCl crystals at 1000K while using {\sc spme}, {\sc sf}
1916 chrisfen 2927 ($\alpha$ = 0, 0.1, \& 0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ =
1917     0.2\AA$^{-1}$). The inset shows the frequency region below 100
1918     cm$^{-1}$ to highlight where the spectra differ.}
1919 chrisfen 2918 \label{fig:methodPS}
1920     \end{figure}
1921    
1922     While the high frequency regions of the power spectra for the
1923     alternative methods are quantitatively identical with Ewald spectrum,
1924     the low frequency region shows how the summation methods differ.
1925     Considering the low-frequency inset (expanded in the upper frame of
1926     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1927     correlated motions are blue-shifted when using undamped or weakly
1928     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1929     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1930     correlated motion to the Ewald method (which has a convergence
1931     parameter of 0.3119\AA$^{-1}$). This weakening of the electrostatic
1932     interaction with increased damping explains why the long-ranged
1933     correlated motions are at lower frequencies for the moderately damped
1934     methods than for undamped or weakly damped methods.
1935    
1936     To isolate the role of the damping constant, we have computed the
1937     spectra for a single method ({\sc sf}) with a range of damping
1938     constants and compared this with the {\sc spme} spectrum.
1939     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1940     electrostatic damping red-shifts the lowest frequency phonon modes.
1941     However, even without any electrostatic damping, the {\sc sf} method
1942     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1943     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1944     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1945     Most} of the collective behavior in the crystal is accurately captured
1946     using the {\sc sf} method. Quantitative agreement with Ewald can be
1947     obtained using moderate damping in addition to the shifting at the
1948     cutoff distance.
1949    
1950     \begin{figure}
1951     \centering
1952     \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1953     \caption{Effect of damping on the two lowest-frequency phonon modes in
1954     the NaCl crystal at 1000K. The undamped shifted force ({\sc sf})
1955     method is off by less than 10 cm$^{-1}$, and increasing the
1956     electrostatic damping to 0.25\AA$^{-1}$ gives quantitative agreement
1957 chrisfen 2957 with the power spectrum obtained using the Ewald sum. Over-damping can
1958 chrisfen 2918 result in underestimates of frequencies of the long-wavelength
1959     motions.}
1960     \label{fig:dampInc}
1961     \end{figure}
1962    
1963 chrisfen 2957 \section{An Application: TIP5P-E Water}\label{sec:t5peApplied}
1964 chrisfen 2918
1965 chrisfen 2957 The above sections focused on the energetics and dynamics of a variety
1966     of systems when utilizing the {\sc sp} and {\sc sf} pairwise
1967     techniques. A unitary correlation with results obtained using the
1968     Ewald summation should result in a successful reproduction of both the
1969     static and dynamic properties of any selected system. To test this,
1970     we decided to calculate a series of properties for the TIP5P-E water
1971     model when using the {\sc sf} technique.
1972    
1973     The TIP5P-E water model is a variant of Mahoney and Jorgensen's
1974     five-point transferable intermolecular potential (TIP5P) model for
1975     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
1976     maximum anomaly present in liquid water near 4$^\circ$C. As with many
1977     previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
1978     and SPC/E), TIP5P was parametrized using a simple cutoff with no
1979     long-range electrostatic
1980     correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
1981     Without this correction, the pressure term on the central particle
1982     from the surroundings is missing. Because they expand to compensate
1983     for this added pressure term when this correction is included, systems
1984     composed of these particles tend to underpredict the density of water
1985     under standard conditions. When using any form of long-range
1986     electrostatic correction, it has become common practice to develop or
1987     utilize a reparametrized water model that corrects for this
1988     effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
1989     this practice and was optimized specifically for use with the Ewald
1990     summation.\cite{Rick04} In his publication, Rick preserved the
1991     geometry and point charge magnitudes in TIP5P and focused on altering
1992     the Lennard-Jones parameters to correct the density at
1993     298K.\cite{Rick04} With the density corrected, he compared common
1994     water properties for TIP5P-E using the Ewald sum with TIP5P using a
1995     9\AA\ cutoff.
1996    
1997     In the following sections, we compared these same water properties
1998     calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
1999     {\sc sf} technique. In the above evaluation of the pairwise
2000     techniques, we observed some flexibility in the choice of parameters.
2001     Because of this, the following comparisons include the {\sc sf}
2002     technique with a 12\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
2003     0.2\AA$^{-1}$, as well as a 9\AA\ cutoff with an $\alpha$ =
2004     0.2\AA$^{-1}$.
2005    
2006     \subsection{Density}\label{sec:t5peDensity}
2007    
2008     As stated previously, the property that prompted the development of
2009     TIP5P-E was the density at 1 atm. The density depends upon the
2010     internal pressure of the system in the $NPT$ ensemble, and the
2011     calculation of the pressure includes a components from both the
2012     kinetic energy and the virial. More specifically, the instantaneous
2013     molecular pressure ($P(t)$) is given by
2014     \begin{equation}
2015     P(t) = \frac{1}{\textrm{d}V}\sum_\mu
2016     \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
2017     + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
2018     \label{eq:MolecularPressure}
2019     \end{equation}
2020     where $V$ is the volume, $\mathbf{P}_{\mu}$ is the momentum of
2021     molecule $\mu$, $\mathbf{R}_\mu$ is the position of the center of mass
2022     ($M_\mu$) of molecule $\mu$, and $\mathbf{F}_{\mu i}$ is the force on
2023     atom $i$ of molecule $\mu$.\cite{Melchionna93} The virial term (the
2024     right term in the brackets of eq. \ref{eq:MolecularPressure}) is
2025     directly dependent on the interatomic forces. Since the {\sc sp}
2026     method does not modify the forces (see
2027     sec. \ref{sec:PairwiseDerivation}), the pressure using {\sc sp} will
2028     be identical to that obtained without an electrostatic correction.
2029     The {\sc sf} method does alter the virial component and, by way of the
2030     modified pressures, should provide densities more in line with those
2031     obtained using the Ewald summation.
2032    
2033     To compare densities, $NPT$ simulations were performed with the same
2034     temperatures as those selected by Rick in his Ewald summation
2035     simulations.\cite{Rick04} In order to improve statistics around the
2036     density maximum, 3ns trajectories were accumulated at 0, 12.5, and
2037     25$^\circ$C, while 2ns trajectories were obtained at all other
2038     temperatures. The average densities were calculated from the later
2039     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
2040     method for accumulating statistics, these sequences were spliced into
2041     200 segments to calculate the average density and standard deviation
2042     at each temperature.\cite{Mahoney00}
2043    
2044     \begin{figure}
2045     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
2046     \caption{Density versus temperature for the TIP5P-E water model when
2047     using the Ewald summation (Ref. \cite{Rick04}) and the {\sc sf} method
2048     with various parameters. The pressure term from the image-charge shell
2049     is larger than that provided by the reciprocal-space portion of the
2050     Ewald summation, leading to slightly lower densities. This effect is
2051     more visible with the 9\AA\ cutoff, where the image charges exert a
2052     greater force on the central particle. The error bars for the {\sc sf}
2053     methods show plus or minus the standard deviation of the density
2054     measurement at each temperature.}
2055     \label{fig:t5peDensities}
2056     \end{figure}
2057    
2058     Figure \ref{fig:t5peDensities} shows the densities calculated for
2059     TIP5P-E using differing electrostatic corrections overlaid on the
2060     experimental values.\cite{CRC80} The densities when using the {\sc sf}
2061     technique are close to, though typically lower than, those calculated
2062     while using the Ewald summation. These slightly reduced densities
2063     indicate that the pressure component from the image charges at
2064     R$_\textrm{c}$ is larger than that exerted by the reciprocal-space
2065     portion of the Ewald summation. Bringing the image charges closer to
2066     the central particle by choosing a 9\AA\ R$_\textrm{c}$ (rather than
2067     the preferred 12\AA\ R$_\textrm{c}$) increases the strength of their
2068     interactions, resulting in a further reduction of the densities.
2069    
2070     Because the strength of the image charge interactions has a noticable
2071     effect on the density, we would expect the use of electrostatic
2072     damping to also play a role in these calculations. Larger values of
2073     $\alpha$ weaken the pair-interactions; and since electrostatic damping
2074     is distance-dependent, force components from the image charges will be
2075     reduced more than those from particles close the the central
2076     charge. This effect is visible in figure \ref{fig:t5peDensities} with
2077     the damped {\sc sf} sums showing slightly higher densities; however,
2078     it is apparent that the choice of cutoff radius plays a much more
2079     important role in the resulting densities.
2080    
2081     As a final note, all of the above density calculations were performed
2082     with systems of 512 water molecules. Rick observed a system sized
2083     dependence of the computed densities when using the Ewald summation,
2084     most likely due to his tying of the convergence parameter to the box
2085     dimensions.\cite{Rick04} For systems of 256 water molecules, the
2086     calculated densities were on average 0.002 to 0.003 g/cm$^3$ lower. A
2087     system size of 256 molecules would force the use of a shorter
2088     R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
2089     lower the densities. Moving to larger systems, as long as the
2090     R$_\textrm{c}$ remains at a fixed value, we would expect the densities
2091     to remain constant.
2092    
2093     \subsection{Liquid Structure}\label{sec:t5peLiqStructure}
2094    
2095     A common function considered when developing and comparing water
2096     models is the oxygen-oxygen radial distribution function
2097     ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
2098     finding a pair of oxygen atoms some distance ($r$) apart relative to a
2099     random distribution at the same density.\cite{Allen87} It is
2100     calculated via
2101     \begin{equation}
2102     g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
2103     \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
2104     \label{eq:GOOofR}
2105     \end{equation}
2106     where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
2107     atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
2108     neutron scattering experiments through the oxygen-oxygen structure
2109     factor ($S_\textrm{OO}(k)$) by the following relationship:
2110     \begin{equation}
2111     S_\textrm{OO}(k) = 1 + 4\pi\rho
2112     \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
2113     \label{eq:SOOofK}
2114     \end{equation}
2115     Thus, $S_\textrm{OO}(k)$ is related to the Fourier transform of
2116     $g_\textrm{OO}(r)$.
2117    
2118     The expermentally determined $g_\textrm{OO}(r)$ for liquid water has
2119     been compared in great detail with the various common water models,
2120     and TIP5P was found to be in better agreement than other rigid,
2121     non-polarizable models.\cite{Sorenson00} This excellent agreement with
2122     experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
2123     check whether the choice of using the Ewald summation or the {\sc sf}
2124     technique alters the liquid structure, the $g_\textrm{OO}(r)$s at 298K
2125     and 1atm were determined for the systems compared in the previous
2126     section.
2127    
2128     \begin{figure}
2129     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
2130     \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298K and
2131     1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
2132     sf} technique with varying parameters. Even with the reduced densities
2133     using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
2134     identical.}
2135     \label{fig:t5peGofRs}
2136     \end{figure}
2137    
2138     The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
2139     sf} technique with a various parameters are overlaid on the
2140     $g_\textrm{OO}(r)$ while using the Ewald summation. The differences in
2141     density do not appear to have any effect on the liquid structure as
2142     the $g_\textrm{OO}(r)$s are indistinquishable. These results indicate
2143     that the $g_\textrm{OO}(r)$ is insensitive to the choice of
2144     electrostatic correction.
2145    
2146     \subsection{Thermodynamic Properties}
2147    
2148     \subsection{Dynamic Properties}
2149    
2150 chrisfen 2968 \section{Damping of Point Multipoles}
2151    
2152     \section{Damping and Dielectric Constants}
2153    
2154 chrisfen 2957 \section{Conclusions}\label{sec:PairwiseConclusions}
2155    
2156 chrisfen 2927 The above investigation of pairwise electrostatic summation techniques
2157     shows that there are viable and computationally efficient alternatives
2158     to the Ewald summation. These methods are derived from the damped and
2159     cutoff-neutralized Coulombic sum originally proposed by Wolf
2160     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
2161     method, reformulated above as eqs. (\ref{eq:DSFPot}) and
2162     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
2163     energetic and dynamic characteristics exhibited by simulations
2164     employing lattice summation techniques. The cumulative energy
2165     difference results showed the undamped {\sc sf} and moderately damped
2166     {\sc sp} methods produced results nearly identical to {\sc spme}.
2167     Similarly for the dynamic features, the undamped or moderately damped
2168     {\sc sf} and moderately damped {\sc sp} methods produce force and
2169     torque vector magnitude and directions very similar to the expected
2170     values. These results translate into long-time dynamic behavior
2171     equivalent to that produced in simulations using {\sc spme}.
2172    
2173     As in all purely-pairwise cutoff methods, these methods are expected
2174     to scale approximately {\it linearly} with system size, and they are
2175     easily parallelizable. This should result in substantial reductions
2176     in the computational cost of performing large simulations.
2177    
2178     Aside from the computational cost benefit, these techniques have
2179     applicability in situations where the use of the Ewald sum can prove
2180     problematic. Of greatest interest is their potential use in
2181     interfacial systems, where the unmodified lattice sum techniques
2182     artificially accentuate the periodicity of the system in an
2183     undesirable manner. There have been alterations to the standard Ewald
2184     techniques, via corrections and reformulations, to compensate for
2185     these systems; but the pairwise techniques discussed here require no
2186     modifications, making them natural tools to tackle these problems.
2187     Additionally, this transferability gives them benefits over other
2188     pairwise methods, like reaction field, because estimations of physical
2189     properties (e.g. the dielectric constant) are unnecessary.
2190    
2191     If a researcher is using Monte Carlo simulations of large chemical
2192     systems containing point charges, most structural features will be
2193     accurately captured using the undamped {\sc sf} method or the {\sc sp}
2194     method with an electrostatic damping of 0.2\AA$^{-1}$. These methods
2195     would also be appropriate for molecular dynamics simulations where the
2196     data of interest is either structural or short-time dynamical
2197     quantities. For long-time dynamics and collective motions, the safest
2198     pairwise method we have evaluated is the {\sc sf} method with an
2199     electrostatic damping between 0.2 and 0.25\AA$^{-1}$.
2200    
2201     We are not suggesting that there is any flaw with the Ewald sum; in
2202     fact, it is the standard by which these simple pairwise sums have been
2203     judged. However, these results do suggest that in the typical
2204     simulations performed today, the Ewald summation may no longer be
2205     required to obtain the level of accuracy most researchers have come to
2206     expect.
2207    
2208 chrisfen 2918 \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
2209    
2210     \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}
2211    
2212     \chapter{\label{chap:shapes}SPHERICAL HARMONIC APPROXIMATIONS FOR MOLECULAR
2213     SIMULATIONS}
2214    
2215     \chapter{\label{chap:conclusion}CONCLUSION}
2216    
2217     \backmatter
2218    
2219     \bibliographystyle{ndthesis}
2220 chrisfen 2927 \bibliography{dissertation}
2221 chrisfen 2918
2222     \end{document}
2223    
2224    
2225     \endinput