ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/electrostaticsChapter.tex
Revision: 2975
Committed: Tue Aug 22 18:04:10 2006 UTC (18 years ago) by chrisfen
Content type: application/x-tex
File size: 134654 byte(s)
Log Message:
adding the figures to cvs

File Contents

# User Rev Content
1 chrisfen 2973 \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION
2     TECHNIQUES}
3    
4     In molecular simulations, proper accumulation of the electrostatic
5     interactions is essential and is one of the most
6     computationally-demanding tasks. The common molecular mechanics force
7     fields represent atomic sites with full or partial charges protected
8     by Lennard-Jones (short range) interactions. This means that nearly
9     every pair interaction involves a calculation of charge-charge forces.
10     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
11     interactions quickly become the most expensive part of molecular
12     simulations. Historically, the electrostatic pair interaction would
13     not have decayed appreciably within the typical box lengths that could
14     be feasibly simulated. In the larger systems that are more typical of
15     modern simulations, large cutoffs should be used to incorporate
16     electrostatics correctly.
17    
18     There have been many efforts to address the proper and practical
19     handling of electrostatic interactions, and these have resulted in a
20     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
21     typically classified as implicit methods (i.e., continuum dielectrics,
22     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
23     (i.e., Ewald summations, interaction shifting or
24     truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
25     reaction field type methods, fast multipole
26     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
27     often preferred because they physically incorporate solvent molecules
28     in the system of interest, but these methods are sometimes difficult
29     to utilize because of their high computational cost.\cite{Roux99} In
30     addition to the computational cost, there have been some questions
31     regarding possible artifacts caused by the inherent periodicity of the
32     explicit Ewald summation.\cite{Tobias01}
33    
34     In this chapter, we focus on a new set of pairwise methods devised by
35     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
36     methods along with a few other mixed methods (i.e. reaction field) are
37     compared with the smooth particle mesh Ewald
38     sum,\cite{Onsager36,Essmann99} which is our reference method for
39     handling long-range electrostatic interactions. The new methods for
40     handling electrostatics have the potential to scale linearly with
41     increasing system size since they involve only a simple modification
42     to the direct pairwise sum. They also lack the added periodicity of
43     the Ewald sum, so they can be used for systems which are non-periodic
44     or which have one- or two-dimensional periodicity. Below, these
45     methods are evaluated using a variety of model systems to
46     establish their usability in molecular simulations.
47    
48     \section{The Ewald Sum}
49    
50     The complete accumulation of the electrostatic interactions in a system with
51     periodic boundary conditions (PBC) requires the consideration of the
52     effect of all charges within a (cubic) simulation box as well as those
53     in the periodic replicas,
54     \begin{equation}
55     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
56     \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
57     \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
58     \right],
59     \label{eq:PBCSum}
60     \end{equation}
61     where the sum over $\mathbf{n}$ is a sum over all periodic box
62     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
63     prime indicates $i = j$ are neglected for $\mathbf{n} =
64     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
65     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
66     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
67     $j$, and $\phi$ is the solution to Poisson's equation
68     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
69     charge-charge interactions). In the case of monopole electrostatics,
70     equation (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
71     non-neutral systems.
72    
73     The electrostatic summation problem was originally studied by Ewald
74     for the case of an infinite crystal.\cite{Ewald21}. The approach he
75     took was to convert this conditionally convergent sum into two
76     absolutely convergent summations: a short-ranged real-space summation
77     and a long-ranged reciprocal-space summation,
78     \begin{equation}
79     \begin{split}
80     V_\textrm{elec} = \frac{1}{2}&
81     \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
82     \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
83     {|\mathbf{r}_{ij}+\mathbf{n}|} \\
84     &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
85     \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
86     \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
87     &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
88     + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
89     \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
90     \end{split}
91     \label{eq:EwaldSum}
92     \end{equation}
93     where $\alpha$ is the damping or convergence parameter with units of
94     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
95     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
96     constant of the surrounding medium. The final two terms of
97     equation (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
98     for interacting with a surrounding dielectric.\cite{Allen87} This
99     dipolar term was neglected in early applications in molecular
100     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
101     Leeuw {\it et al.} to address situations where the unit cell has a
102     dipole moment which is magnified through replication of the periodic
103     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
104     system is said to be using conducting (or ``tin-foil'') boundary
105     conditions, $\epsilon_{\rm S} = \infty$. Figure
106     \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
107     time. Initially, due to the small system sizes that could be
108     simulated feasibly, the entire simulation box was replicated to
109     convergence. In more modern simulations, the systems have grown large
110     enough that a real-space cutoff could potentially give convergent
111     behavior. Indeed, it has been observed that with the choice of a
112     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
113     rapidly convergent and small relative to the real-space
114     portion.\cite{Karasawa89,Kolafa92}
115    
116     \begin{figure}
117     \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
118     \caption{The change in the need for the Ewald sum with
119     increasing computational power. A:~Initially, only small systems
120     could be studied, and the Ewald sum replicated the simulation box to
121     convergence. B:~Now, radial cutoff methods should be able to reach
122     convergence for the larger systems of charges that are common today.}
123     \label{fig:ewaldTime}
124     \end{figure}
125    
126     The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
127     convergence parameter $(\alpha)$ plays an important role in balancing
128     the computational cost between the direct and reciprocal-space
129     portions of the summation. The choice of this value allows one to
130     select whether the real-space or reciprocal space portion of the
131     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
132     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
133     $\alpha$ and thoughtful algorithm development, this cost can be
134     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
135     taken to reduce the cost of the Ewald summation even further is to set
136     $\alpha$ such that the real-space interactions decay rapidly, allowing
137     for a short spherical cutoff. Then the reciprocal space summation is
138     optimized. These optimizations usually involve utilization of the
139     fast Fourier transform (FFT),\cite{Hockney81} leading to the
140     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
141     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
142     methods, the cost of the reciprocal-space portion of the Ewald
143     summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
144     \log N)$.
145    
146     These developments and optimizations have made the use of the Ewald
147     summation routine in simulations with periodic boundary
148     conditions. However, in certain systems, such as vapor-liquid
149     interfaces and membranes, the intrinsic three-dimensional periodicity
150     can prove problematic. The Ewald sum has been reformulated to handle
151     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
152     methods are computationally expensive.\cite{Spohr97,Yeh99} More
153     recently, there have been several successful efforts toward reducing
154     the computational cost of 2-D lattice
155     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
156     bringing them more in line with the cost of the full 3-D summation.
157    
158     Several studies have recognized that the inherent periodicity in the
159     Ewald sum can also have an effect on three-dimensional
160     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
161     Solvated proteins are essentially kept at high concentration due to
162     the periodicity of the electrostatic summation method. In these
163     systems, the more compact folded states of a protein can be
164     artificially stabilized by the periodic replicas introduced by the
165     Ewald summation.\cite{Weber00} Thus, care must be taken when
166     considering the use of the Ewald summation where the assumed
167     periodicity would introduce spurious effects in the system dynamics.
168    
169    
170     \section{The Wolf and Zahn Methods}
171    
172     In a recent paper by Wolf \textit{et al.}, a procedure was outlined
173     for the accurate accumulation of electrostatic interactions in an
174     efficient pairwise fashion. This procedure lacks the inherent
175     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
176     observed that the electrostatic interaction is effectively
177     short-ranged in condensed phase systems and that neutralization of the
178     charge contained within the cutoff radius is crucial for potential
179     stability. They devised a pairwise summation method that ensures
180     charge neutrality and gives results similar to those obtained with the
181     Ewald summation. The resulting shifted Coulomb potential includes
182     image-charges subtracted out through placement on the cutoff sphere
183     and a distance-dependent damping function (identical to that seen in
184     the real-space portion of the Ewald sum) to aid convergence
185     \begin{equation}
186     V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
187     - \lim_{r_{ij}\rightarrow R_\textrm{c}}
188     \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
189     \label{eq:WolfPot}
190     \end{equation}
191     Equation (\ref{eq:WolfPot}) is essentially the common form of a shifted
192     potential. However, neutralizing the charge contained within each
193     cutoff sphere requires the placement of a self-image charge on the
194     surface of the cutoff sphere. This additional self-term in the total
195     potential enabled Wolf {\it et al.} to obtain excellent estimates of
196     Madelung energies for many crystals.
197    
198     In order to use their charge-neutralized potential in molecular
199     dynamics simulations, Wolf \textit{et al.} suggested taking the
200     derivative of this potential prior to evaluation of the limit. This
201     procedure gives an expression for the forces,
202     \begin{equation}
203     \begin{split}
204     F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
205     \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
206     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
207     \Biggr]\\
208     &-\Biggr[
209     \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
210     + \frac{2\alpha}{\pi^{1/2}}
211     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
212     \Biggr]\Biggr\},
213     \end{split}
214     \label{eq:WolfForces}
215     \end{equation}
216     that incorporates both image charges and damping of the electrostatic
217     interaction.
218    
219     More recently, Zahn \textit{et al.} investigated these potential and
220     force expressions for use in simulations involving water.\cite{Zahn02}
221     In their work, they pointed out that the forces and derivative of
222     the potential are not commensurate. Attempts to use both
223     equations (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
224     to poor energy conservation. They correctly observed that taking the
225     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
226     derivatives gives forces for a different potential energy function
227     than the one shown in equation (\ref{eq:WolfPot}).
228    
229     Zahn \textit{et al.} introduced a modified form of this summation
230     method as a way to use the technique in Molecular Dynamics
231     simulations. They proposed a new damped Coulomb potential,
232     \begin{equation}
233     \begin{split}
234     V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
235     \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
236     &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
237     + \frac{2\alpha}{\pi^{1/2}}
238     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
239     \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
240     \end{split}
241     \label{eq:ZahnPot}
242     \end{equation}
243     and showed that this potential does fairly well at capturing the
244     structural and dynamic properties of water compared the same
245     properties obtained using the Ewald sum.
246    
247     \section{Simple Forms for Pairwise Electrostatics}\label{sec:PairwiseDerivation}
248    
249     The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
250     al.} are constructed using two different (and separable) computational
251     tricks:
252    
253     \begin{enumerate}[itemsep=0pt]
254     \item shifting through the use of image charges, and
255     \item damping the electrostatic interaction.
256     \end{enumerate}
257     Wolf \textit{et al.} treated the development of their summation method
258     as a progressive application of these techniques,\cite{Wolf99} while
259     Zahn \textit{et al.} founded their damped Coulomb modification
260     (Eq. (\ref{eq:ZahnPot})) on the post-limit forces
261     (Eq. (\ref{eq:WolfForces})) which were derived using both techniques.
262     It is possible, however, to separate these tricks and study their
263     effects independently.
264    
265     Starting with the original observation that the effective range of the
266     electrostatic interaction in condensed phases is considerably less
267     than $r^{-1}$, either the cutoff sphere neutralization or the
268     distance-dependent damping technique could be used as a foundation for
269     a new pairwise summation method. Wolf \textit{et al.} made the
270     observation that charge neutralization within the cutoff sphere plays
271     a significant role in energy convergence; therefore we will begin our
272     analysis with the various shifted forms that maintain this charge
273     neutralization. We can evaluate the methods of Wolf {\it et al.} and
274     Zahn {\it et al.} by considering the standard shifted potential,
275     \begin{equation}
276     V_\textrm{SP}(r) = \begin{cases}
277     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
278     R_\textrm{c}
279     \end{cases},
280     \label{eq:shiftingPotForm}
281     \end{equation}
282     and shifted force,
283     \begin{equation}
284     V_\textrm{SF}(r) = \begin{cases}
285     v(r) - v_\textrm{c}
286     - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
287     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
288     \end{cases},
289     \label{eq:shiftingForm}
290     \end{equation}
291     functions where $v(r)$ is the unshifted form of the potential, and
292     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
293     that both the potential and the forces goes to zero at the cutoff
294     radius, while the Shifted Potential ({\sc sp}) form only ensures the
295     potential is smooth at the cutoff radius
296     ($R_\textrm{c}$).\cite{Allen87}
297    
298     The forces associated with the shifted potential are simply the forces
299     of the unshifted potential itself (when inside the cutoff sphere),
300     \begin{equation}
301     F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
302     \end{equation}
303     and are zero outside. Inside the cutoff sphere, the forces associated
304     with the shifted force form can be written,
305     \begin{equation}
306     F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
307     v(r)}{dr} \right)_{r=R_\textrm{c}}.
308     \end{equation}
309    
310     If the potential, $v(r)$, is taken to be the normal Coulomb potential,
311     \begin{equation}
312     v(r) = \frac{q_i q_j}{r},
313     \label{eq:Coulomb}
314     \end{equation}
315     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
316     al.}'s undamped prescription:
317     \begin{equation}
318     V_\textrm{SP}(r) =
319     q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
320     r\leqslant R_\textrm{c},
321     \label{eq:SPPot}
322     \end{equation}
323     with associated forces,
324     \begin{equation}
325     F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
326     \quad r\leqslant R_\textrm{c}.
327     \label{eq:SPForces}
328     \end{equation}
329     These forces are identical to the forces of the standard Coulomb
330     interaction, and cutting these off at $R_c$ was addressed by Wolf
331     \textit{et al.} as undesirable. They pointed out that the effect of
332     the image charges is neglected in the forces when this form is
333     used,\cite{Wolf99} thereby eliminating any benefit from the method in
334     molecular dynamics. Additionally, there is a discontinuity in the
335     forces at the cutoff radius which results in energy drift during MD
336     simulations.
337    
338     The shifted force ({\sc sf}) form using the normal Coulomb potential
339     will give,
340     \begin{equation}
341     V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
342     + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
343     \quad r\leqslant R_\textrm{c}.
344     \label{eq:SFPot}
345     \end{equation}
346     with associated forces,
347     \begin{equation}
348     F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
349     \quad r\leqslant R_\textrm{c}.
350     \label{eq:SFForces}
351     \end{equation}
352     This formulation has the benefits that there are no discontinuities at
353     the cutoff radius, while the neutralizing image charges are present in
354     both the energy and force expressions. It would be simple to add the
355     self-neutralizing term back when computing the total energy of the
356     system, thereby maintaining the agreement with the Madelung energies.
357     A side effect of this treatment is the alteration in the shape of the
358     potential that comes from the derivative term. Thus, a degree of
359     clarity about agreement with the empirical potential is lost in order
360     to gain functionality in dynamics simulations.
361    
362     Wolf \textit{et al.} originally discussed the energetics of the
363     shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
364     insufficient for accurate determination of the energy with reasonable
365     cutoff distances. The calculated Madelung energies fluctuated around
366     the expected value as the cutoff radius was increased, but the
367     oscillations converged toward the correct value.\cite{Wolf99} A
368     damping function was incorporated to accelerate the convergence; and
369     though alternative forms for the damping function could be
370     used,\cite{Jones56,Heyes81} the complimentary error function was
371     chosen to mirror the effective screening used in the Ewald summation.
372     Incorporating this error function damping into the simple Coulomb
373     potential,
374     \begin{equation}
375     v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
376     \label{eq:dampCoulomb}
377     \end{equation}
378     the shifted potential (Eq. (\ref{eq:SPPot})) becomes
379     \begin{equation}
380     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
381     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
382     \quad r\leqslant R_\textrm{c},
383     \label{eq:DSPPot}
384     \end{equation}
385     with associated forces,
386     \begin{equation}
387     F_{\textrm{DSP}}(r) = q_iq_j
388     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
389     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
390     \quad r\leqslant R_\textrm{c}.
391     \label{eq:DSPForces}
392     \end{equation}
393     Again, this damped shifted potential suffers from a
394     force-discontinuity at the cutoff radius, and the image charges play
395     no role in the forces. To remedy these concerns, one may derive a
396     {\sc sf} variant by including the derivative term in
397     equation (\ref{eq:shiftingForm}),
398     \begin{equation}
399     \begin{split}
400     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
401     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
402     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
403     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
404     + \frac{2\alpha}{\pi^{1/2}}
405     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
406     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
407     \quad r\leqslant R_\textrm{c}.
408     \label{eq:DSFPot}
409     \end{split}
410     \end{equation}
411     The derivative of the above potential will lead to the following forces,
412     \begin{equation}
413     \begin{split}
414     F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
415     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
416     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
417     &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
418     {R_{\textrm{c}}^2}
419     + \frac{2\alpha}{\pi^{1/2}}
420     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
421     \right)\Biggr{]}
422     \quad r\leqslant R_\textrm{c}.
423     \label{eq:DSFForces}
424     \end{split}
425     \end{equation}
426     If the damping parameter $(\alpha)$ is set to zero, the undamped case,
427     equations (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
428     recovered from equations (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
429    
430     This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
431     derived by Zahn \textit{et al.}; however, there are two important
432     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from equation
433     (\ref{eq:shiftingForm}) is equal to equation (\ref{eq:dampCoulomb})
434     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
435     in the Zahn potential, resulting in a potential discontinuity as
436     particles cross $R_\textrm{c}$. Second, the sign of the derivative
437     portion is different. The missing $v_\textrm{c}$ term would not
438     affect molecular dynamics simulations (although the computed energy
439     would be expected to have sudden jumps as particle distances crossed
440     $R_c$). The sign problem is a potential source of errors, however.
441     In fact, it introduces a discontinuity in the forces at the cutoff,
442     because the force function is shifted in the wrong direction and
443     doesn't cross zero at $R_\textrm{c}$.
444    
445     Equations (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
446     electrostatic summation method in which the potential and forces are
447     continuous at the cutoff radius and which incorporates the damping
448     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
449     this paper, we will evaluate exactly how good these methods ({\sc sp},
450     {\sc sf}, damping) are at reproducing the correct electrostatic
451     summation performed by the Ewald sum.
452    
453    
454     \section{Evaluating Pairwise Summation Techniques}
455    
456     In classical molecular mechanics simulations, there are two primary
457     techniques utilized to obtain information about the system of
458     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
459     techniques utilize pairwise summations of interactions between
460     particle sites, but they use these summations in different ways.
461    
462     In MC, the potential energy difference between configurations dictates
463     the progression of MC sampling. Going back to the origins of this
464     method, the acceptance criterion for the canonical ensemble laid out
465     by Metropolis \textit{et al.} states that a subsequent configuration
466     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
467     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
468     Maintaining the correct $\Delta E$ when using an alternate method for
469     handling the long-range electrostatics will ensure proper sampling
470     from the ensemble.
471    
472     In MD, the derivative of the potential governs how the system will
473     progress in time. Consequently, the force and torque vectors on each
474     body in the system dictate how the system evolves. If the magnitude
475     and direction of these vectors are similar when using alternate
476     electrostatic summation techniques, the dynamics in the short term
477     will be indistinguishable. Because error in MD calculations is
478     cumulative, one should expect greater deviation at longer times,
479     although methods which have large differences in the force and torque
480     vectors will diverge from each other more rapidly.
481    
482     \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
483    
484     The pairwise summation techniques (outlined in section
485     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
486     studying the energy differences between conformations. We took the
487     {\sc spme}-computed energy difference between two conformations to be the
488     correct behavior. An ideal performance by an alternative method would
489     reproduce these energy differences exactly (even if the absolute
490     energies calculated by the methods are different). Since none of the
491     methods provide exact energy differences, we used linear least squares
492     regressions of energy gap data to evaluate how closely the methods
493     mimicked the Ewald energy gaps. Unitary results for both the
494     correlation (slope) and correlation coefficient for these regressions
495     indicate perfect agreement between the alternative method and {\sc spme}.
496     Sample correlation plots for two alternate methods are shown in
497     Fig. \ref{fig:linearFit}.
498    
499     \begin{figure}
500     \centering
501 chrisfen 2975 \includegraphics[width = 3.5in]{./figures/dualLinear.pdf}
502 chrisfen 2973 \caption{Example least squares regressions of the configuration energy
503     differences for SPC/E water systems. The upper plot shows a data set
504     with a poor correlation coefficient ($R^2$), while the lower plot
505     shows a data set with a good correlation coefficient.}
506     \label{fig:linearFit}
507     \end{figure}
508    
509     Each of the seven system types (detailed in section \ref{sec:RepSims})
510     were represented using 500 independent configurations. Thus, each of
511     the alternative (non-Ewald) electrostatic summation methods was
512     evaluated using an accumulated 873,250 configurational energy
513     differences.
514    
515     Results and discussion for the individual analysis of each of the
516     system types appear in sections \ref{sec:IndividualResults}, while the
517     cumulative results over all the investigated systems appear below in
518     sections \ref{sec:EnergyResults}.
519    
520     \subsection{Molecular Dynamics and the Force and Torque
521     Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
522     (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
523     comparing the force and torque vectors with those obtained using the
524     reference Ewald summation ({\sc spme}). Both the magnitude and the
525     direction of these vectors on each of the bodies in the system were
526     analyzed. For the magnitude of these vectors, linear least squares
527     regression analyses were performed as described previously for
528     comparing $\Delta E$ values. Instead of a single energy difference
529     between two system configurations, we compared the magnitudes of the
530     forces (and torques) on each molecule in each configuration. For a
531     system of 1000 water molecules and 40 ions, there are 1040 force
532     vectors and 1000 torque vectors. With 500 configurations, this
533     results in 520,000 force and 500,000 torque vector comparisons.
534     Additionally, data from seven different system types was aggregated
535     before the comparison was made.
536    
537     The {\it directionality} of the force and torque vectors was
538     investigated through measurement of the angle ($\theta$) formed
539     between those computed from the particular method and those from {\sc spme},
540     \begin{equation}
541     \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
542     \cdot \hat{F}_\textrm{M}\right),
543     \end{equation}
544     where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
545     vector computed using method M. Each of these $\theta$ values was
546     accumulated in a distribution function and weighted by the area on the
547     unit sphere. Since this distribution is a measure of angular error
548     between two different electrostatic summation methods, there is no
549     {\it a priori} reason for the profile to adhere to any specific
550     shape. Thus, gaussian fits were used to measure the width of the
551     resulting distributions. The variance ($\sigma^2$) was extracted from
552     each of these fits and was used to compare distribution widths.
553     Values of $\sigma^2$ near zero indicate vector directions
554     indistinguishable from those calculated when using the reference
555     method ({\sc spme}).
556    
557     \subsection{Short-time Dynamics}
558    
559     The effects of the alternative electrostatic summation methods on the
560     short-time dynamics of charged systems were evaluated by considering a
561     NaCl crystal at a temperature of 1000 K. A subset of the best
562     performing pairwise methods was used in this comparison. The NaCl
563     crystal was chosen to avoid possible complications from the treatment
564     of orientational motion in molecular systems. All systems were
565     started with the same initial positions and velocities. Simulations
566     were performed under the microcanonical ensemble, and velocity
567     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
568     of the trajectories,
569     \begin{equation}
570     C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
571     \label{eq:vCorr}
572     \end{equation}
573     Velocity autocorrelation functions require detailed short time data,
574     thus velocity information was saved every 2fs over 10ps
575     trajectories. Because the NaCl crystal is composed of two different
576     atom types, the average of the two resulting velocity autocorrelation
577     functions was used for comparisons.
578    
579     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
580    
581     The effects of the same subset of alternative electrostatic methods on
582     the {\it long-time} dynamics of charged systems were evaluated using
583     the same model system (NaCl crystals at 1000K). The power spectrum
584     ($I(\omega)$) was obtained via Fourier transform of the velocity
585     autocorrelation function,
586     \begin{equation} I(\omega) =
587     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
588     \label{eq:powerSpec}
589     \end{equation}
590     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
591     NaCl crystal is composed of two different atom types, the average of
592     the two resulting power spectra was used for comparisons. Simulations
593     were performed under the microcanonical ensemble, and velocity
594     information was saved every 5fs over 100ps trajectories.
595    
596     \subsection{Representative Simulations}\label{sec:RepSims}
597     A variety of representative molecular simulations were analyzed to
598     determine the relative effectiveness of the pairwise summation
599     techniques in reproducing the energetics and dynamics exhibited by
600     {\sc spme}. We wanted to span the space of typical molecular
601     simulations (i.e. from liquids of neutral molecules to ionic
602     crystals), so the systems studied were:
603    
604     \begin{enumerate}[itemsep=0pt]
605     \item liquid water (SPC/E),\cite{Berendsen87}
606     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
607     \item NaCl crystals,
608     \item NaCl melts,
609     \item a low ionic strength solution of NaCl in water (0.11 M),
610     \item a high ionic strength solution of NaCl in water (1.1 M), and
611     \item a 6\AA\ radius sphere of Argon in water.
612     \end{enumerate}
613    
614     By utilizing the pairwise techniques (outlined in section
615     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
616     charged particles, and mixtures of the two, we hope to discern under
617     which conditions it will be possible to use one of the alternative
618     summation methodologies instead of the Ewald sum.
619    
620     For the solid and liquid water configurations, configurations were
621     taken at regular intervals from high temperature trajectories of 1000
622     SPC/E water molecules. Each configuration was equilibrated
623     independently at a lower temperature (300K for the liquid, 200K for
624     the crystal). The solid and liquid NaCl systems consisted of 500
625     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
626     these systems were selected and equilibrated in the same manner as the
627     water systems. In order to introduce measurable fluctuations in the
628     configuration energy differences, the crystalline simulations were
629     equilibrated at 1000K, near the $T_\textrm{m}$ for NaCl. The liquid
630     NaCl configurations needed to represent a fully disordered array of
631     point charges, so the high temperature of 7000K was selected for
632     equilibration. The ionic solutions were made by solvating 4 (or 40)
633     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
634     water positions were then randomly swapped, and the resulting
635     configurations were again equilibrated individually. Finally, for the
636     Argon / Water ``charge void'' systems, the identities of all the SPC/E
637     waters within 6\AA\ of the center of the equilibrated water
638     configurations were converted to argon.
639    
640     These procedures guaranteed us a set of representative configurations
641     from chemically-relevant systems sampled from appropriate
642     ensembles. Force field parameters for the ions and Argon were taken
643     from the force field utilized by {\sc oopse}.\cite{Meineke05}
644    
645     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
646     We compared the following alternative summation methods with results
647     from the reference method ({\sc spme}):
648    
649     \begin{enumerate}[itemsep=0pt]
650     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
651     and 0.3\AA$^{-1}$,
652     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
653     and 0.3\AA$^{-1}$,
654     \item reaction field with an infinite dielectric constant, and
655     \item an unmodified cutoff.
656     \end{enumerate}
657    
658     Group-based cutoffs with a fifth-order polynomial switching function
659     were utilized for the reaction field simulations. Additionally, we
660     investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
661     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
662     implementation of {\sc spme},\cite{Ponder87} while all other calculations
663     were performed using the {\sc oopse} molecular mechanics
664     package.\cite{Meineke05} All other portions of the energy calculation
665     (i.e. Lennard-Jones interactions) were handled in exactly the same
666     manner across all systems and configurations.
667    
668     The alternative methods were also evaluated with three different
669     cutoff radii (9, 12, and 15\AA). As noted previously, the
670     convergence parameter ($\alpha$) plays a role in the balance of the
671     real-space and reciprocal-space portions of the Ewald calculation.
672     Typical molecular mechanics packages set this to a value dependent on
673     the cutoff radius and a tolerance (typically less than $1 \times
674     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
675     increasing accuracy at the expense of computational time spent on the
676     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
677     The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
678     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
679     0.3119, and 0.2476\AA$^{-1}$ for cutoff radii of 9, 12, and 15\AA\
680     respectively.
681    
682     \section{Configuration Energy Difference Results}\label{sec:EnergyResults}
683     In order to evaluate the performance of the pairwise electrostatic
684     summation methods for Monte Carlo (MC) simulations, the energy
685     differences between configurations were compared to the values
686     obtained when using {\sc spme}. The results for the combined
687     regression analysis of all of the systems are shown in figure
688     \ref{fig:delE}.
689    
690     \begin{figure}
691     \centering
692     \includegraphics[width=4.75in]{./figures/delEplot.pdf}
693     \caption{Statistical analysis of the quality of configurational energy
694     differences for a given electrostatic method compared with the
695     reference Ewald sum. Results with a value equal to 1 (dashed line)
696     indicate $\Delta E$ values indistinguishable from those obtained using
697     {\sc spme}. Different values of the cutoff radius are indicated with
698     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
699     inverted triangles).}
700     \label{fig:delE}
701     \end{figure}
702    
703     The most striking feature of this plot is how well the Shifted Force
704     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
705     differences. For the undamped {\sc sf} method, and the
706     moderately-damped {\sc sp} methods, the results are nearly
707     indistinguishable from the Ewald results. The other common methods do
708     significantly less well.
709    
710     The unmodified cutoff method is essentially unusable. This is not
711     surprising since hard cutoffs give large energy fluctuations as atoms
712     or molecules move in and out of the cutoff
713     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
714     some degree by using group based cutoffs with a switching
715     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
716     significant improvement using the group-switched cutoff because the
717     salt and salt solution systems contain non-neutral groups. Section
718     \ref{sec:IndividualResults} includes results for systems comprised entirely
719     of neutral groups.
720    
721     For the {\sc sp} method, inclusion of electrostatic damping improves
722     the agreement with Ewald, and using an $\alpha$ of 0.2\AA $^{-1}$
723     shows an excellent correlation and quality of fit with the {\sc spme}
724     results, particularly with a cutoff radius greater than 12
725     \AA . Use of a larger damping parameter is more helpful for the
726     shortest cutoff shown, but it has a detrimental effect on simulations
727     with larger cutoffs.
728    
729     In the {\sc sf} sets, increasing damping results in progressively {\it
730     worse} correlation with Ewald. Overall, the undamped case is the best
731     performing set, as the correlation and quality of fits are
732     consistently superior regardless of the cutoff distance. The undamped
733     case is also less computationally demanding (because no evaluation of
734     the complementary error function is required).
735    
736     The reaction field results illustrates some of that method's
737     limitations, primarily that it was developed for use in homogeneous
738     systems; although it does provide results that are an improvement over
739     those from an unmodified cutoff.
740    
741     \section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults}
742    
743     Evaluation of pairwise methods for use in Molecular Dynamics
744     simulations requires consideration of effects on the forces and
745     torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
746     regression results for the force and torque vector magnitudes,
747     respectively. The data in these figures was generated from an
748     accumulation of the statistics from all of the system types.
749    
750     \begin{figure}
751     \centering
752     \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
753     \caption{Statistical analysis of the quality of the force vector
754     magnitudes for a given electrostatic method compared with the
755     reference Ewald sum. Results with a value equal to 1 (dashed line)
756     indicate force magnitude values indistinguishable from those obtained
757     using {\sc spme}. Different values of the cutoff radius are indicated with
758     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
759     inverted triangles).}
760     \label{fig:frcMag}
761     \end{figure}
762    
763     Again, it is striking how well the Shifted Potential and Shifted Force
764     methods are doing at reproducing the {\sc spme} forces. The undamped and
765     weakly-damped {\sc sf} method gives the best agreement with Ewald.
766     This is perhaps expected because this method explicitly incorporates a
767     smooth transition in the forces at the cutoff radius as well as the
768     neutralizing image charges.
769    
770     Figure \ref{fig:frcMag}, for the most part, parallels the results seen
771     in the previous $\Delta E$ section. The unmodified cutoff results are
772     poor, but using group based cutoffs and a switching function provides
773     an improvement much more significant than what was seen with $\Delta
774     E$.
775    
776     With moderate damping and a large enough cutoff radius, the {\sc sp}
777     method is generating usable forces. Further increases in damping,
778     while beneficial for simulations with a cutoff radius of 9\AA\ , is
779     detrimental to simulations with larger cutoff radii.
780    
781     The reaction field results are surprisingly good, considering the poor
782     quality of the fits for the $\Delta E$ results. There is still a
783     considerable degree of scatter in the data, but the forces correlate
784     well with the Ewald forces in general. We note that the reaction
785     field calculations do not include the pure NaCl systems, so these
786     results are partly biased towards conditions in which the method
787     performs more favorably.
788    
789     \begin{figure}
790     \centering
791     \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
792     \caption{Statistical analysis of the quality of the torque vector
793     magnitudes for a given electrostatic method compared with the
794     reference Ewald sum. Results with a value equal to 1 (dashed line)
795     indicate torque magnitude values indistinguishable from those obtained
796     using {\sc spme}. Different values of the cutoff radius are indicated with
797     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
798     inverted triangles).}
799     \label{fig:trqMag}
800     \end{figure}
801    
802     Molecular torques were only available from the systems which contained
803     rigid molecules (i.e. the systems containing water). The data in
804     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
805    
806     Torques appear to be much more sensitive to charges at a longer
807     distance. The striking feature in comparing the new electrostatic
808     methods with {\sc spme} is how much the agreement improves with increasing
809     cutoff radius. Again, the weakly damped and undamped {\sc sf} method
810     appears to be reproducing the {\sc spme} torques most accurately.
811    
812     Water molecules are dipolar, and the reaction field method reproduces
813     the effect of the surrounding polarized medium on each of the
814     molecular bodies. Therefore it is not surprising that reaction field
815     performs best of all of the methods on molecular torques.
816    
817     \section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults}
818    
819     It is clearly important that a new electrostatic method can reproduce
820     the magnitudes of the force and torque vectors obtained via the Ewald
821     sum. However, the {\it directionality} of these vectors will also be
822     vital in calculating dynamical quantities accurately. Force and
823     torque directionalities were investigated by measuring the angles
824     formed between these vectors and the same vectors calculated using
825     {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
826     variance ($\sigma^2$) of the Gaussian fits of the angle error
827     distributions of the combined set over all system types.
828    
829     \begin{figure}
830     \centering
831     \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
832     \caption{Statistical analysis of the width of the angular distribution
833     that the force and torque vectors from a given electrostatic method
834     make with their counterparts obtained using the reference Ewald sum.
835     Results with a variance ($\sigma^2$) equal to zero (dashed line)
836     indicate force and torque directions indistinguishable from those
837     obtained using {\sc spme}. Different values of the cutoff radius are
838     indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
839     and 15\AA\ = inverted triangles).}
840     \label{fig:frcTrqAng}
841     \end{figure}
842    
843     Both the force and torque $\sigma^2$ results from the analysis of the
844     total accumulated system data are tabulated in figure
845     \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
846     sp}) method would be essentially unusable for molecular dynamics
847     unless the damping function is added. The Shifted Force ({\sc sf})
848     method, however, is generating force and torque vectors which are
849     within a few degrees of the Ewald results even with weak (or no)
850     damping.
851    
852     All of the sets (aside from the over-damped case) show the improvement
853     afforded by choosing a larger cutoff radius. Increasing the cutoff
854     from 9 to 12\AA\ typically results in a halving of the width of the
855     distribution, with a similar improvement when going from 12 to 15
856     \AA .
857    
858     The undamped {\sc sf}, group-based cutoff, and reaction field methods
859     all do equivalently well at capturing the direction of both the force
860     and torque vectors. Using the electrostatic damping improves the
861     angular behavior significantly for the {\sc sp} and moderately for the
862     {\sc sf} methods. Over-damping is detrimental to both methods. Again
863     it is important to recognize that the force vectors cover all
864     particles in all seven systems, while torque vectors are only
865     available for neutral molecular groups. Damping is more beneficial to
866     charged bodies, and this observation is investigated further in
867     section \ref{sec:IndividualResults}.
868    
869     Although not discussed previously, group based cutoffs can be applied
870     to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
871     will reintroduce small discontinuities at the cutoff radius, but the
872     effects of these can be minimized by utilizing a switching function.
873     Though there are no significant benefits or drawbacks observed in
874     $\Delta E$ and the force and torque magnitudes when doing this, there
875     is a measurable improvement in the directionality of the forces and
876     torques. Table \ref{tab:groupAngle} shows the angular variances
877     obtained both without (N) and with (Y) group based cutoffs and a
878     switching function. Note that the $\alpha$ values have units of
879     \AA$^{-1}$ and the variance values have units of degrees$^2$. The
880     {\sc sp} (with an $\alpha$ of 0.2\AA$^{-1}$ or smaller) shows much
881     narrower angular distributions when using group-based cutoffs. The
882     {\sc sf} method likewise shows improvement in the undamped and lightly
883     damped cases.
884    
885     \begin{table}
886     \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
887     THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
888     GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
889     THE REFERENCE EWALD SUMMATION}
890    
891     \footnotesize
892     \begin{center}
893     \begin{tabular}{@{} ccrrrrrrrr @{}}
894     \toprule
895     \toprule
896     & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
897     Force} \\
898     \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
899     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
900     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
901    
902     \midrule
903    
904     9\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
905     & \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} \\
906     12\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
907     & \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} \\
908     15\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
909     & \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} \\
910    
911     \midrule
912    
913     9\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
914     & \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} \\
915     12\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
916     & \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} \\
917     15\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
918     & \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} \\
919    
920     \bottomrule
921     \end{tabular}
922     \end{center}
923     \label{tab:groupAngle}
924     \end{table}
925    
926     One additional trend in table \ref{tab:groupAngle} is that the
927     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
928     increases, something that is more obvious with group-based cutoffs.
929     The complimentary error function inserted into the potential weakens
930     the electrostatic interaction as the value of $\alpha$ is increased.
931     However, at larger values of $\alpha$, it is possible to over-damp the
932     electrostatic interaction and to remove it completely. Kast
933     \textit{et al.} developed a method for choosing appropriate $\alpha$
934     values for these types of electrostatic summation methods by fitting
935     to $g(r)$ data, and their methods indicate optimal values of 0.34,
936     0.25, and 0.16\AA$^{-1}$ for cutoff values of 9, 12, and 15\AA\
937     respectively.\cite{Kast03} These appear to be reasonable choices to
938     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
939     these findings, choices this high would introduce error in the
940     molecular torques, particularly for the shorter cutoffs. Based on our
941     observations, empirical damping up to 0.2\AA$^{-1}$ is beneficial,
942     but damping may be unnecessary when using the {\sc sf} method.
943    
944     \section{Individual System Analysis Results}\label{sec:IndividualResults}
945    
946     The combined results of the previous sections show how the pairwise
947     methods compare to the Ewald summation in the general sense over all
948     of the system types. It is also useful to consider each of the
949     studied systems in an individual fashion, so that we can identify
950     conditions that are particularly difficult for a selected pairwise
951     method to address. This allows us to further establish the limitations
952     of these pairwise techniques. Below, the energy difference, force
953     vector, and torque vector analyses are presented on an individual
954     system basis.
955    
956     \subsection{SPC/E Water Results}\label{sec:WaterResults}
957    
958     The first system considered was liquid water at 300K using the SPC/E
959     model of water.\cite{Berendsen87} The results for the energy gap
960     comparisons and the force and torque vector magnitude comparisons are
961     shown in table \ref{tab:spce}. The force and torque vector
962     directionality results are displayed separately in table
963     \ref{tab:spceAng}, where the effect of group-based cutoffs and
964     switching functions on the {\sc sp} and {\sc sf} potentials are also
965     investigated. In all of the individual results table, the method
966     abbreviations are as follows:
967    
968     \begin{itemize}[itemsep=0pt]
969     \item PC = Pure Cutoff,
970     \item SP = Shifted Potential,
971     \item SF = Shifted Force,
972     \item GSC = Group Switched Cutoff,
973     \item RF = Reaction Field (where $\varepsilon \approx\infty$),
974     \item GSSP = Group Switched Shifted Potential, and
975     \item GSSF = Group Switched Shifted Force.
976     \end{itemize}
977    
978     \begin{table}[htbp]
979     \centering
980     \caption{REGRESSION RESULTS OF THE LIQUID WATER SYSTEM FOR THE
981     $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it middle})
982     AND TORQUE VECTOR MAGNITUDES ({\it lower})}
983    
984     \footnotesize
985     \begin{tabular}{@{} ccrrrrrr @{}}
986     \toprule
987     \toprule
988     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
989     \cmidrule(lr){3-4}
990     \cmidrule(lr){5-6}
991     \cmidrule(l){7-8}
992     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
993     \midrule
994     PC & & 3.046 & 0.002 & -3.018 & 0.002 & 4.719 & 0.005 \\
995     SP & 0.0 & 1.035 & 0.218 & 0.908 & 0.313 & 1.037 & 0.470 \\
996     & 0.1 & 1.021 & 0.387 & 0.965 & 0.752 & 1.006 & 0.947 \\
997     & 0.2 & 0.997 & 0.962 & 1.001 & 0.994 & 0.994 & 0.996 \\
998     & 0.3 & 0.984 & 0.980 & 0.997 & 0.985 & 0.982 & 0.987 \\
999     SF & 0.0 & 0.977 & 0.974 & 0.996 & 0.992 & 0.991 & 0.997 \\
1000     & 0.1 & 0.983 & 0.974 & 1.001 & 0.994 & 0.996 & 0.998 \\
1001     & 0.2 & 0.992 & 0.989 & 1.001 & 0.995 & 0.994 & 0.996 \\
1002     & 0.3 & 0.984 & 0.980 & 0.996 & 0.985 & 0.982 & 0.987 \\
1003     GSC & & 0.918 & 0.862 & 0.852 & 0.756 & 0.801 & 0.700 \\
1004     RF & & 0.971 & 0.958 & 0.975 & 0.987 & 0.959 & 0.983 \\
1005     \midrule
1006     PC & & -1.647 & 0.000 & -0.127 & 0.000 & -0.979 & 0.000 \\
1007     SP & 0.0 & 0.735 & 0.368 & 0.813 & 0.537 & 0.865 & 0.659 \\
1008     & 0.1 & 0.850 & 0.612 & 0.956 & 0.887 & 0.992 & 0.979 \\
1009     & 0.2 & 0.996 & 0.989 & 1.000 & 1.000 & 1.000 & 1.000 \\
1010     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1011     SF & 0.0 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 0.999 \\
1012     & 0.1 & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1013     & 0.2 & 0.999 & 0.998 & 1.000 & 1.000 & 1.000 & 1.000 \\
1014     & 0.3 & 0.996 & 0.998 & 0.997 & 0.998 & 0.996 & 0.998 \\
1015     GSC & & 0.998 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1016     RF & & 0.999 & 0.995 & 1.000 & 0.999 & 1.000 & 1.000 \\
1017     \midrule
1018     PC & & 2.387 & 0.000 & 0.183 & 0.000 & 1.282 & 0.000 \\
1019     SP & 0.0 & 0.847 & 0.543 & 0.904 & 0.694 & 0.935 & 0.786 \\
1020     & 0.1 & 0.922 & 0.749 & 0.980 & 0.934 & 0.996 & 0.988 \\
1021     & 0.2 & 0.987 & 0.985 & 0.989 & 0.992 & 0.990 & 0.993 \\
1022     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1023     SF & 0.0 & 0.978 & 0.990 & 0.988 & 0.997 & 0.993 & 0.999 \\
1024     & 0.1 & 0.983 & 0.991 & 0.993 & 0.997 & 0.997 & 0.999 \\
1025     & 0.2 & 0.986 & 0.989 & 0.989 & 0.992 & 0.990 & 0.993 \\
1026     & 0.3 & 0.965 & 0.973 & 0.967 & 0.975 & 0.967 & 0.976 \\
1027     GSC & & 0.995 & 0.981 & 0.999 & 0.991 & 1.001 & 0.994 \\
1028     RF & & 0.993 & 0.989 & 0.998 & 0.996 & 1.000 & 0.999 \\
1029     \bottomrule
1030     \end{tabular}
1031     \label{tab:spce}
1032     \end{table}
1033    
1034     \begin{table}[htbp]
1035     \centering
1036     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1037     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE LIQUID WATER
1038     SYSTEM}
1039    
1040     \footnotesize
1041     \begin{tabular}{@{} ccrrrrrr @{}}
1042     \toprule
1043     \toprule
1044     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1045     \cmidrule(lr){3-5}
1046     \cmidrule(l){6-8}
1047     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1048     \midrule
1049     PC & & 783.759 & 481.353 & 332.677 & 248.674 & 144.382 & 98.535 \\
1050     SP & 0.0 & 659.440 & 380.699 & 250.002 & 235.151 & 134.661 & 88.135 \\
1051     & 0.1 & 293.849 & 67.772 & 11.609 & 105.090 & 23.813 & 4.369 \\
1052     & 0.2 & 5.975 & 0.136 & 0.094 & 5.553 & 1.784 & 1.536 \\
1053     & 0.3 & 0.725 & 0.707 & 0.693 & 7.293 & 6.933 & 6.748 \\
1054     SF & 0.0 & 2.238 & 0.713 & 0.292 & 3.290 & 1.090 & 0.416 \\
1055     & 0.1 & 2.238 & 0.524 & 0.115 & 3.184 & 0.945 & 0.326 \\
1056     & 0.2 & 0.374 & 0.102 & 0.094 & 2.598 & 1.755 & 1.537 \\
1057     & 0.3 & 0.721 & 0.707 & 0.693 & 7.322 & 6.933 & 6.748 \\
1058     GSC & & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1059     RF & & 2.091 & 0.403 & 0.113 & 3.583 & 1.071 & 0.399 \\
1060     \midrule
1061     GSSP & 0.0 & 2.431 & 0.614 & 0.274 & 5.135 & 2.133 & 1.339 \\
1062     & 0.1 & 1.879 & 0.291 & 0.057 & 3.983 & 1.117 & 0.370 \\
1063     & 0.2 & 0.443 & 0.103 & 0.093 & 2.821 & 1.794 & 1.532 \\
1064     & 0.3 & 0.728 & 0.694 & 0.692 & 7.387 & 6.942 & 6.748 \\
1065     GSSF & 0.0 & 1.298 & 0.270 & 0.083 & 3.098 & 0.992 & 0.375 \\
1066     & 0.1 & 1.296 & 0.210 & 0.044 & 3.055 & 0.922 & 0.330 \\
1067     & 0.2 & 0.433 & 0.104 & 0.093 & 2.895 & 1.797 & 1.532 \\
1068     & 0.3 & 0.728 & 0.694 & 0.692 & 7.410 & 6.942 & 6.748 \\
1069     \bottomrule
1070     \end{tabular}
1071     \label{tab:spceAng}
1072     \end{table}
1073    
1074     The water results parallel the combined results seen in sections
1075     \ref{sec:EnergyResults} through \ref{sec:FTDirResults}. There is good
1076     agreement with {\sc spme} in both energetic and dynamic behavior when
1077     using the {\sc sf} method with and without damping. The {\sc sp}
1078     method does well with an $\alpha$ around 0.2\AA$^{-1}$, particularly
1079     with cutoff radii greater than 12\AA. Over-damping the electrostatics
1080     reduces the agreement between both these methods and {\sc spme}.
1081    
1082     The pure cutoff ({\sc pc}) method performs poorly, again mirroring the
1083     observations from the combined results. In contrast to these results, however, the use of a switching function and group
1084     based cutoffs greatly improves the results for these neutral water
1085     molecules. The group switched cutoff ({\sc gsc}) does not mimic the
1086     energetics of {\sc spme} as well as the {\sc sp} (with moderate
1087     damping) and {\sc sf} methods, but the dynamics are quite good. The
1088     switching functions correct discontinuities in the potential and
1089     forces, leading to these improved results. Such improvements with the
1090     use of a switching function have been recognized in previous
1091     studies,\cite{Andrea83,Steinbach94} and this proves to be a useful
1092     tactic for stably incorporating local area electrostatic effects.
1093    
1094     The reaction field ({\sc rf}) method simply extends upon the results
1095     observed in the {\sc gsc} case. Both methods are similar in form
1096     (i.e. neutral groups, switching function), but {\sc rf} incorporates
1097     an added effect from the external dielectric. This similarity
1098     translates into the same good dynamic results and improved energetic
1099     agreement with {\sc spme}. Though this agreement is not to the level
1100     of the moderately damped {\sc sp} and {\sc sf} methods, these results
1101     show how incorporating some implicit properties of the surroundings
1102     (i.e. $\epsilon_\textrm{S}$) can improve the solvent depiction.
1103    
1104     As a final note for the liquid water system, use of group cutoffs and a
1105     switching function leads to noticeable improvements in the {\sc sp}
1106     and {\sc sf} methods, primarily in directionality of the force and
1107     torque vectors (table \ref{tab:spceAng}). The {\sc sp} method shows
1108     significant narrowing of the angle distribution when using little to
1109     no damping and only modest improvement for the recommended conditions
1110     ($\alpha = 0.2$\AA$^{-1}$ and $R_\textrm{c}~\geqslant~12$\AA). The
1111     {\sc sf} method shows modest narrowing across all damping and cutoff
1112     ranges of interest. When over-damping these methods, group cutoffs and
1113     the switching function do not improve the force and torque
1114     directionalities.
1115    
1116     \subsection{SPC/E Ice I$_\textrm{c}$ Results}\label{sec:IceResults}
1117    
1118     In addition to the disordered molecular system above, the ordered
1119     molecular system of ice I$_\textrm{c}$ was also considered. Ice
1120     polymorph could have been used to fit this role; however, ice
1121     I$_\textrm{c}$ was chosen because it can form an ideal periodic
1122     lattice with the same number of water molecules used in the disordered
1123     liquid state case. The results for the energy gap comparisons and the
1124     force and torque vector magnitude comparisons are shown in table
1125     \ref{tab:ice}. The force and torque vector directionality results are
1126     displayed separately in table \ref{tab:iceAng}, where the effect of
1127     group-based cutoffs and switching functions on the {\sc sp} and {\sc
1128     sf} potentials are also displayed.
1129    
1130     \begin{table}[htbp]
1131     \centering
1132     \caption{REGRESSION RESULTS OF THE ICE I$_\textrm{c}$ SYSTEM FOR
1133     $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES ({\it
1134     middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1135    
1136     \footnotesize
1137     \begin{tabular}{@{} ccrrrrrr @{}}
1138     \toprule
1139     \toprule
1140     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1141     \cmidrule(lr){3-4}
1142     \cmidrule(lr){5-6}
1143     \cmidrule(l){7-8}
1144     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1145     \midrule
1146     PC & & 19.897 & 0.047 & -29.214 & 0.048 & -3.771 & 0.001 \\
1147     SP & 0.0 & -0.014 & 0.000 & 2.135 & 0.347 & 0.457 & 0.045 \\
1148     & 0.1 & 0.321 & 0.017 & 1.490 & 0.584 & 0.886 & 0.796 \\
1149     & 0.2 & 0.896 & 0.872 & 1.011 & 0.998 & 0.997 & 0.999 \\
1150     & 0.3 & 0.983 & 0.997 & 0.992 & 0.997 & 0.991 & 0.997 \\
1151     SF & 0.0 & 0.943 & 0.979 & 1.048 & 0.978 & 0.995 & 0.999 \\
1152     & 0.1 & 0.948 & 0.979 & 1.044 & 0.983 & 1.000 & 0.999 \\
1153     & 0.2 & 0.982 & 0.997 & 0.969 & 0.960 & 0.997 & 0.999 \\
1154     & 0.3 & 0.985 & 0.997 & 0.961 & 0.961 & 0.991 & 0.997 \\
1155     GSC & & 0.983 & 0.985 & 0.966 & 0.994 & 1.003 & 0.999 \\
1156     RF & & 0.924 & 0.944 & 0.990 & 0.996 & 0.991 & 0.998 \\
1157     \midrule
1158     PC & & -4.375 & 0.000 & 6.781 & 0.000 & -3.369 & 0.000 \\
1159     SP & 0.0 & 0.515 & 0.164 & 0.856 & 0.426 & 0.743 & 0.478 \\
1160     & 0.1 & 0.696 & 0.405 & 0.977 & 0.817 & 0.974 & 0.964 \\
1161     & 0.2 & 0.981 & 0.980 & 1.001 & 1.000 & 1.000 & 1.000 \\
1162     & 0.3 & 0.996 & 0.998 & 0.997 & 0.999 & 0.997 & 0.999 \\
1163     SF & 0.0 & 0.991 & 0.995 & 1.003 & 0.998 & 0.999 & 1.000 \\
1164     & 0.1 & 0.992 & 0.995 & 1.003 & 0.998 & 1.000 & 1.000 \\
1165     & 0.2 & 0.998 & 0.998 & 0.981 & 0.962 & 1.000 & 1.000 \\
1166     & 0.3 & 0.996 & 0.998 & 0.976 & 0.957 & 0.997 & 0.999 \\
1167     GSC & & 0.997 & 0.996 & 0.998 & 0.999 & 1.000 & 1.000 \\
1168     RF & & 0.988 & 0.989 & 1.000 & 0.999 & 1.000 & 1.000 \\
1169     \midrule
1170     PC & & -6.367 & 0.000 & -3.552 & 0.000 & -3.447 & 0.000 \\
1171     SP & 0.0 & 0.643 & 0.409 & 0.833 & 0.607 & 0.961 & 0.805 \\
1172     & 0.1 & 0.791 & 0.683 & 0.957 & 0.914 & 1.000 & 0.989 \\
1173     & 0.2 & 0.974 & 0.991 & 0.993 & 0.998 & 0.993 & 0.998 \\
1174     & 0.3 & 0.976 & 0.992 & 0.977 & 0.992 & 0.977 & 0.992 \\
1175     SF & 0.0 & 0.979 & 0.997 & 0.992 & 0.999 & 0.994 & 1.000 \\
1176     & 0.1 & 0.984 & 0.997 & 0.996 & 0.999 & 0.998 & 1.000 \\
1177     & 0.2 & 0.991 & 0.997 & 0.974 & 0.958 & 0.993 & 0.998 \\
1178     & 0.3 & 0.977 & 0.992 & 0.956 & 0.948 & 0.977 & 0.992 \\
1179     GSC & & 0.999 & 0.997 & 0.996 & 0.999 & 1.002 & 1.000 \\
1180     RF & & 0.994 & 0.997 & 0.997 & 0.999 & 1.000 & 1.000 \\
1181     \bottomrule
1182     \end{tabular}
1183     \label{tab:ice}
1184     \end{table}
1185    
1186     \begin{table}[htbp]
1187     \centering
1188     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1189     OF THE FORCE AND TORQUE VECTORS IN THE ICE I$_\textrm{c}$ SYSTEM}
1190    
1191     \footnotesize
1192     \begin{tabular}{@{} ccrrrrrr @{}}
1193     \toprule
1194     \toprule
1195     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque
1196     $\sigma^2$} \\
1197     \cmidrule(lr){3-5}
1198     \cmidrule(l){6-8}
1199     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1200     \midrule
1201     PC & & 2128.921 & 603.197 & 715.579 & 329.056 & 221.397 & 81.042 \\
1202     SP & 0.0 & 1429.341 & 470.320 & 447.557 & 301.678 & 197.437 & 73.840 \\
1203     & 0.1 & 590.008 & 107.510 & 18.883 & 118.201 & 32.472 & 3.599 \\
1204     & 0.2 & 10.057 & 0.105 & 0.038 & 2.875 & 0.572 & 0.518 \\
1205     & 0.3 & 0.245 & 0.260 & 0.262 & 2.365 & 2.396 & 2.327 \\
1206     SF & 0.0 & 1.745 & 1.161 & 0.212 & 1.135 & 0.426 & 0.155 \\
1207     & 0.1 & 1.721 & 0.868 & 0.082 & 1.118 & 0.358 & 0.118 \\
1208     & 0.2 & 0.201 & 0.040 & 0.038 & 0.786 & 0.555 & 0.518 \\
1209     & 0.3 & 0.241 & 0.260 & 0.262 & 2.368 & 2.400 & 2.327 \\
1210     GSC & & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1211     RF & & 2.887 & 0.217 & 0.107 & 1.006 & 0.281 & 0.085 \\
1212     \midrule
1213     GSSP & 0.0 & 1.483 & 0.261 & 0.099 & 0.926 & 0.295 & 0.095 \\
1214     & 0.1 & 1.341 & 0.123 & 0.037 & 0.835 & 0.234 & 0.085 \\
1215     & 0.2 & 0.558 & 0.040 & 0.037 & 0.823 & 0.557 & 0.519 \\
1216     & 0.3 & 0.250 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1217     GSSF & 0.0 & 2.124 & 0.132 & 0.069 & 0.919 & 0.263 & 0.099 \\
1218     & 0.1 & 2.165 & 0.101 & 0.035 & 0.895 & 0.244 & 0.096 \\
1219     & 0.2 & 0.706 & 0.040 & 0.037 & 0.870 & 0.559 & 0.519 \\
1220     & 0.3 & 0.251 & 0.251 & 0.259 & 2.387 & 2.395 & 2.328 \\
1221     \bottomrule
1222     \end{tabular}
1223     \label{tab:iceAng}
1224     \end{table}
1225    
1226     Highly ordered systems are a difficult test for the pairwise methods
1227     in that they lack the implicit periodicity of the Ewald summation. As
1228     expected, the energy gap agreement with {\sc spme} is reduced for the
1229     {\sc sp} and {\sc sf} methods with parameters that were ideal for the
1230     disordered liquid system. Moving to higher $R_\textrm{c}$ helps
1231     improve the agreement, though at an increase in computational cost.
1232     The dynamics of this crystalline system (both in magnitude and
1233     direction) are little affected. Both methods still reproduce the Ewald
1234     behavior with the same parameter recommendations from the previous
1235     section.
1236    
1237     It is also worth noting that {\sc rf} exhibits improved energy gap
1238     results over the liquid water system. One possible explanation is
1239     that the ice I$_\textrm{c}$ crystal is ordered such that the net
1240     dipole moment of the crystal is zero. With $\epsilon_\textrm{S} =
1241     \infty$, the reaction field incorporates this structural organization
1242     by actively enforcing a zeroed dipole moment within each cutoff
1243     sphere.
1244    
1245     \subsection{NaCl Melt Results}\label{sec:SaltMeltResults}
1246    
1247     A high temperature NaCl melt was tested to gauge the accuracy of the
1248     pairwise summation methods in a disordered system of charges. The
1249     results for the energy gap comparisons and the force vector magnitude
1250     comparisons are shown in table \ref{tab:melt}. The force vector
1251     directionality results are displayed separately in table
1252     \ref{tab:meltAng}.
1253    
1254     \begin{table}[htbp]
1255     \centering
1256     \caption{REGRESSION RESULTS OF THE MOLTEN SODIUM CHLORIDE SYSTEM FOR
1257     $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES ({\it
1258     lower})}
1259    
1260     \footnotesize
1261     \begin{tabular}{@{} ccrrrrrr @{}}
1262     \toprule
1263     \toprule
1264     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1265     \cmidrule(lr){3-4}
1266     \cmidrule(lr){5-6}
1267     \cmidrule(l){7-8}
1268     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1269     \midrule
1270     PC & & -0.008 & 0.000 & -0.049 & 0.005 & -0.136 & 0.020 \\
1271     SP & 0.0 & 0.928 & 0.996 & 0.931 & 0.998 & 0.950 & 0.999 \\
1272     & 0.1 & 0.977 & 0.998 & 0.998 & 1.000 & 0.997 & 1.000 \\
1273     & 0.2 & 0.960 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1274     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1275     SF & 0.0 & 0.996 & 1.000 & 0.995 & 1.000 & 0.997 & 1.000 \\
1276     & 0.1 & 1.021 & 1.000 & 1.024 & 1.000 & 1.007 & 1.000 \\
1277     & 0.2 & 0.966 & 1.000 & 0.813 & 0.996 & 0.811 & 0.954 \\
1278     & 0.3 & 0.671 & 0.994 & 0.439 & 0.929 & 0.535 & 0.831 \\
1279     \midrule
1280     PC & & 1.103 & 0.000 & 0.989 & 0.000 & 0.802 & 0.000 \\
1281     SP & 0.0 & 0.973 & 0.981 & 0.975 & 0.988 & 0.979 & 0.992 \\
1282     & 0.1 & 0.987 & 0.992 & 0.993 & 0.998 & 0.997 & 0.999 \\
1283     & 0.2 & 0.993 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1284     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1285     SF & 0.0 & 0.996 & 0.997 & 0.997 & 0.999 & 0.998 & 1.000 \\
1286     & 0.1 & 1.000 & 0.997 & 1.001 & 0.999 & 1.000 & 1.000 \\
1287     & 0.2 & 0.994 & 0.996 & 0.985 & 0.988 & 0.986 & 0.981 \\
1288     & 0.3 & 0.956 & 0.956 & 0.940 & 0.912 & 0.948 & 0.929 \\
1289     \bottomrule
1290     \end{tabular}
1291     \label{tab:melt}
1292     \end{table}
1293    
1294     \begin{table}[htbp]
1295     \centering
1296     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1297     OF THE FORCE VECTORS IN THE MOLTEN SODIUM CHLORIDE SYSTEM}
1298    
1299     \footnotesize
1300     \begin{tabular}{@{} ccrrrrrr @{}}
1301     \toprule
1302     \toprule
1303     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1304     \cmidrule(lr){3-5}
1305     \cmidrule(l){6-8}
1306     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1307     \midrule
1308     PC & & 13.294 & 8.035 & 5.366 \\
1309     SP & 0.0 & 13.316 & 8.037 & 5.385 \\
1310     & 0.1 & 5.705 & 1.391 & 0.360 \\
1311     & 0.2 & 2.415 & 7.534 & 13.927 \\
1312     & 0.3 & 23.769 & 67.306 & 57.252 \\
1313     SF & 0.0 & 1.693 & 0.603 & 0.256 \\
1314     & 0.1 & 1.687 & 0.653 & 0.272 \\
1315     & 0.2 & 2.598 & 7.523 & 13.930 \\
1316     & 0.3 & 23.734 & 67.305 & 57.252 \\
1317     \bottomrule
1318     \end{tabular}
1319     \label{tab:meltAng}
1320     \end{table}
1321    
1322     The molten NaCl system shows more sensitivity to the electrostatic
1323     damping than the water systems. The most noticeable point is that the
1324     undamped {\sc sf} method does very well at replicating the {\sc spme}
1325     configurational energy differences and forces. Light damping appears
1326     to minimally improve the dynamics, but this comes with a deterioration
1327     of the energy gap results. In contrast, this light damping improves
1328     the {\sc sp} energy gaps and forces. Moderate and heavy electrostatic
1329     damping reduce the agreement with {\sc spme} for both methods. From
1330     these observations, the undamped {\sc sf} method is the best choice
1331     for disordered systems of charges.
1332    
1333     \subsection{NaCl Crystal Results}\label{sec:SaltCrystalResults}
1334    
1335     Similar to the use of ice I$_\textrm{c}$ to investigate the role of
1336     order in molecular systems on the effectiveness of the pairwise
1337     methods, the 1000K NaCl crystal system was used to investigate the
1338     accuracy of the pairwise summation methods in an ordered system of
1339     charged particles. The results for the energy gap comparisons and the
1340     force vector magnitude comparisons are shown in table \ref{tab:salt}.
1341     The force vector directionality results are displayed separately in
1342     table \ref{tab:saltAng}.
1343    
1344     \begin{table}[htbp]
1345     \centering
1346     \caption{REGRESSION RESULTS OF THE CRYSTALLINE SODIUM CHLORIDE
1347     SYSTEM FOR $\Delta E$ VALUES ({\it upper}) AND FORCE VECTOR MAGNITUDES
1348     ({\it lower})}
1349    
1350     \footnotesize
1351     \begin{tabular}{@{} ccrrrrrr @{}}
1352     \toprule
1353     \toprule
1354     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1355     \cmidrule(lr){3-4}
1356     \cmidrule(lr){5-6}
1357     \cmidrule(l){7-8}
1358     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1359     \midrule
1360     PC & & -20.241 & 0.228 & -20.248 & 0.229 & -20.239 & 0.228 \\
1361     SP & 0.0 & 1.039 & 0.733 & 2.037 & 0.565 & 1.225 & 0.743 \\
1362     & 0.1 & 1.049 & 0.865 & 1.424 & 0.784 & 1.029 & 0.980 \\
1363     & 0.2 & 0.982 & 0.976 & 0.969 & 0.980 & 0.960 & 0.980 \\
1364     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.945 \\
1365     SF & 0.0 & 1.041 & 0.967 & 0.994 & 0.989 & 0.957 & 0.993 \\
1366     & 0.1 & 1.050 & 0.968 & 0.996 & 0.991 & 0.972 & 0.995 \\
1367     & 0.2 & 0.982 & 0.975 & 0.959 & 0.980 & 0.960 & 0.980 \\
1368     & 0.3 & 0.873 & 0.944 & 0.872 & 0.945 & 0.872 & 0.944 \\
1369     \midrule
1370     PC & & 0.795 & 0.000 & 0.792 & 0.000 & 0.793 & 0.000 \\
1371     SP & 0.0 & 0.916 & 0.829 & 1.086 & 0.791 & 1.010 & 0.936 \\
1372     & 0.1 & 0.958 & 0.917 & 1.049 & 0.943 & 1.001 & 0.995 \\
1373     & 0.2 & 0.981 & 0.981 & 0.982 & 0.984 & 0.981 & 0.984 \\
1374     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1375     SF & 0.0 & 1.002 & 0.983 & 0.997 & 0.994 & 0.991 & 0.997 \\
1376     & 0.1 & 1.003 & 0.984 & 0.996 & 0.995 & 0.993 & 0.997 \\
1377     & 0.2 & 0.983 & 0.980 & 0.981 & 0.984 & 0.981 & 0.984 \\
1378     & 0.3 & 0.950 & 0.952 & 0.950 & 0.953 & 0.950 & 0.953 \\
1379     \bottomrule
1380     \end{tabular}
1381     \label{tab:salt}
1382     \end{table}
1383    
1384     \begin{table}[htbp]
1385     \centering
1386     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1387     DISTRIBUTIONS OF THE FORCE VECTORS IN THE CRYSTALLINE SODIUM CHLORIDE
1388     SYSTEM}
1389    
1390     \footnotesize
1391     \begin{tabular}{@{} ccrrrrrr @{}}
1392     \toprule
1393     \toprule
1394     & & \multicolumn{3}{c}{Force $\sigma^2$} \\
1395     \cmidrule(lr){3-5}
1396     \cmidrule(l){6-8}
1397     Method & $\alpha$ & 9\AA & 12\AA & 15\AA \\
1398     \midrule
1399     PC & & 111.945 & 111.824 & 111.866 \\
1400     SP & 0.0 & 112.414 & 152.215 & 38.087 \\
1401     & 0.1 & 52.361 & 42.574 & 2.819 \\
1402     & 0.2 & 10.847 & 9.709 & 9.686 \\
1403     & 0.3 & 31.128 & 31.104 & 31.029 \\
1404     SF & 0.0 & 10.025 & 3.555 & 1.648 \\
1405     & 0.1 & 9.462 & 3.303 & 1.721 \\
1406     & 0.2 & 11.454 & 9.813 & 9.701 \\
1407     & 0.3 & 31.120 & 31.105 & 31.029 \\
1408     \bottomrule
1409     \end{tabular}
1410     \label{tab:saltAng}
1411     \end{table}
1412    
1413     The crystalline NaCl system is the most challenging test case for the
1414     pairwise summation methods, as evidenced by the results in tables
1415     \ref{tab:salt} and \ref{tab:saltAng}. The undamped and weakly damped
1416     {\sc sf} methods seem to be the best choices. These methods match well
1417     with {\sc spme} across the energy gap, force magnitude, and force
1418     directionality tests. The {\sc sp} method struggles in all cases,
1419     with the exception of good dynamics reproduction when using weak
1420     electrostatic damping with a large cutoff radius.
1421    
1422     The moderate electrostatic damping case is not as good as we would
1423     expect given the long-time dynamics results observed for this system
1424     (see section \ref{sec:LongTimeDynamics}). Since the data tabulated in
1425     tables \ref{tab:salt} and \ref{tab:saltAng} are a test of
1426     instantaneous dynamics, this indicates that good long-time dynamics
1427     comes in part at the expense of short-time dynamics.
1428    
1429     \subsection{0.11M NaCl Solution Results}
1430    
1431     In an effort to bridge the charged atomic and neutral molecular
1432     systems, Na$^+$ and Cl$^-$ ion charge defects were incorporated into
1433     the liquid water system. This low ionic strength system consists of 4
1434     ions in the 1000 SPC/E water solvent ($\approx$0.11 M). The results
1435     for the energy gap comparisons and the force and torque vector
1436     magnitude comparisons are shown in table \ref{tab:solnWeak}. The
1437     force and torque vector directionality results are displayed
1438     separately in table \ref{tab:solnWeakAng}, where the effect of
1439     group-based cutoffs and switching functions on the {\sc sp} and {\sc
1440     sf} potentials are investigated.
1441    
1442     \begin{table}[htbp]
1443     \centering
1444     \caption{REGRESSION RESULTS OF THE WEAK SODIUM CHLORIDE SOLUTION
1445     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1446     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1447    
1448     \footnotesize
1449     \begin{tabular}{@{} ccrrrrrr @{}}
1450     \toprule
1451     \toprule
1452     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1453     \cmidrule(lr){3-4}
1454     \cmidrule(lr){5-6}
1455     \cmidrule(l){7-8}
1456     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1457     \midrule
1458     PC & & 0.247 & 0.000 & -1.103 & 0.001 & 5.480 & 0.015 \\
1459     SP & 0.0 & 0.935 & 0.388 & 0.984 & 0.541 & 1.010 & 0.685 \\
1460     & 0.1 & 0.951 & 0.603 & 0.993 & 0.875 & 1.001 & 0.979 \\
1461     & 0.2 & 0.969 & 0.968 & 0.996 & 0.997 & 0.994 & 0.997 \\
1462     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1463     SF & 0.0 & 0.963 & 0.971 & 0.989 & 0.996 & 0.991 & 0.998 \\
1464     & 0.1 & 0.970 & 0.971 & 0.995 & 0.997 & 0.997 & 0.999 \\
1465     & 0.2 & 0.972 & 0.975 & 0.996 & 0.997 & 0.994 & 0.997 \\
1466     & 0.3 & 0.955 & 0.966 & 0.984 & 0.992 & 0.978 & 0.991 \\
1467     GSC & & 0.964 & 0.731 & 0.984 & 0.704 & 1.005 & 0.770 \\
1468     RF & & 0.968 & 0.605 & 0.974 & 0.541 & 1.014 & 0.614 \\
1469     \midrule
1470     PC & & 1.354 & 0.000 & -1.190 & 0.000 & -0.314 & 0.000 \\
1471     SP & 0.0 & 0.720 & 0.338 & 0.808 & 0.523 & 0.860 & 0.643 \\
1472     & 0.1 & 0.839 & 0.583 & 0.955 & 0.882 & 0.992 & 0.978 \\
1473     & 0.2 & 0.995 & 0.987 & 0.999 & 1.000 & 0.999 & 1.000 \\
1474     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1475     SF & 0.0 & 0.998 & 0.994 & 1.000 & 0.998 & 1.000 & 0.999 \\
1476     & 0.1 & 0.997 & 0.994 & 1.000 & 0.999 & 1.000 & 1.000 \\
1477     & 0.2 & 0.999 & 0.998 & 0.999 & 1.000 & 0.999 & 1.000 \\
1478     & 0.3 & 0.995 & 0.996 & 0.996 & 0.998 & 0.996 & 0.998 \\
1479     GSC & & 0.995 & 0.990 & 0.998 & 0.997 & 0.998 & 0.996 \\
1480     RF & & 0.998 & 0.993 & 0.999 & 0.998 & 0.999 & 0.996 \\
1481     \midrule
1482     PC & & 2.437 & 0.000 & -1.872 & 0.000 & 2.138 & 0.000 \\
1483     SP & 0.0 & 0.838 & 0.525 & 0.901 & 0.686 & 0.932 & 0.779 \\
1484     & 0.1 & 0.914 & 0.733 & 0.979 & 0.932 & 0.995 & 0.987 \\
1485     & 0.2 & 0.977 & 0.969 & 0.988 & 0.990 & 0.989 & 0.990 \\
1486     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1487     SF & 0.0 & 0.969 & 0.977 & 0.987 & 0.996 & 0.993 & 0.998 \\
1488     & 0.1 & 0.975 & 0.978 & 0.993 & 0.996 & 0.997 & 0.998 \\
1489     & 0.2 & 0.976 & 0.973 & 0.988 & 0.990 & 0.989 & 0.990 \\
1490     & 0.3 & 0.952 & 0.950 & 0.964 & 0.971 & 0.965 & 0.970 \\
1491     GSC & & 0.980 & 0.959 & 0.990 & 0.983 & 0.992 & 0.989 \\
1492     RF & & 0.984 & 0.975 & 0.996 & 0.995 & 0.998 & 0.998 \\
1493     \bottomrule
1494     \end{tabular}
1495     \label{tab:solnWeak}
1496     \end{table}
1497    
1498     \begin{table}[htbp]
1499     \centering
1500     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1501     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE WEAK SODIUM
1502     CHLORIDE SOLUTION SYSTEM}
1503    
1504     \footnotesize
1505     \begin{tabular}{@{} ccrrrrrr @{}}
1506     \toprule
1507     \toprule
1508     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1509     \cmidrule(lr){3-5}
1510     \cmidrule(l){6-8}
1511     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1512     \midrule
1513     PC & & 882.863 & 510.435 & 344.201 & 277.691 & 154.231 & 100.131 \\
1514     SP & 0.0 & 732.569 & 405.704 & 257.756 & 261.445 & 142.245 & 91.497 \\
1515     & 0.1 & 329.031 & 70.746 & 12.014 & 118.496 & 25.218 & 4.711 \\
1516     & 0.2 & 6.772 & 0.153 & 0.118 & 9.780 & 2.101 & 2.102 \\
1517     & 0.3 & 0.951 & 0.774 & 0.784 & 12.108 & 7.673 & 7.851 \\
1518     SF & 0.0 & 2.555 & 0.762 & 0.313 & 6.590 & 1.328 & 0.558 \\
1519     & 0.1 & 2.561 & 0.560 & 0.123 & 6.464 & 1.162 & 0.457 \\
1520     & 0.2 & 0.501 & 0.118 & 0.118 & 5.698 & 2.074 & 2.099 \\
1521     & 0.3 & 0.943 & 0.774 & 0.784 & 12.118 & 7.674 & 7.851 \\
1522     GSC & & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1523     RF & & 2.415 & 0.452 & 0.130 & 6.915 & 1.423 & 0.507 \\
1524     \midrule
1525     GSSP & 0.0 & 2.915 & 0.643 & 0.261 & 9.576 & 3.133 & 1.812 \\
1526     & 0.1 & 2.251 & 0.324 & 0.064 & 7.628 & 1.639 & 0.497 \\
1527     & 0.2 & 0.590 & 0.118 & 0.116 & 6.080 & 2.096 & 2.103 \\
1528     & 0.3 & 0.953 & 0.759 & 0.780 & 12.347 & 7.683 & 7.849 \\
1529     GSSF & 0.0 & 1.541 & 0.301 & 0.096 & 6.407 & 1.316 & 0.496 \\
1530     & 0.1 & 1.541 & 0.237 & 0.050 & 6.356 & 1.202 & 0.457 \\
1531     & 0.2 & 0.568 & 0.118 & 0.116 & 6.166 & 2.105 & 2.105 \\
1532     & 0.3 & 0.954 & 0.759 & 0.780 & 12.337 & 7.684 & 7.849 \\
1533     \bottomrule
1534     \end{tabular}
1535     \label{tab:solnWeakAng}
1536     \end{table}
1537    
1538     Because this system is a perturbation of the pure liquid water system,
1539     comparisons are best drawn between these two sets. The {\sc sp} and
1540     {\sc sf} methods are not significantly affected by the inclusion of a
1541     few ions. The aspect of cutoff sphere neutralization aids in the
1542     smooth incorporation of these ions; thus, all of the observations
1543     regarding these methods carry over from section
1544     \ref{sec:WaterResults}. The differences between these systems are more
1545     visible for the {\sc rf} method. Though good force agreement is still
1546     maintained, the energy gaps show a significant increase in the scatter
1547     of the data.
1548    
1549     \subsection{1.1M NaCl Solution Results}
1550    
1551     The bridging of the charged atomic and neutral molecular systems was
1552     further developed by considering a high ionic strength system
1553     consisting of 40 ions in the 1000 SPC/E water solvent ($\approx$1.1
1554     M). The results for the energy gap comparisons and the force and
1555     torque vector magnitude comparisons are shown in table
1556     \ref{tab:solnStr}. The force and torque vector directionality
1557     results are displayed separately in table \ref{tab:solnStrAng}, where
1558     the effect of group-based cutoffs and switching functions on the {\sc
1559     sp} and {\sc sf} potentials are investigated.
1560    
1561     \begin{table}[htbp]
1562     \centering
1563     \caption{REGRESSION RESULTS OF THE STRONG SODIUM CHLORIDE SOLUTION
1564     SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR MAGNITUDES
1565     ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1566    
1567     \footnotesize
1568     \begin{tabular}{@{} ccrrrrrr @{}}
1569     \toprule
1570     \toprule
1571     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1572     \cmidrule(lr){3-4}
1573     \cmidrule(lr){5-6}
1574     \cmidrule(l){7-8}
1575     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1576     \midrule
1577     PC & & -0.081 & 0.000 & 0.945 & 0.001 & 0.073 & 0.000 \\
1578     SP & 0.0 & 0.978 & 0.469 & 0.996 & 0.672 & 0.975 & 0.668 \\
1579     & 0.1 & 0.944 & 0.645 & 0.997 & 0.886 & 0.991 & 0.978 \\
1580     & 0.2 & 0.873 & 0.896 & 0.985 & 0.993 & 0.980 & 0.993 \\
1581     & 0.3 & 0.831 & 0.860 & 0.960 & 0.979 & 0.955 & 0.977 \\
1582     SF & 0.0 & 0.858 & 0.905 & 0.985 & 0.970 & 0.990 & 0.998 \\
1583     & 0.1 & 0.865 & 0.907 & 0.992 & 0.974 & 0.994 & 0.999 \\
1584     & 0.2 & 0.862 & 0.894 & 0.985 & 0.993 & 0.980 & 0.993 \\
1585     & 0.3 & 0.831 & 0.859 & 0.960 & 0.979 & 0.955 & 0.977 \\
1586     GSC & & 1.985 & 0.152 & 0.760 & 0.031 & 1.106 & 0.062 \\
1587     RF & & 2.414 & 0.116 & 0.813 & 0.017 & 1.434 & 0.047 \\
1588     \midrule
1589     PC & & -7.028 & 0.000 & -9.364 & 0.000 & 0.925 & 0.865 \\
1590     SP & 0.0 & 0.701 & 0.319 & 0.909 & 0.773 & 0.861 & 0.665 \\
1591     & 0.1 & 0.824 & 0.565 & 0.970 & 0.930 & 0.990 & 0.979 \\
1592     & 0.2 & 0.988 & 0.981 & 0.995 & 0.998 & 0.991 & 0.998 \\
1593     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1594     SF & 0.0 & 0.993 & 0.988 & 0.992 & 0.984 & 0.998 & 0.999 \\
1595     & 0.1 & 0.993 & 0.989 & 0.993 & 0.986 & 0.998 & 1.000 \\
1596     & 0.2 & 0.993 & 0.992 & 0.995 & 0.998 & 0.991 & 0.998 \\
1597     & 0.3 & 0.983 & 0.985 & 0.985 & 0.991 & 0.978 & 0.990 \\
1598     GSC & & 0.964 & 0.897 & 0.970 & 0.917 & 0.925 & 0.865 \\
1599     RF & & 0.994 & 0.864 & 0.988 & 0.865 & 0.980 & 0.784 \\
1600     \midrule
1601     PC & & -2.212 & 0.000 & -0.588 & 0.000 & 0.953 & 0.925 \\
1602     SP & 0.0 & 0.800 & 0.479 & 0.930 & 0.804 & 0.924 & 0.759 \\
1603     & 0.1 & 0.883 & 0.694 & 0.976 & 0.942 & 0.993 & 0.986 \\
1604     & 0.2 & 0.952 & 0.943 & 0.980 & 0.984 & 0.980 & 0.983 \\
1605     & 0.3 & 0.914 & 0.909 & 0.943 & 0.948 & 0.944 & 0.946 \\
1606     SF & 0.0 & 0.945 & 0.953 & 0.980 & 0.984 & 0.991 & 0.998 \\
1607     & 0.1 & 0.951 & 0.954 & 0.987 & 0.986 & 0.995 & 0.998 \\
1608     & 0.2 & 0.951 & 0.946 & 0.980 & 0.984 & 0.980 & 0.983 \\
1609     & 0.3 & 0.914 & 0.908 & 0.943 & 0.948 & 0.944 & 0.946 \\
1610     GSC & & 0.882 & 0.818 & 0.939 & 0.902 & 0.953 & 0.925 \\
1611     RF & & 0.949 & 0.939 & 0.988 & 0.988 & 0.992 & 0.993 \\
1612     \bottomrule
1613     \end{tabular}
1614     \label{tab:solnStr}
1615     \end{table}
1616    
1617     \begin{table}[htbp]
1618     \centering
1619     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR DISTRIBUTIONS
1620     OF THE FORCE AND TORQUE VECTORS IN THE STRONG SODIUM CHLORIDE SOLUTION
1621     SYSTEM}
1622    
1623     \footnotesize
1624     \begin{tabular}{@{} ccrrrrrr @{}}
1625     \toprule
1626     \toprule
1627     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1628     \cmidrule(lr){3-5}
1629     \cmidrule(l){6-8}
1630     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1631     \midrule
1632     PC & & 957.784 & 513.373 & 2.260 & 340.043 & 179.443 & 13.079 \\
1633     SP & 0.0 & 786.244 & 139.985 & 259.289 & 311.519 & 90.280 & 105.187 \\
1634     & 0.1 & 354.697 & 38.614 & 12.274 & 144.531 & 23.787 & 5.401 \\
1635     & 0.2 & 7.674 & 0.363 & 0.215 & 16.655 & 3.601 & 3.634 \\
1636     & 0.3 & 1.745 & 1.456 & 1.449 & 23.669 & 14.376 & 14.240 \\
1637     SF & 0.0 & 3.282 & 8.567 & 0.369 & 11.904 & 6.589 & 0.717 \\
1638     & 0.1 & 3.263 & 7.479 & 0.142 & 11.634 & 5.750 & 0.591 \\
1639     & 0.2 & 0.686 & 0.324 & 0.215 & 10.809 & 3.580 & 3.635 \\
1640     & 0.3 & 1.749 & 1.456 & 1.449 & 23.635 & 14.375 & 14.240 \\
1641     GSC & & 6.181 & 2.904 & 2.263 & 44.349 & 19.442 & 12.873 \\
1642     RF & & 3.891 & 0.847 & 0.323 & 18.628 & 3.995 & 2.072 \\
1643     \midrule
1644     GSSP & 0.0 & 6.197 & 2.929 & 2.290 & 44.441 & 19.442 & 12.873 \\
1645     & 0.1 & 4.688 & 1.064 & 0.260 & 31.208 & 6.967 & 2.303 \\
1646     & 0.2 & 1.021 & 0.218 & 0.213 & 14.425 & 3.629 & 3.649 \\
1647     & 0.3 & 1.752 & 1.454 & 1.451 & 23.540 & 14.390 & 14.245 \\
1648     GSSF & 0.0 & 2.494 & 0.546 & 0.217 & 16.391 & 3.230 & 1.613 \\
1649     & 0.1 & 2.448 & 0.429 & 0.106 & 16.390 & 2.827 & 1.159 \\
1650     & 0.2 & 0.899 & 0.214 & 0.213 & 13.542 & 3.583 & 3.645 \\
1651     & 0.3 & 1.752 & 1.454 & 1.451 & 23.587 & 14.390 & 14.245 \\
1652     \bottomrule
1653     \end{tabular}
1654     \label{tab:solnStrAng}
1655     \end{table}
1656    
1657     The {\sc rf} method struggles with the jump in ionic strength. The
1658     configuration energy differences degrade to unusable levels while the
1659     forces and torques show a more modest reduction in the agreement with
1660     {\sc spme}. The {\sc rf} method was designed for homogeneous systems,
1661     and this attribute is apparent in these results.
1662    
1663     The {\sc sp} and {\sc sf} methods require larger cutoffs to maintain
1664     their agreement with {\sc spme}. With these results, we still
1665     recommend undamped to moderate damping for the {\sc sf} method and
1666     moderate damping for the {\sc sp} method, both with cutoffs greater
1667     than 12\AA.
1668    
1669     \subsection{6\AA\ Argon Sphere in SPC/E Water Results}
1670    
1671     The final model system studied was a 6\AA\ sphere of Argon solvated
1672     by SPC/E water. This serves as a test case of a specifically sized
1673     electrostatic defect in a disordered molecular system. The results for
1674     the energy gap comparisons and the force and torque vector magnitude
1675     comparisons are shown in table \ref{tab:argon}. The force and torque
1676     vector directionality results are displayed separately in table
1677     \ref{tab:argonAng}, where the effect of group-based cutoffs and
1678     switching functions on the {\sc sp} and {\sc sf} potentials are
1679     investigated.
1680    
1681     \begin{table}[htbp]
1682     \centering
1683     \caption{REGRESSION RESULTS OF THE 6\AA\ ARGON SPHERE IN LIQUID
1684     WATER SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECTOR
1685     MAGNITUDES ({\it middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1686    
1687     \footnotesize
1688     \begin{tabular}{@{} ccrrrrrr @{}}
1689     \toprule
1690     \toprule
1691     & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
1692     \cmidrule(lr){3-4}
1693     \cmidrule(lr){5-6}
1694     \cmidrule(l){7-8}
1695     Method & $\alpha$ & slope & $R^2$ & slope & $R^2$ & slope & $R^2$ \\
1696     \midrule
1697     PC & & 2.320 & 0.008 & -0.650 & 0.001 & 3.848 & 0.029 \\
1698     SP & 0.0 & 1.053 & 0.711 & 0.977 & 0.820 & 0.974 & 0.882 \\
1699     & 0.1 & 1.032 & 0.846 & 0.989 & 0.965 & 0.992 & 0.994 \\
1700     & 0.2 & 0.993 & 0.995 & 0.982 & 0.998 & 0.986 & 0.998 \\
1701     & 0.3 & 0.968 & 0.995 & 0.954 & 0.992 & 0.961 & 0.994 \\
1702     SF & 0.0 & 0.982 & 0.996 & 0.992 & 0.999 & 0.993 & 1.000 \\
1703     & 0.1 & 0.987 & 0.996 & 0.996 & 0.999 & 0.997 & 1.000 \\
1704     & 0.2 & 0.989 & 0.998 & 0.984 & 0.998 & 0.989 & 0.998 \\
1705     & 0.3 & 0.971 & 0.995 & 0.957 & 0.992 & 0.965 & 0.994 \\
1706     GSC & & 1.002 & 0.983 & 0.992 & 0.973 & 0.996 & 0.971 \\
1707     RF & & 0.998 & 0.995 & 0.999 & 0.998 & 0.998 & 0.998 \\
1708     \midrule
1709     PC & & -36.559 & 0.002 & -44.917 & 0.004 & -52.945 & 0.006 \\
1710     SP & 0.0 & 0.890 & 0.786 & 0.927 & 0.867 & 0.949 & 0.909 \\
1711     & 0.1 & 0.942 & 0.895 & 0.984 & 0.974 & 0.997 & 0.995 \\
1712     & 0.2 & 0.999 & 0.997 & 1.000 & 1.000 & 1.000 & 1.000 \\
1713     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1714     SF & 0.0 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1715     & 0.1 & 1.000 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1716     & 0.2 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 & 1.000 \\
1717     & 0.3 & 1.001 & 0.999 & 1.001 & 1.000 & 1.001 & 1.000 \\
1718     GSC & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1719     RF & & 0.999 & 0.999 & 1.000 & 1.000 & 1.000 & 1.000 \\
1720     \midrule
1721     PC & & 1.984 & 0.000 & 0.012 & 0.000 & 1.357 & 0.000 \\
1722     SP & 0.0 & 0.850 & 0.552 & 0.907 & 0.703 & 0.938 & 0.793 \\
1723     & 0.1 & 0.924 & 0.755 & 0.980 & 0.936 & 0.995 & 0.988 \\
1724     & 0.2 & 0.985 & 0.983 & 0.986 & 0.988 & 0.987 & 0.988 \\
1725     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1726     SF & 0.0 & 0.977 & 0.989 & 0.987 & 0.995 & 0.992 & 0.998 \\
1727     & 0.1 & 0.982 & 0.989 & 0.992 & 0.996 & 0.997 & 0.998 \\
1728     & 0.2 & 0.984 & 0.987 & 0.986 & 0.987 & 0.987 & 0.988 \\
1729     & 0.3 & 0.961 & 0.966 & 0.959 & 0.964 & 0.960 & 0.966 \\
1730     GSC & & 0.995 & 0.981 & 0.999 & 0.990 & 1.000 & 0.993 \\
1731     RF & & 0.993 & 0.988 & 0.997 & 0.995 & 0.999 & 0.998 \\
1732     \bottomrule
1733     \end{tabular}
1734     \label{tab:argon}
1735     \end{table}
1736    
1737     \begin{table}[htbp]
1738     \centering
1739     \caption{VARIANCE RESULTS FROM GAUSSIAN FITS TO ANGULAR
1740     DISTRIBUTIONS OF THE FORCE AND TORQUE VECTORS IN THE 6\AA\ SPHERE OF
1741     ARGON IN LIQUID WATER SYSTEM}
1742    
1743     \footnotesize
1744     \begin{tabular}{@{} ccrrrrrr @{}}
1745     \toprule
1746     \toprule
1747     & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
1748     \cmidrule(lr){3-5}
1749     \cmidrule(l){6-8}
1750     Method & $\alpha$ & 9\AA & 12\AA & 15\AA & 9\AA & 12\AA & 15\AA \\
1751     \midrule
1752     PC & & 568.025 & 265.993 & 195.099 & 246.626 & 138.600 & 91.654 \\
1753     SP & 0.0 & 504.578 & 251.694 & 179.932 & 231.568 & 131.444 & 85.119 \\
1754     & 0.1 & 224.886 & 49.746 & 9.346 & 104.482 & 23.683 & 4.480 \\
1755     & 0.2 & 4.889 & 0.197 & 0.155 & 6.029 & 2.507 & 2.269 \\
1756     & 0.3 & 0.817 & 0.833 & 0.812 & 8.286 & 8.436 & 8.135 \\
1757     SF & 0.0 & 1.924 & 0.675 & 0.304 & 3.658 & 1.448 & 0.600 \\
1758     & 0.1 & 1.937 & 0.515 & 0.143 & 3.565 & 1.308 & 0.546 \\
1759     & 0.2 & 0.407 & 0.166 & 0.156 & 3.086 & 2.501 & 2.274 \\
1760     & 0.3 & 0.815 & 0.833 & 0.812 & 8.330 & 8.437 & 8.135 \\
1761     GSC & & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1762     RF & & 1.822 & 0.408 & 0.142 & 3.799 & 1.362 & 0.550 \\
1763     \midrule
1764     GSSP & 0.0 & 2.098 & 0.584 & 0.284 & 5.391 & 2.414 & 1.501 \\
1765     & 0.1 & 1.652 & 0.309 & 0.087 & 4.197 & 1.401 & 0.590 \\
1766     & 0.2 & 0.465 & 0.165 & 0.153 & 3.323 & 2.529 & 2.273 \\
1767     & 0.3 & 0.813 & 0.825 & 0.816 & 8.316 & 8.447 & 8.132 \\
1768     GSSF & 0.0 & 1.173 & 0.292 & 0.113 & 3.452 & 1.347 & 0.583 \\
1769     & 0.1 & 1.166 & 0.240 & 0.076 & 3.381 & 1.281 & 0.575 \\
1770     & 0.2 & 0.459 & 0.165 & 0.153 & 3.430 & 2.542 & 2.273 \\
1771     & 0.3 & 0.814 & 0.825 & 0.816 & 8.325 & 8.447 & 8.132 \\
1772     \bottomrule
1773     \end{tabular}
1774     \label{tab:argonAng}
1775     \end{table}
1776    
1777     This system does not appear to show any significant deviations from
1778     the previously observed results. The {\sc sp} and {\sc sf} methods
1779     have agreements similar to those observed in section
1780     \ref{sec:WaterResults}. The only significant difference is the
1781     improvement in the configuration energy differences for the {\sc rf}
1782     method. This is surprising in that we are introducing an inhomogeneity
1783     to the system; however, this inhomogeneity is charge-neutral and does
1784     not result in charged cutoff spheres. The charge-neutrality of the
1785     cutoff spheres, which the {\sc sp} and {\sc sf} methods explicitly
1786     enforce, seems to play a greater role in the stability of the {\sc rf}
1787     method than the required homogeneity of the environment.
1788    
1789    
1790     \section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics}
1791    
1792     Zahn {\it et al.} investigated the structure and dynamics of water
1793     using equations (\ref{eq:ZahnPot}) and
1794     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
1795     that a method similar (but not identical with) the damped {\sc sf}
1796     method resulted in properties very similar to those obtained when
1797     using the Ewald summation. The properties they studied (pair
1798     distribution functions, diffusion constants, and velocity and
1799     orientational correlation functions) may not be particularly sensitive
1800     to the long-range and collective behavior that governs the
1801     low-frequency behavior in crystalline systems. Additionally, the
1802     ionic crystals are the worst case scenario for the pairwise methods
1803     because they lack the reciprocal space contribution contained in the
1804     Ewald summation.
1805    
1806     We are using two separate measures to probe the effects of these
1807     alternative electrostatic methods on the dynamics in crystalline
1808     materials. For short- and intermediate-time dynamics, we are
1809     computing the velocity autocorrelation function, and for long-time
1810     and large length-scale collective motions, we are looking at the
1811     low-frequency portion of the power spectrum.
1812    
1813     \begin{figure}
1814     \centering
1815     \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
1816     \caption{Velocity autocorrelation functions of NaCl crystals at
1817     1000K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \&
1818     0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2\AA$^{-1}$). The inset is
1819     a magnification of the area around the first minimum. The times to
1820     first collision are nearly identical, but differences can be seen in
1821     the peaks and troughs, where the undamped and weakly damped methods
1822     are stiffer than the moderately damped and {\sc spme} methods.}
1823     \label{fig:vCorrPlot}
1824     \end{figure}
1825    
1826     The short-time decay of the velocity autocorrelation function through
1827     the first collision are nearly identical in figure
1828     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1829     how the methods differ. The undamped {\sc sf} method has deeper
1830     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1831     any of the other methods. As the damping parameter ($\alpha$) is
1832     increased, these peaks are smoothed out, and the {\sc sf} method
1833     approaches the {\sc spme} results. With $\alpha$ values of 0.2\AA$^{-1}$,
1834     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1835     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1836     and {\sc sp} potentials become nearly identical with increased
1837     damping. However, this appears to indicate that once damping is
1838     utilized, the details of the form of the potential (and forces)
1839     constructed out of the damped electrostatic interaction are less
1840     important.
1841    
1842     \section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics}
1843    
1844     To evaluate how the differences between the methods affect the
1845     collective long-time motion, we computed power spectra from long-time
1846     traces of the velocity autocorrelation function. The power spectra for
1847     the best-performing alternative methods are shown in
1848     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1849     a cubic switching function between 40 and 50ps was used to reduce the
1850     ringing resulting from data truncation. This procedure had no
1851     noticeable effect on peak location or magnitude.
1852    
1853     \begin{figure}
1854     \centering
1855     \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
1856     \caption{Power spectra obtained from the velocity auto-correlation
1857     functions of NaCl crystals at 1000K while using {\sc spme}, {\sc sf}
1858     ($\alpha$ = 0, 0.1, \& 0.2\AA$^{-1}$), and {\sc sp} ($\alpha$ =
1859     0.2\AA$^{-1}$). The inset shows the frequency region below 100
1860     cm$^{-1}$ to highlight where the spectra differ.}
1861     \label{fig:methodPS}
1862     \end{figure}
1863    
1864     While the high frequency regions of the power spectra for the
1865     alternative methods are quantitatively identical with Ewald spectrum,
1866     the low frequency region shows how the summation methods differ.
1867     Considering the low-frequency inset (expanded in the upper frame of
1868     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1869     correlated motions are blue-shifted when using undamped or weakly
1870     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1871     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1872     correlated motion to the Ewald method (which has a convergence
1873     parameter of 0.3119\AA$^{-1}$). This weakening of the electrostatic
1874     interaction with increased damping explains why the long-ranged
1875     correlated motions are at lower frequencies for the moderately damped
1876     methods than for undamped or weakly damped methods.
1877    
1878     To isolate the role of the damping constant, we have computed the
1879     spectra for a single method ({\sc sf}) with a range of damping
1880     constants and compared this with the {\sc spme} spectrum.
1881     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1882     electrostatic damping red-shifts the lowest frequency phonon modes.
1883     However, even without any electrostatic damping, the {\sc sf} method
1884     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1885     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1886     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1887     Most} of the collective behavior in the crystal is accurately captured
1888     using the {\sc sf} method. Quantitative agreement with Ewald can be
1889     obtained using moderate damping in addition to the shifting at the
1890     cutoff distance.
1891    
1892     \begin{figure}
1893     \centering
1894     \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1895     \caption{Effect of damping on the two lowest-frequency phonon modes in
1896     the NaCl crystal at 1000K. The undamped shifted force ({\sc sf})
1897     method is off by less than 10 cm$^{-1}$, and increasing the
1898     electrostatic damping to 0.25\AA$^{-1}$ gives quantitative agreement
1899     with the power spectrum obtained using the Ewald sum. Over-damping can
1900     result in underestimates of frequencies of the long-wavelength
1901     motions.}
1902     \label{fig:dampInc}
1903     \end{figure}
1904    
1905     \section{An Application: TIP5P-E Water}\label{sec:t5peApplied}
1906    
1907     The above sections focused on the energetics and dynamics of a variety
1908     of systems when utilizing the {\sc sp} and {\sc sf} pairwise
1909     techniques. A unitary correlation with results obtained using the
1910     Ewald summation should result in a successful reproduction of both the
1911     static and dynamic properties of any selected system. To test this,
1912     we decided to calculate a series of properties for the TIP5P-E water
1913     model when using the {\sc sf} technique.
1914    
1915     The TIP5P-E water model is a variant of Mahoney and Jorgensen's
1916     five-point transferable intermolecular potential (TIP5P) model for
1917     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
1918     maximum anomaly present in liquid water near 4$^\circ$C. As with many
1919     previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
1920     and SPC/E), TIP5P was parametrized using a simple cutoff with no
1921     long-range electrostatic
1922     correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
1923     Without this correction, the pressure term on the central particle
1924     from the surroundings is missing. Because they expand to compensate
1925     for this added pressure term when this correction is included, systems
1926     composed of these particles tend to under-predict the density of water
1927     under standard conditions. When using any form of long-range
1928     electrostatic correction, it has become common practice to develop or
1929     utilize a reparametrized water model that corrects for this
1930     effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
1931     this practice and was optimized specifically for use with the Ewald
1932     summation.\cite{Rick04} In his publication, Rick preserved the
1933     geometry and point charge magnitudes in TIP5P and focused on altering
1934     the Lennard-Jones parameters to correct the density at
1935     298K.\cite{Rick04} With the density corrected, he compared common
1936     water properties for TIP5P-E using the Ewald sum with TIP5P using a
1937     9\AA\ cutoff.
1938    
1939     In the following sections, we compared these same water properties
1940     calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
1941     {\sc sf} technique. In the above evaluation of the pairwise
1942     techniques, we observed some flexibility in the choice of parameters.
1943     Because of this, the following comparisons include the {\sc sf}
1944     technique with a 12\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
1945     0.2\AA$^{-1}$, as well as a 9\AA\ cutoff with an $\alpha$ =
1946     0.2\AA$^{-1}$.
1947    
1948     \subsection{Density}\label{sec:t5peDensity}
1949    
1950     As stated previously, the property that prompted the development of
1951     TIP5P-E was the density at 1 atm. The density depends upon the
1952     internal pressure of the system in the $NPT$ ensemble, and the
1953     calculation of the pressure includes a components from both the
1954     kinetic energy and the virial. More specifically, the instantaneous
1955     molecular pressure ($p(t)$) is given by
1956     \begin{equation}
1957     p(t) = \frac{1}{\textrm{d}V}\sum_\mu
1958     \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
1959     + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
1960     \label{eq:MolecularPressure}
1961     \end{equation}
1962     where $V$ is the volume, $\mathbf{P}_{\mu}$ is the momentum of
1963     molecule $\mu$, $\mathbf{R}_\mu$ is the position of the center of mass
1964     ($M_\mu$) of molecule $\mu$, and $\mathbf{F}_{\mu i}$ is the force on
1965     atom $i$ of molecule $\mu$.\cite{Melchionna93} The virial term (the
1966     right term in the brackets of equation \ref{eq:MolecularPressure}) is
1967     directly dependent on the interatomic forces. Since the {\sc sp}
1968     method does not modify the forces (see
1969     section. \ref{sec:PairwiseDerivation}), the pressure using {\sc sp} will
1970     be identical to that obtained without an electrostatic correction.
1971     The {\sc sf} method does alter the virial component and, by way of the
1972     modified pressures, should provide densities more in line with those
1973     obtained using the Ewald summation.
1974    
1975     To compare densities, $NPT$ simulations were performed with the same
1976     temperatures as those selected by Rick in his Ewald summation
1977     simulations.\cite{Rick04} In order to improve statistics around the
1978     density maximum, 3ns trajectories were accumulated at 0, 12.5, and
1979     25$^\circ$C, while 2ns trajectories were obtained at all other
1980     temperatures. The average densities were calculated from the later
1981     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
1982     method for accumulating statistics, these sequences were spliced into
1983     200 segments to calculate the average density and standard deviation
1984     at each temperature.\cite{Mahoney00}
1985    
1986     \begin{figure}
1987     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
1988     \caption{Density versus temperature for the TIP5P-E water model when
1989     using the Ewald summation (Ref. \cite{Rick04}) and the {\sc sf} method
1990     with various parameters. The pressure term from the image-charge shell
1991     is larger than that provided by the reciprocal-space portion of the
1992     Ewald summation, leading to slightly lower densities. This effect is
1993     more visible with the 9\AA\ cutoff, where the image charges exert a
1994     greater force on the central particle. The error bars for the {\sc sf}
1995     methods show plus or minus the standard deviation of the density
1996     measurement at each temperature.}
1997     \label{fig:t5peDensities}
1998     \end{figure}
1999    
2000     Figure \ref{fig:t5peDensities} shows the densities calculated for
2001     TIP5P-E using differing electrostatic corrections overlaid on the
2002     experimental values.\cite{CRC80} The densities when using the {\sc sf}
2003     technique are close to, though typically lower than, those calculated
2004     while using the Ewald summation. These slightly reduced densities
2005     indicate that the pressure component from the image charges at
2006     R$_\textrm{c}$ is larger than that exerted by the reciprocal-space
2007     portion of the Ewald summation. Bringing the image charges closer to
2008     the central particle by choosing a 9\AA\ R$_\textrm{c}$ (rather than
2009     the preferred 12\AA\ R$_\textrm{c}$) increases the strength of their
2010     interactions, resulting in a further reduction of the densities.
2011    
2012     Because the strength of the image charge interactions has a noticeable
2013     effect on the density, we would expect the use of electrostatic
2014     damping to also play a role in these calculations. Larger values of
2015     $\alpha$ weaken the pair-interactions; and since electrostatic damping
2016     is distance-dependent, force components from the image charges will be
2017     reduced more than those from particles close the the central
2018     charge. This effect is visible in figure \ref{fig:t5peDensities} with
2019     the damped {\sc sf} sums showing slightly higher densities; however,
2020     it is apparent that the choice of cutoff radius plays a much more
2021     important role in the resulting densities.
2022    
2023     As a final note, all of the above density calculations were performed
2024     with systems of 512 water molecules. Rick observed a system sized
2025     dependence of the computed densities when using the Ewald summation,
2026     most likely due to his tying of the convergence parameter to the box
2027     dimensions.\cite{Rick04} For systems of 256 water molecules, the
2028     calculated densities were on average 0.002 to 0.003 g/cm$^3$ lower. A
2029     system size of 256 molecules would force the use of a shorter
2030     R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
2031     lower the densities. Moving to larger systems, as long as the
2032     R$_\textrm{c}$ remains at a fixed value, we would expect the densities
2033     to remain constant.
2034    
2035     \subsection{Liquid Structure}\label{sec:t5peLiqStructure}
2036    
2037     A common function considered when developing and comparing water
2038     models is the oxygen-oxygen radial distribution function
2039     ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
2040     finding a pair of oxygen atoms some distance ($r$) apart relative to a
2041     random distribution at the same density.\cite{Allen87} It is
2042     calculated via
2043     \begin{equation}
2044     g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
2045     \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
2046     \label{eq:GOOofR}
2047     \end{equation}
2048     where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
2049     atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
2050     neutron scattering experiments through the oxygen-oxygen structure
2051     factor ($S_\textrm{OO}(k)$) by the following relationship:
2052     \begin{equation}
2053     S_\textrm{OO}(k) = 1 + 4\pi\rho
2054     \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
2055     \label{eq:SOOofK}
2056     \end{equation}
2057     Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
2058     of $g_\textrm{OO}(r)$.
2059    
2060     The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
2061     been compared in great detail with the various common water models,
2062     and TIP5P was found to be in better agreement than other rigid,
2063     non-polarizable models.\cite{Sorenson00} This excellent agreement with
2064     experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
2065     check whether the choice of using the Ewald summation or the {\sc sf}
2066     technique alters the liquid structure, the $g_\textrm{OO}(r)$s at 298K
2067     and 1atm were determined for the systems compared in the previous
2068     section.
2069    
2070     \begin{figure}
2071     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
2072     \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298K and
2073     1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
2074     sf} technique with varying parameters. Even with the reduced densities
2075     using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
2076     identical.}
2077     \label{fig:t5peGofRs}
2078     \end{figure}
2079    
2080     The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
2081     sf} technique with a various parameters are overlaid on the
2082     $g_\textrm{OO}(r)$ while using the Ewald summation. The differences in
2083     density do not appear to have any effect on the liquid structure as
2084     the $g_\textrm{OO}(r)$s are indistinguishable. These results indicate
2085     that the $g_\textrm{OO}(r)$ is insensitive to the choice of
2086     electrostatic correction.
2087    
2088     \subsection{Thermodynamic Properties}\label{sec:t5peThermo}
2089    
2090     In addition to the density, there are a variety of thermodynamic
2091     quantities that can be calculated for water and compared directly to
2092     experimental values. Some of these additional quantities include the
2093     latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
2094     pressure heat capacity ($C_p$), the isothermal compressibility
2095     ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
2096     dielectric constant ($\epsilon$). All of these properties were
2097     calculated for TIP5P-E with the Ewald summation, so they provide a
2098     good set for comparisons involving the {\sc sf} technique.
2099    
2100     The $\Delta H_\textrm{vap}$ is the enthalpy change required to
2101     transform one mol of substance from the liquid phase to the gas
2102     phase.\cite{Berry00} In molecular simulations, this quantity can be
2103     determined via
2104     \begin{equation}
2105     \begin{split}
2106     \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
2107     &= E_\textrm{gas} - E_\textrm{liq.}
2108     + p(V_\textrm{gas} - V_\textrm{liq.}) \\
2109     &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
2110     \end{split}
2111     \label{eq:DeltaHVap}
2112     \end{equation}
2113     where $E$ is the total energy, $U$ is the potential energy, $p$ is the
2114     pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
2115     the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
2116     seen in the last line of equation (\ref{eq:DeltaHVap}), we can
2117     approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
2118     state. This allows us to cancel the kinetic energy terms, leaving only
2119     the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
2120     several orders of magnitude larger than that of the liquid, so we can
2121     neglect the liquid $pV$ term.
2122    
2123     The remaining thermodynamic properties can all be calculated from
2124     fluctuations of the enthalpy, volume, and system dipole
2125     moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
2126     enthalpy in constant pressure simulations via
2127     \begin{equation}
2128     \begin{split}
2129     C_p = \left(\frac{\partial H}{\partial T}\right)_{N,p}
2130     = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
2131     \end{split}
2132     \label{eq:Cp}
2133     \end{equation}
2134     where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
2135     \begin{equation}
2136     \begin{split}
2137     \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
2138     = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
2139     {k_BT\langle V\rangle_{N,P,T}},
2140     \end{split}
2141     \label{eq:kappa}
2142     \end{equation}
2143     and $\alpha_p$ can be calculated via
2144     \begin{equation}
2145     \begin{split}
2146     \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
2147     = \frac{(\langle VH\rangle_{N,P,T}
2148     - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
2149     {k_BT^2\langle V\rangle_{N,P,T}}.
2150     \end{split}
2151     \label{eq:alpha}
2152     \end{equation}
2153     Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
2154     be calculated for systems of non-polarizable substances via
2155     \begin{equation}
2156     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
2157     \label{eq:staticDielectric}
2158     \end{equation}
2159     where $\epsilon_0$ is the permittivity of free space and $\langle
2160     M^2\rangle$ is the fluctuation of the system dipole
2161     moment.\cite{Allen87} The numerator in the fractional term in equation
2162     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
2163     dipole moment, identical to the quantity calculated in the
2164     finite-system Kirkwood $g$ factor ($G_k$):
2165     \begin{equation}
2166     G_k = \frac{\langle M^2\rangle}{N\mu^2},
2167     \label{eq:KirkwoodFactor}
2168     \end{equation}
2169     where $\mu$ is the dipole moment of a single molecule of the
2170     homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
2171     fluctuation term in both equation (\ref{eq:staticDielectric}) and
2172     \ref{eq:KirkwoodFactor} is calculated as follows,
2173     \begin{equation}
2174     \begin{split}
2175     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
2176     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
2177     &= \langle M_x^2+M_y^2+M_z^2\rangle
2178     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
2179     + \langle M_x\rangle^2).
2180     \end{split}
2181     \label{eq:fluctBoxDipole}
2182     \end{equation}
2183     This fluctuation term can be accumulated during the simulation;
2184     however, it converges rather slowly, thus requiring multi-nanosecond
2185     simulation times.\cite{Horn04} In the case of tin-foil boundary
2186     conditions, the dielectric/surface term of equation (\ref{eq:EwaldSum})
2187     is equal to zero. Since the {\sc sf} method also lacks this
2188     dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
2189     valid for determining static dielectric constants.
2190    
2191     All of the above properties were calculated from the same trajectories
2192     used to determine the densities in section \ref{sec:t5peDensity}
2193     except for the static dielectric constants. The $\epsilon$ values were
2194     accumulated from 2ns $NVE$ ensemble trajectories with system densities
2195     fixed at the average values from the $NPT$ simulations at each of the
2196     temperatures. The resulting values are displayed in figure
2197     \ref{fig:t5peThermo}.
2198     \begin{figure}
2199     \centering
2200     \includegraphics[width=5.5in]{./figures/t5peThermo.pdf}
2201     \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
2202     and the {\sc sf} techniques along with the experimental values. Units
2203     for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
2204     cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
2205     and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
2206     reference \cite{Rick04}. Experimental values for $\Delta
2207     H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
2208     \cite{Kell75}. Experimental values for $C_p$ are from reference
2209     \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
2210     \cite{Malmberg56}.}
2211     \label{fig:t5peThermo}
2212     \end{figure}
2213    
2214     As observed for the density in section \ref{sec:t5peDensity}, the
2215     property trends with temperature seen when using the Ewald summation
2216     are reproduced with the {\sc sf} technique. Differences include the
2217     calculated values of $\Delta H_\textrm{vap}$ under-predicting the Ewald
2218     values. This is to be expected due to the direct weakening of the
2219     electrostatic interaction through forced neutralization in {\sc
2220     sf}. This results in an increase of the intermolecular potential
2221     producing lower values from equation (\ref{eq:DeltaHVap}). The slopes of
2222     these values with temperature are similar to that seen using the Ewald
2223     summation; however, they are both steeper than the experimental trend,
2224     indirectly resulting in the inflated $C_p$ values at all temperatures.
2225    
2226     Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$
2227     values all overlap within error. As indicated for the $\Delta
2228     H_\textrm{vap}$ and $C_p$ results discussed in the previous paragraph,
2229     the deviations between experiment and simulation in this region are
2230     not the fault of the electrostatic summation methods but are due to
2231     the TIP5P class model itself. Like most rigid, non-polarizable,
2232     point-charge water models, the density decreases with temperature at a
2233     much faster rate than experiment (see figure
2234     \ref{fig:t5peDensities}). The reduced density leads to the inflated
2235     compressibility and expansivity values at higher temperatures seen
2236     here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
2237     and many-body effects are required in order for simulation to overcome
2238     these differences with experiment.\cite{Laasonen93,Donchev06}
2239    
2240     At temperatures below the freezing point for experimental water, the
2241     differences between {\sc sf} and the Ewald summation results are more
2242     apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
2243     indicate a more pronounced transition in the supercooled regime,
2244     particularly in the case of {\sc sf} without damping. This points to
2245     the onset of a more frustrated or glassy behavior for TIP5P-E at
2246     temperatures below 250K in these simulations. Because the systems are
2247     locked in different regions of phase-space, comparisons between
2248     properties at these temperatures are not exactly fair. This
2249     observation is explored in more detail in section
2250     \ref{sec:t5peDynamics}.
2251    
2252     The final thermodynamic property displayed in figure
2253     \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
2254     between the Ewald summation and the {\sc sf} technique (and experiment
2255     for that matter). It is known that the dielectric constant is
2256     dependent upon and quite sensitive to the imposed boundary
2257     conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
2258     converged $\epsilon$ values accumulated for the {\sc sf}
2259     simulations. Lack of a damping function results in dielectric
2260     constants significantly smaller than that obtained using the Ewald
2261     sum. Increasing the damping coefficient to 0.2\AA$^{-1}$ improves the
2262     agreement considerably. It should be noted that the choice of the
2263     ``Ewald coefficient'' value also has a significant effect on the
2264     calculated value when using the Ewald summation. In the simulations of
2265     TIP5P-E with the Ewald sum, this screening parameter was tethered to
2266     the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04} In
2267     general, systems with larger screening parameters reported larger
2268     dielectric constant values, the same behavior we see here with {\sc
2269     sf}; however, the choice of cutoff radius also plays an important
2270     role. In section \ref{sec:dampingDielectric}, this connection is
2271     further explored as optimal damping coefficients for different choices
2272     of $R_\textrm{c}$ are determined for {\sc sf} for capturing the
2273     dielectric behavior.
2274    
2275     \subsection{Dynamic Properties}\label{sec:t5peDynamics}
2276    
2277     To look at the dynamic properties of TIP5P-E when using the {\sc sf}
2278     method, 200ps $NVE$ simulations were performed for each temperature at
2279     the average density reported by the $NPT$ simulations. The
2280     self-diffusion constants ($D$) were calculated with the Einstein
2281     relation using the mean square displacement (MSD),
2282     \begin{equation}
2283     D = \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
2284     \label{eq:MSD}
2285     \end{equation}
2286     where $t$ is time, and $\mathbf{r}_i$ is the position of particle
2287     $i$.\cite{Allen87} Figure \ref{fig:ExampleMSD} shows an example MSD
2288     plot. As labeled in the figure, MSD plots consist of three distinct
2289     regions:
2290    
2291     \begin{enumerate}[itemsep=0pt]
2292     \item parabolic short-time ballistic motion,
2293     \item linear diffusive regime, and
2294     \item poor statistic region at long-time.
2295     \end{enumerate}
2296     The slope from the linear region (region 2) is used to calculate $D$.
2297     \begin{figure}
2298     \centering
2299     \includegraphics[width=3.5in]{./figures/ExampleMSD.pdf}
2300     \caption{Example plot of mean square displacement verses time. The
2301     left red region is the ballistic motion regime, the middle green
2302     region is the linear diffusive regime, and the right blue region is
2303     the region with poor statistics.}
2304     \label{fig:ExampleMSD}
2305     \end{figure}
2306    
2307     \begin{figure}
2308     \centering
2309     \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
2310     \caption{Body-fixed coordinate frame for a water molecule. The
2311     respective molecular principle axes point in the direction of the
2312     labeled frame axes.}
2313     \label{fig:waterFrame}
2314     \end{figure}
2315     In addition to translational diffusion, reorientational time constants
2316     were calculated for comparisons with the Ewald simulations and with
2317     experiments. These values were determined from 25ps $NVE$ trajectories
2318     through calculation of the orientational time correlation function,
2319     \begin{equation}
2320     C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
2321     \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
2322     \label{eq:OrientCorr}
2323     \end{equation}
2324     where $P_l$ is the Legendre polynomial of order $l$ and
2325     $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
2326     principle axis $\alpha$. The principle axis frame for these water
2327     molecules is shown in figure \ref{fig:waterFrame}. As an example,
2328     $C_l^y$ is calculated from the time evolution of the unit vector
2329     connecting the two hydrogen atoms.
2330    
2331     \begin{figure}
2332     \centering
2333     \includegraphics[width=3.5in]{./figures/ExampleOrientCorr.pdf}
2334     \caption{Example plots of the orientational autocorrelation functions
2335     for the first and second Legendre polynomials. These curves show the
2336     time decay of the unit vector along the $y$ principle axis.}
2337     \label{fig:OrientCorr}
2338     \end{figure}
2339     From the orientation autocorrelation functions, we can obtain time
2340     constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
2341     some example plots of orientational autocorrelation functions for the
2342     first and second Legendre polynomials. The relatively short time
2343     portions (between 1 and 3ps for water) of these curves can be fit to
2344     an exponential decay to obtain these constants, and they are directly
2345     comparable to water orientational relaxation times from nuclear
2346     magnetic resonance (NMR). The relaxation constant obtained from
2347     $C_2^y(t)$ is of particular interest because it describes the
2348     relaxation of the principle axis connecting the hydrogen atoms. Thus,
2349     $C_2^y(t)$ can be compared to the intermolecular portion of the
2350     dipole-dipole relaxation from a proton NMR signal and should provide
2351     the best estimate of the NMR relaxation time constant.\cite{Impey82}
2352    
2353     \begin{figure}
2354     \centering
2355 chrisfen 2975 \includegraphics[width=3.5in]{./figures/t5peDynamics.pdf}
2356 chrisfen 2973 \caption{Diffusion constants ({\it upper}) and reorientational time
2357     constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
2358     technique compared with experiment. Data at temperatures less that
2359     0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
2360     easier comparisons in the more relevant temperature regime.}
2361     \label{fig:t5peDynamics}
2362     \end{figure}
2363     Results for the diffusion constants and reorientational time constants
2364     are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
2365     apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
2366     the Ewald sum are reproduced with the {\sc sf} technique. The enhanced
2367     diffusion at high temperatures are again the product of the lower
2368     densities in comparison with experiment and do not provide any special
2369     insight into differences between the electrostatic summation
2370     techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
2371     diffuse a little faster than with the Ewald sum; however, use of light
2372     to moderate damping results in indistinguishable $D$ values. Though not
2373     apparent in this figure, {\sc sf} values at the lowest temperature are
2374     approximately an order of magnitude lower than with Ewald. These
2375     values support the observation from section \ref{sec:t5peThermo} that
2376     there appeared to be a change to a more glassy-like phase with the
2377     {\sc sf} technique at these lower temperatures.
2378    
2379     The $\tau_2^y$ results in the lower frame of figure
2380     \ref{fig:t5peDynamics} show a much greater difference between the {\sc
2381     sf} results and the Ewald results. At all temperatures shown, TIP5P-E
2382     relaxes faster than experiment with the Ewald sum while tracking
2383     experiment fairly well when using the {\sc sf} technique, independent
2384     of the choice of damping constant. Their are several possible reasons
2385     for this deviation between techniques. The Ewald results were taken
2386     shorter (10ps) trajectories than the {\sc sf} results (25ps). A quick
2387     calculation from a 10ps trajectory with {\sc sf} with an $\alpha$ of
2388     0.2\AA$^-1$ at 25$^\circ$C showed a 0.4ps drop in $\tau_2^y$, placing
2389     the result more in line with that obtained using the Ewald sum. These
2390     results support this explanation; however, recomputing the results to
2391     meet a poorer statistical standard is counter-productive. Assuming the
2392     Ewald results are not the product of poor statistics, differences in
2393     techniques to integrate the orientational motion could also play a
2394     role. {\sc shake} is the most commonly used technique for
2395     approximating rigid-body orientational motion,\cite{Ryckaert77} where
2396     as in {\sc oopse}, we maintain and integrate the entire rotation
2397     matrix using the {\sc dlm} method.\cite{Meineke05} Since {\sc shake}
2398     is an iterative constraint technique, if the convergence tolerances
2399     are raised for increased performance, error will accumulate in the
2400     orientational motion. Finally, the Ewald results were calculated using
2401     the $NVT$ ensemble, while the $NVE$ ensemble was used for {\sc sf}
2402     calculations. The additional mode of motion due to the thermostat will
2403     alter the dynamics, resulting in differences between $NVT$ and $NVE$
2404     results. These differences are increasingly noticeable as the
2405     thermostat time constant decreases.
2406    
2407     \section{Damping of Point Multipoles}\label{sec:dampingMultipoles}
2408    
2409     As discussed above, the {\sc sp} and {\sc sf} methods operate by
2410     neutralizing the cutoff sphere with charge-charge interaction shifting
2411     and by damping the electrostatic interactions. Now we would like to
2412     consider an extension of these techniques to include point multipole
2413     interactions. How will the shifting and damping need to develop in
2414     order to accommodate point multipoles?
2415    
2416     Of the two techniques, the least to vary is shifting. Shifting is
2417     employed to neutralize the cutoff sphere; however, in a system
2418     composed purely of point multipoles, the cutoff sphere is already
2419     neutralized. This means that shifting is not necessary between point
2420     multipoles. In a mixed system of monopoles and multipoles, the
2421     undamped {\sc sf} potential needs only to shift the force terms of the
2422     monopole (and use the monopole potential of equation (\ref{eq:SFPot}))
2423     and smoothly cutoff the multipole interactions with a switching
2424     function. The switching function is required in order to conserve
2425     energy, because a discontinuity will exist at $R_\textrm{c}$ in the
2426     absence of shifting terms.
2427    
2428     If we consider damping the {\sc sf} potential (Eq. (\ref{eq:DSFPot})),
2429     then we need to incorporate the complimentary error function term into
2430     the multipole potentials. The most direct way to do this is by
2431     replacing $r^{-1}$ with erfc$(\alpha r)\cdot r^{-1}$ in the multipole
2432     expansion.\cite{Hirschfelder67} In the multipole expansion, rather
2433     than considering only the interactions between single point charges,
2434     the electrostatic interactions is reformulated such that it describes
2435     the interaction between charge distributions about central sites of
2436     the respective sets of charges. This procedure is what leads to the
2437     familiar charge-dipole,
2438     \begin{equation}
2439     V_\textrm{cd} = -q_i\frac{\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}{r^3_{ij}}
2440     = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}},
2441     \label{eq:chargeDipole}
2442     \end{equation}
2443     and dipole-dipole,
2444     \begin{equation}
2445     V_\textrm{dd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
2446     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} -
2447     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}},
2448     \label{eq:dipoleDipole}
2449     \end{equation}
2450     interaction potentials.
2451    
2452     Using the charge-dipole interaction as an example, if we insert
2453     erfc$(\alpha r)\cdot r^{-1}$ in place of $r^{-1}$, a damped
2454     charge-dipole results,
2455     \begin{equation}
2456     V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
2457     \label{eq:dChargeDipole}
2458     \end{equation}
2459     where $c_1(r_{ij})$ is
2460     \begin{equation}
2461     c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
2462     + \textrm{erfc}(\alpha r_{ij}).
2463     \label{eq:c1Func}
2464     \end{equation}
2465     Thus, $c_1(r_{ij})$ is the resulting damping term that modifies the
2466     standard charge-dipole potential (Eq. (\ref{eq:chargeDipole})). Note
2467     that this damping term is dependent upon distance and not upon
2468     orientation, and that it is acting on what was originally an
2469 chrisfen 2975 $r^{-3}$ function. By writing the damped form in this manner, we
2470 chrisfen 2973 can collect the damping into one function and apply it to the original
2471     potential when damping is desired. This works well for potentials that
2472     have only one $r^{-n}$ term (where $n$ is an odd positive integer);
2473     but in the case of the dipole-dipole potential, there is one part
2474     dependent on $r^{-3}$ and another dependent on $r^{-5}$. In order to
2475     properly damping this potential, each of these parts is dampened with
2476     separate damping functions. We can determine the necessary damping
2477     functions by continuing with the multipole expansion; however, it
2478     quickly becomes more complex with ``two-center'' systems, like the
2479     dipole-dipole potential, and is typically approached with a spherical
2480     harmonic formalism.\cite{Hirschfelder67} A simpler method for
2481     determining these functions arises from adopting the tensor formalism
2482     for expressing the electrostatic interactions.\cite{Stone02}
2483    
2484     The tensor formalism for electrostatic interactions involves obtaining
2485     the multipole interactions from successive gradients of the monopole
2486     potential. Thus, tensors of rank one through three are
2487     \begin{equation}
2488     T = \frac{1}{4\pi\epsilon_0r_{ij}},
2489     \label{eq:tensorRank1}
2490     \end{equation}
2491     \begin{equation}
2492     T_\alpha = \frac{1}{4\pi\epsilon_0}\nabla_\alpha \frac{1}{r_{ij}},
2493     \label{eq:tensorRank2}
2494     \end{equation}
2495     \begin{equation}
2496     T_{\alpha\beta} = \frac{1}{4\pi\epsilon_0}
2497     \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
2498     \label{eq:tensorRank3}
2499     \end{equation}
2500     where the form of the first tensor gives the monopole-monopole
2501     potential, the second gives the monopole-dipole potential, and the
2502     third gives the monopole-quadrupole and dipole-dipole
2503     potentials.\cite{Stone02} Since the force is $-\nabla V$, the forces
2504     for each potential come from the next higher tensor.
2505    
2506     To obtain the damped electrostatic forms, we replace $r^{-1}$ with
2507     erfc$(\alpha r)\cdot r^{-1}$. Equation \ref{eq:tensorRank2} generates
2508     $c_1(r_{ij})$, just like the multipole expansion, while equation
2509     \ref{eq:tensorRank3} generates $c_2(r_{ij})$, where
2510     \begin{equation}
2511     c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
2512     + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
2513     + \textrm{erfc}(\alpha r_{ij}).
2514     \end{equation}
2515     Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
2516     term. Continuing with higher rank tensors, we can obtain the damping
2517     functions for higher multipoles as well as the forces. Each subsequent
2518     damping function includes one additional term, and we can simplify the
2519     procedure for obtaining these terms by writing out the following
2520     generating function,
2521     \begin{equation}
2522     c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
2523 chrisfen 2975 {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
2524 chrisfen 2973 \label{eq:dampingGeneratingFunc}
2525     \end{equation}
2526     where,
2527     \begin{equation}
2528     m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
2529     m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
2530     m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
2531     1 & m = -1\textrm{ or }0,
2532     \end{array}\right.
2533     \label{eq:doubleFactorial}
2534     \end{equation}
2535     and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
2536     is similar in form to those obtained by researchers for the
2537     application of the Ewald sum to
2538     multipoles.\cite{Smith82,Smith98,Aguado03}
2539    
2540     Returning to the dipole-dipole example, the potential consists of a
2541     portion dependent upon $r^{-5}$ and another dependent upon
2542     $r^{-3}$. In the damped dipole-dipole potential,
2543     \begin{equation}
2544     V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
2545     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
2546     c_2(r_{ij}) -
2547     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
2548     c_1(r_{ij}),
2549     \label{eq:dampDipoleDipole}
2550     \end{equation}
2551     $c_2(r_{ij})$ and $c_1(r_{ij})$ respectively dampen these two
2552     parts. The forces for the damped dipole-dipole interaction,
2553     \begin{equation}
2554     \begin{split}
2555     F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
2556     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
2557     c_3(r_{ij})\\ &-
2558     3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
2559     \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
2560     \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
2561     {r^5_{ij}} c_2(r_{ij}),
2562     \end{split}
2563     \label{eq:dampDipoleDipoleForces}
2564     \end{equation}
2565     rely on higher order damping functions because we perform another
2566     gradient operation. In this manner, we can dampen higher order
2567     multipolar interactions along with the monopole interactions, allowing
2568     us to include multipoles in simulations involving damped electrostatic
2569     interactions.
2570    
2571    
2572     \section{Damping and Dielectric Constants}\label{sec:dampingDielectric}
2573    
2574     In section \ref{sec:t5peThermo}, we observed that the choice of
2575     damping coefficient plays a major role in the calculated dielectric
2576     constant. This is not too surprising given the results for damping
2577     parameter influence on the long-time correlated motions of the NaCl
2578     crystal in section \ref{sec:LongTimeDynamics}. The static dielectric
2579     constant is calculated from the long-time fluctuations of the system's
2580     accumulated dipole moment (Eq. (\ref{eq:staticDielectric})), so it is
2581     going to be quite sensitive to the choice of damping parameter. We
2582     would like to choose an optimal damping constant for any particular
2583     cutoff radius choice that would properly capture the dielectric
2584     behavior of the liquid.
2585    
2586     In order to find these optimal values, we mapped out the static
2587     dielectric constant as a function of both the damping parameter and
2588     cutoff radius for several different water models. To calculate the
2589 chrisfen 2975 static dielectric constant, we performed 5ns $NPT$ calculations on
2590     systems of 512 water molecules, using the TIP5P-E, TIP4P-Ew, SPC/E,
2591 chrisfen 2973 and SSD/RF water models. TIP4P-Ew is a reparametrized version of the
2592     four-point transferable intermolecular potential (TIP4P) for water
2593     targeted for use with the Ewald summation.\cite{Horn04} SSD/RF is the
2594     reaction field modified variant of the soft sticky dipole (SSD) model
2595 chrisfen 2975 for water\cite{Fennell04} This model is discussed in more detail in
2596     the next chapter. One thing to note about it, electrostatic
2597     interactions are handled via dipole-dipole interactions rather than
2598     charge-charge interactions like the other three models. Damping of the
2599     dipole-dipole interaction was handled as described in section
2600     \ref{sec:dampingMultipoles}. Each of these systems were studied with
2601     cutoff radii of 9, 10, 11, and 12\AA\ and with damping parameter values
2602     ranging from 0 to 0.35\AA$^{-1}$.
2603 chrisfen 2973 \begin{figure}
2604     \centering
2605 chrisfen 2975 \includegraphics[width=\linewidth]{./figures/dielectricMap.pdf}
2606     \caption{The static dielectric constant for the TIP5P-E (A), TIP4P-Ew
2607     (B), SPC/E (C), and SSD/RF (D) water models as a function of cutoff
2608     radius ($R_\textrm{c}$) and damping coefficient ($\alpha$).}
2609 chrisfen 2974 \label{fig:dielectricMap}
2610 chrisfen 2973 \end{figure}
2611    
2612 chrisfen 2974 The results of these calculations are displayed in figure
2613     \ref{fig:dielectricMap} in the form of shaded contour plots. An
2614     interesting aspect of all four contour plots is that the dielectric
2615     constant is effectively linear with respect to $\alpha$ and
2616 chrisfen 2975 $R_\textrm{c}$ in the low to moderate damping regions, and the slope
2617     is 0.025\AA$^{-1}\cdot R_\textrm{c}^{-1}$. Another point to note is
2618     that choosing $\alpha$ and $R_\textrm{c}$ identical to those used in
2619     studies with the Ewald summation results in the same calculated
2620     dielectric constant. As an example, in the paper outlining the
2621     development of TIP5P-E, the real-space cutoff and Ewald coefficient
2622     were tethered to the system size, and for a 512 molecule system are
2623     approximately 12\AA\ and 0.25\AA$^{-1}$ respectively.\cite{Rick04}
2624     These parameters resulted in a dielectric constant of 92$\pm$14, while
2625     with {\sc sf} these parameters give a dielectric constant of
2626     90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
2627     $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5\AA\ and
2628     0.35\AA$^{-1}$, and these parameters resulted in a $\epsilon_0$ equal
2629     to 63$\pm$1.\cite{Horn04} We did not perform calculations with these
2630     exact parameters, but interpolating between surrounding values gives a
2631     $\epsilon_0$ of 61$\pm$1. Seeing a dependence of the dielectric
2632     constant on $\alpha$ and $R_\textrm{c}$ with the {\sc sf} technique,
2633     it might be interesting to investigate the dielectric dependence of
2634     the real-space Ewald parameters.
2635 chrisfen 2974
2636 chrisfen 2975 Although it is tempting to choose damping parameters equivalent to
2637     these Ewald examples, the results discussed in sections
2638     \ref{sec:EnergyResults} through \ref{sec:IndividualResults} indicate
2639     that values this high are destructive to both the energetics and
2640     dynamics. Ideally, $\alpha$ should not exceed 0.3\AA$^{-1}$ for any of
2641     the cutoff values in this range. If the optimal damping parameter is
2642     chosen to be midway between 0.275 and 0.3\AA$^{-1}$ (0.2875\AA$^{-1}$)
2643     at the 9\AA\ cutoff, then the linear trend with $R_\textrm{c}$ will
2644     always keep $\alpha$ below 0.3\AA$^{-1}$. This linear progression
2645     would give values of 0.2875, 0.2625, 0.2375, and 0.2125\AA$^{-1}$ for
2646     cutoff radii of 9, 10, 11, and 12\AA. Setting this to be the default
2647     behavior for the damped {\sc sf} technique will result in consistent
2648     dielectric behavior for these and other condensed molecular systems,
2649     regardless of the chosen cutoff radius. The static dielectric
2650     constants for TIP5P-E, TIP4P-Ew, SPC/E, and SSD/RF will be
2651     approximately fixed at 74, 52, 58, and 89 respectively. These values
2652     are generally lower than the values reported in the literature;
2653     however, the relative dielectric behavior scales as expected when
2654     comparing the models to one another.
2655 chrisfen 2974
2656 chrisfen 2973 \section{Conclusions}\label{sec:PairwiseConclusions}
2657    
2658     The above investigation of pairwise electrostatic summation techniques
2659     shows that there are viable and computationally efficient alternatives
2660     to the Ewald summation. These methods are derived from the damped and
2661     cutoff-neutralized Coulombic sum originally proposed by Wolf
2662     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
2663     method, reformulated above as equations (\ref{eq:DSFPot}) and
2664     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
2665     energetic and dynamic characteristics exhibited by simulations
2666     employing lattice summation techniques. The cumulative energy
2667     difference results showed the undamped {\sc sf} and moderately damped
2668 chrisfen 2975 {\sc sp} methods produced results nearly identical to the Ewald
2669     summation. Similarly for the dynamic features, the undamped or
2670     moderately damped {\sc sf} and moderately damped {\sc sp} methods
2671     produce force and torque vector magnitude and directions very similar
2672     to the expected values. These results translate into long-time
2673     dynamic behavior equivalent to that produced in simulations using the
2674     Ewald summation. A detailed study of water simulations showed that
2675     liquid properties calculated when using {\sc sf} will also be
2676     equivalent to those obtained using the Ewald summation.
2677 chrisfen 2973
2678     As in all purely-pairwise cutoff methods, these methods are expected
2679     to scale approximately {\it linearly} with system size, and they are
2680     easily parallelizable. This should result in substantial reductions
2681     in the computational cost of performing large simulations.
2682    
2683     Aside from the computational cost benefit, these techniques have
2684     applicability in situations where the use of the Ewald sum can prove
2685     problematic. Of greatest interest is their potential use in
2686     interfacial systems, where the unmodified lattice sum techniques
2687     artificially accentuate the periodicity of the system in an
2688     undesirable manner. There have been alterations to the standard Ewald
2689     techniques, via corrections and reformulations, to compensate for
2690     these systems; but the pairwise techniques discussed here require no
2691     modifications, making them natural tools to tackle these problems.
2692     Additionally, this transferability gives them benefits over other
2693     pairwise methods, like reaction field, because estimations of physical
2694     properties (e.g. the dielectric constant) are unnecessary.
2695    
2696     If a researcher is using Monte Carlo simulations of large chemical
2697     systems containing point charges, most structural features will be
2698     accurately captured using the undamped {\sc sf} method or the {\sc sp}
2699     method with an electrostatic damping of 0.2\AA$^{-1}$. These methods
2700     would also be appropriate for molecular dynamics simulations where the
2701     data of interest is either structural or short-time dynamical
2702     quantities. For long-time dynamics and collective motions, the safest
2703     pairwise method we have evaluated is the {\sc sf} method with an
2704 chrisfen 2975 electrostatic damping between 0.2 and 0.25\AA$^{-1}$. It is also
2705     important to note that the static dielectric constant in water
2706     simulations is highly dependent on both $\alpha$ and
2707     $R_\textrm{c}$. For consistent dielectric behavior, the damped {\sc
2708     sf} method should use an $\alpha$ of 0.2175\AA$^{-1}$ for an
2709     $R_\textrm{c}$ of 12\AA, and $\alpha$ should decrease by
2710     0.025\AA$^{-1}$ for every 1\AA\ increase in cutoff radius.
2711 chrisfen 2973
2712     We are not suggesting that there is any flaw with the Ewald sum; in
2713     fact, it is the standard by which these simple pairwise sums have been
2714     judged. However, these results do suggest that in the typical
2715     simulations performed today, the Ewald summation may no longer be
2716     required to obtain the level of accuracy most researchers have come to
2717     expect.