ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chrisDissertation/Electrostatics.tex
Revision: 3023
Committed: Tue Sep 26 03:07:59 2006 UTC (17 years, 11 months ago) by chrisfen
Content type: application/x-tex
File size: 96315 byte(s)
Log Message:
Refinements.  Just a little proof checking left...

File Contents

# User Rev Content
1 chrisfen 3001 \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION \\ TECHNIQUES}
2 chrisfen 2987
3 chrisfen 3023 In molecular simulations, proper accumulation of electrostatic
4 chrisfen 2987 interactions is essential and is one of the most
5     computationally-demanding tasks. The common molecular mechanics force
6     fields represent atomic sites with full or partial charges protected
7 chrisfen 3023 by repulsive Lennard-Jones interactions. This means that nearly every
8     pair interaction involves a calculation of charge-charge forces.
9 chrisfen 2987 Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
10     interactions quickly become the most expensive part of molecular
11     simulations. Historically, the electrostatic pair interaction would
12     not have decayed appreciably within the typical box lengths that could
13     be feasibly simulated. In the larger systems that are more typical of
14     modern simulations, large cutoffs should be used to incorporate
15     electrostatics correctly.
16    
17     There have been many efforts to address the proper and practical
18     handling of electrostatic interactions, and these have resulted in a
19     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
20     typically classified as implicit methods (i.e., continuum dielectrics,
21     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
22     (i.e., Ewald summations, interaction shifting or
23     truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
24     reaction field type methods, fast multipole
25     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
26     often preferred because they physically incorporate solvent molecules
27     in the system of interest, but these methods are sometimes difficult
28     to utilize because of their high computational cost.\cite{Roux99} In
29     addition to the computational cost, there have been some questions
30     regarding possible artifacts caused by the inherent periodicity of the
31     explicit Ewald summation.\cite{Tobias01}
32    
33     In this chapter, we focus on a new set of pairwise methods devised by
34     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
35     methods along with a few other mixed methods (i.e. reaction field) are
36     compared with the smooth particle mesh Ewald
37     sum,\cite{Onsager36,Essmann99} which is our reference method for
38     handling long-range electrostatic interactions. The new methods for
39     handling electrostatics have the potential to scale linearly with
40     increasing system size since they involve only a simple modification
41     to the direct pairwise sum. They also lack the added periodicity of
42     the Ewald sum, so they can be used for systems which are non-periodic
43     or which have one- or two-dimensional periodicity. Below, these
44     methods are evaluated using a variety of model systems to
45     establish their usability in molecular simulations.
46    
47     \section{The Ewald Sum}
48    
49     The complete accumulation of the electrostatic interactions in a system with
50     periodic boundary conditions (PBC) requires the consideration of the
51     effect of all charges within a (cubic) simulation box as well as those
52     in the periodic replicas,
53     \begin{equation}
54     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
55     \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
56     \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
57     \right],
58     \label{eq:PBCSum}
59     \end{equation}
60     where the sum over $\mathbf{n}$ is a sum over all periodic box
61     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
62     prime indicates $i = j$ are neglected for $\mathbf{n} =
63     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
64     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
65     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
66     $j$, and $\phi$ is the solution to Poisson's equation
67     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
68     charge-charge interactions). In the case of monopole electrostatics,
69     equation (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
70     non-neutral systems.
71    
72     The electrostatic summation problem was originally studied by Ewald
73     for the case of an infinite crystal.\cite{Ewald21}. The approach he
74     took was to convert this conditionally convergent sum into two
75     absolutely convergent summations: a short-ranged real-space summation
76     and a long-ranged reciprocal-space summation,
77     \begin{equation}
78     \begin{split}
79     V_\textrm{elec} = \frac{1}{2}&
80     \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
81     \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
82     {|\mathbf{r}_{ij}+\mathbf{n}|} \\
83     &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
84     \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
85     \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
86     &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
87     + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
88     \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
89     \end{split}
90     \label{eq:EwaldSum}
91     \end{equation}
92     where $\alpha$ is the damping or convergence parameter with units of
93     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
94     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
95     constant of the surrounding medium. The final two terms of
96     equation (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
97     for interacting with a surrounding dielectric.\cite{Allen87} This
98     dipolar term was neglected in early applications in molecular
99     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
100     Leeuw {\it et al.} to address situations where the unit cell has a
101     dipole moment which is magnified through replication of the periodic
102     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
103     system is said to be using conducting (or ``tin-foil'') boundary
104 chrisfen 3023 conditions, $\epsilon_{\rm S} = \infty$.
105 chrisfen 2987
106     \begin{figure}
107     \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
108     \caption{The change in the need for the Ewald sum with
109     increasing computational power. A:~Initially, only small systems
110     could be studied, and the Ewald sum replicated the simulation box to
111     convergence. B:~Now, radial cutoff methods should be able to reach
112     convergence for the larger systems of charges that are common today.}
113     \label{fig:ewaldTime}
114     \end{figure}
115 chrisfen 3023 Figure \ref{fig:ewaldTime} shows how the Ewald sum has been applied
116     over time. Initially, due to the small system sizes that could be
117     simulated feasibly, the entire simulation box was replicated to
118     convergence. In more modern simulations, the systems have grown large
119     enough that a real-space cutoff could potentially give convergent
120     behavior. Indeed, it has been observed that with the choice of a
121     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
122     rapidly convergent and small relative to the real-space
123     portion.\cite{Karasawa89,Kolafa92}
124 chrisfen 2987
125 chrisfen 3001 The original Ewald summation is an $\mathcal{O}(N^2)$ algorithm. The
126 chrisfen 2987 convergence parameter $(\alpha)$ plays an important role in balancing
127     the computational cost between the direct and reciprocal-space
128     portions of the summation. The choice of this value allows one to
129     select whether the real-space or reciprocal space portion of the
130 chrisfen 3001 summation is an $\mathcal{O}(N^2)$ calculation (with the other being
131     $\mathcal{O}(N)$).\cite{Sagui99} With the appropriate choice of
132 chrisfen 2987 $\alpha$ and thoughtful algorithm development, this cost can be
133 chrisfen 3001 reduced to $\mathcal{O}(N^{3/2})$.\cite{Perram88} The typical route
134 chrisfen 2987 taken to reduce the cost of the Ewald summation even further is to set
135     $\alpha$ such that the real-space interactions decay rapidly, allowing
136     for a short spherical cutoff. Then the reciprocal space summation is
137     optimized. These optimizations usually involve utilization of the
138     fast Fourier transform (FFT),\cite{Hockney81} leading to the
139     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
140     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
141     methods, the cost of the reciprocal-space portion of the Ewald
142 chrisfen 3001 summation is reduced from $\mathcal{O}(N^2)$ down to $\mathcal{O}(N
143 chrisfen 2987 \log N)$.
144    
145     These developments and optimizations have made the use of the Ewald
146     summation routine in simulations with periodic boundary
147     conditions. However, in certain systems, such as vapor-liquid
148     interfaces and membranes, the intrinsic three-dimensional periodicity
149     can prove problematic. The Ewald sum has been reformulated to handle
150     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
151     methods are computationally expensive.\cite{Spohr97,Yeh99} More
152     recently, there have been several successful efforts toward reducing
153     the computational cost of 2-D lattice
154     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
155     bringing them more in line with the cost of the full 3-D summation.
156    
157     Several studies have recognized that the inherent periodicity in the
158     Ewald sum can also have an effect on three-dimensional
159     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
160     Solvated proteins are essentially kept at high concentration due to
161     the periodicity of the electrostatic summation method. In these
162     systems, the more compact folded states of a protein can be
163     artificially stabilized by the periodic replicas introduced by the
164     Ewald summation.\cite{Weber00} Thus, care must be taken when
165     considering the use of the Ewald summation where the assumed
166 chrisfen 3016 periodicity would introduce spurious effects.
167 chrisfen 2987
168    
169     \section{The Wolf and Zahn Methods}
170    
171     In a recent paper by Wolf \textit{et al.}, a procedure was outlined
172     for the accurate accumulation of electrostatic interactions in an
173     efficient pairwise fashion. This procedure lacks the inherent
174     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
175     observed that the electrostatic interaction is effectively
176     short-ranged in condensed phase systems and that neutralization of the
177     charge contained within the cutoff radius is crucial for potential
178     stability. They devised a pairwise summation method that ensures
179     charge neutrality and gives results similar to those obtained with the
180     Ewald summation. The resulting shifted Coulomb potential includes
181     image-charges subtracted out through placement on the cutoff sphere
182     and a distance-dependent damping function (identical to that seen in
183     the real-space portion of the Ewald sum) to aid convergence
184     \begin{equation}
185     V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
186     - \lim_{r_{ij}\rightarrow R_\textrm{c}}
187     \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
188     \label{eq:WolfPot}
189     \end{equation}
190     Equation (\ref{eq:WolfPot}) is essentially the common form of a shifted
191     potential. However, neutralizing the charge contained within each
192     cutoff sphere requires the placement of a self-image charge on the
193     surface of the cutoff sphere. This additional self-term in the total
194     potential enabled Wolf {\it et al.} to obtain excellent estimates of
195     Madelung energies for many crystals.
196    
197     In order to use their charge-neutralized potential in molecular
198     dynamics simulations, Wolf \textit{et al.} suggested taking the
199     derivative of this potential prior to evaluation of the limit. This
200     procedure gives an expression for the forces,
201     \begin{equation}
202     \begin{split}
203     F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
204     \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
205     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
206     \Biggr]\\
207     &-\Biggr[
208     \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
209     + \frac{2\alpha}{\pi^{1/2}}
210     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
211     \Biggr]\Biggr\},
212     \end{split}
213     \label{eq:WolfForces}
214     \end{equation}
215     that incorporates both image charges and damping of the electrostatic
216     interaction.
217    
218     More recently, Zahn \textit{et al.} investigated these potential and
219     force expressions for use in simulations involving water.\cite{Zahn02}
220     In their work, they pointed out that the forces and derivative of
221     the potential are not commensurate. Attempts to use both
222     equations (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
223     to poor energy conservation. They correctly observed that taking the
224     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
225     derivatives gives forces for a different potential energy function
226     than the one shown in equation (\ref{eq:WolfPot}).
227    
228     Zahn \textit{et al.} introduced a modified form of this summation
229     method as a way to use the technique in Molecular Dynamics
230     simulations. They proposed a new damped Coulomb potential,
231     \begin{equation}
232     \begin{split}
233     V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
234     \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
235     &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
236     + \frac{2\alpha}{\pi^{1/2}}
237     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
238     \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
239     \end{split}
240     \label{eq:ZahnPot}
241     \end{equation}
242     and showed that this potential does fairly well at capturing the
243     structural and dynamic properties of water compared the same
244     properties obtained using the Ewald sum.
245    
246     \section{Simple Forms for Pairwise Electrostatics}\label{sec:PairwiseDerivation}
247    
248     The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
249     al.} are constructed using two different (and separable) computational
250     tricks:
251    
252     \begin{enumerate}[itemsep=0pt]
253     \item shifting through the use of image charges, and
254     \item damping the electrostatic interaction.
255     \end{enumerate}
256     Wolf \textit{et al.} treated the development of their summation method
257     as a progressive application of these techniques,\cite{Wolf99} while
258     Zahn \textit{et al.} founded their damped Coulomb modification
259     (Eq. (\ref{eq:ZahnPot})) on the post-limit forces
260     (Eq. (\ref{eq:WolfForces})) which were derived using both techniques.
261     It is possible, however, to separate these tricks and study their
262     effects independently.
263    
264     Starting with the original observation that the effective range of the
265     electrostatic interaction in condensed phases is considerably less
266     than $r^{-1}$, either the cutoff sphere neutralization or the
267     distance-dependent damping technique could be used as a foundation for
268     a new pairwise summation method. Wolf \textit{et al.} made the
269     observation that charge neutralization within the cutoff sphere plays
270     a significant role in energy convergence; therefore we will begin our
271     analysis with the various shifted forms that maintain this charge
272     neutralization. We can evaluate the methods of Wolf {\it et al.} and
273     Zahn {\it et al.} by considering the standard shifted potential,
274     \begin{equation}
275     V_\textrm{SP}(r) = \begin{cases}
276     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
277     R_\textrm{c}
278     \end{cases},
279     \label{eq:shiftingPotForm}
280     \end{equation}
281     and shifted force,
282     \begin{equation}
283     V_\textrm{SF}(r) = \begin{cases}
284     v(r) - v_\textrm{c}
285     - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
286     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
287     \end{cases},
288     \label{eq:shiftingForm}
289     \end{equation}
290     functions where $v(r)$ is the unshifted form of the potential, and
291     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
292     that both the potential and the forces goes to zero at the cutoff
293     radius, while the Shifted Potential ({\sc sp}) form only ensures the
294     potential is smooth at the cutoff radius
295     ($R_\textrm{c}$).\cite{Allen87}
296    
297     The forces associated with the shifted potential are simply the forces
298     of the unshifted potential itself (when inside the cutoff sphere),
299     \begin{equation}
300     F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
301     \end{equation}
302     and are zero outside. Inside the cutoff sphere, the forces associated
303     with the shifted force form can be written,
304     \begin{equation}
305     F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
306     v(r)}{dr} \right)_{r=R_\textrm{c}}.
307     \end{equation}
308    
309     If the potential, $v(r)$, is taken to be the normal Coulomb potential,
310     \begin{equation}
311     v(r) = \frac{q_i q_j}{r},
312     \label{eq:Coulomb}
313     \end{equation}
314     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
315     al.}'s undamped prescription:
316     \begin{equation}
317     V_\textrm{SP}(r) =
318     q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
319     r\leqslant R_\textrm{c},
320     \label{eq:SPPot}
321     \end{equation}
322     with associated forces,
323     \begin{equation}
324     F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
325     \quad r\leqslant R_\textrm{c}.
326     \label{eq:SPForces}
327     \end{equation}
328     These forces are identical to the forces of the standard Coulomb
329     interaction, and cutting these off at $R_c$ was addressed by Wolf
330     \textit{et al.} as undesirable. They pointed out that the effect of
331     the image charges is neglected in the forces when this form is
332     used,\cite{Wolf99} thereby eliminating any benefit from the method in
333     molecular dynamics. Additionally, there is a discontinuity in the
334     forces at the cutoff radius which results in energy drift during MD
335     simulations.
336    
337     The shifted force ({\sc sf}) form using the normal Coulomb potential
338     will give,
339     \begin{equation}
340     V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
341     + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
342     \quad r\leqslant R_\textrm{c}.
343     \label{eq:SFPot}
344     \end{equation}
345     with associated forces,
346     \begin{equation}
347     F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
348     \quad r\leqslant R_\textrm{c}.
349     \label{eq:SFForces}
350     \end{equation}
351     This formulation has the benefits that there are no discontinuities at
352     the cutoff radius, while the neutralizing image charges are present in
353     both the energy and force expressions. It would be simple to add the
354     self-neutralizing term back when computing the total energy of the
355     system, thereby maintaining the agreement with the Madelung energies.
356     A side effect of this treatment is the alteration in the shape of the
357     potential that comes from the derivative term. Thus, a degree of
358     clarity about agreement with the empirical potential is lost in order
359     to gain functionality in dynamics simulations.
360    
361     Wolf \textit{et al.} originally discussed the energetics of the
362     shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
363     insufficient for accurate determination of the energy with reasonable
364     cutoff distances. The calculated Madelung energies fluctuated around
365     the expected value as the cutoff radius was increased, but the
366     oscillations converged toward the correct value.\cite{Wolf99} A
367     damping function was incorporated to accelerate the convergence; and
368     though alternative forms for the damping function could be
369     used,\cite{Jones56,Heyes81} the complimentary error function was
370     chosen to mirror the effective screening used in the Ewald summation.
371     Incorporating this error function damping into the simple Coulomb
372     potential,
373     \begin{equation}
374     v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
375     \label{eq:dampCoulomb}
376     \end{equation}
377     the shifted potential (Eq. (\ref{eq:SPPot})) becomes
378     \begin{equation}
379     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
380     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
381     \quad r\leqslant R_\textrm{c},
382     \label{eq:DSPPot}
383     \end{equation}
384     with associated forces,
385     \begin{equation}
386     F_{\textrm{DSP}}(r) = q_iq_j
387     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
388     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
389     \quad r\leqslant R_\textrm{c}.
390     \label{eq:DSPForces}
391     \end{equation}
392     Again, this damped shifted potential suffers from a
393     force-discontinuity at the cutoff radius, and the image charges play
394     no role in the forces. To remedy these concerns, one may derive a
395     {\sc sf} variant by including the derivative term in
396     equation (\ref{eq:shiftingForm}),
397     \begin{equation}
398     \begin{split}
399     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
400     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
401     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
402     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
403     + \frac{2\alpha}{\pi^{1/2}}
404     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
405     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
406     \quad r\leqslant R_\textrm{c}.
407     \label{eq:DSFPot}
408     \end{split}
409     \end{equation}
410     The derivative of the above potential will lead to the following forces,
411     \begin{equation}
412     \begin{split}
413     F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
414     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
415     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
416     &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
417     {R_{\textrm{c}}^2}
418     + \frac{2\alpha}{\pi^{1/2}}
419     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
420     \right)\Biggr{]}
421     \quad r\leqslant R_\textrm{c}.
422     \label{eq:DSFForces}
423     \end{split}
424     \end{equation}
425     If the damping parameter $(\alpha)$ is set to zero, the undamped case,
426     equations (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
427     recovered from equations (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
428    
429     This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
430     derived by Zahn \textit{et al.}; however, there are two important
431     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from equation
432     (\ref{eq:shiftingForm}) is equal to equation (\ref{eq:dampCoulomb})
433     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
434     in the Zahn potential, resulting in a potential discontinuity as
435     particles cross $R_\textrm{c}$. Second, the sign of the derivative
436     portion is different. The missing $v_\textrm{c}$ term would not
437     affect molecular dynamics simulations (although the computed energy
438     would be expected to have sudden jumps as particle distances crossed
439     $R_c$). The sign problem is a potential source of errors, however.
440     In fact, it introduces a discontinuity in the forces at the cutoff,
441     because the force function is shifted in the wrong direction and
442     doesn't cross zero at $R_\textrm{c}$.
443    
444     Equations (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
445     electrostatic summation method in which the potential and forces are
446     continuous at the cutoff radius and which incorporates the damping
447     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
448     this paper, we will evaluate exactly how good these methods ({\sc sp},
449     {\sc sf}, damping) are at reproducing the correct electrostatic
450     summation performed by the Ewald sum.
451    
452    
453     \section{Evaluating Pairwise Summation Techniques}
454    
455 chrisfen 3001 As mentioned in the introduction, there are two primary techniques
456     utilized to obtain information about the system of interest in
457     classical molecular mechanics simulations: Monte Carlo (MC) and
458     Molecular Dynamics (MD). Both of these techniques utilize pairwise
459     summations of interactions between particle sites, but they use these
460     summations in different ways.
461 chrisfen 2987
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 chrisfen 3023 \begin{figure}
485     \centering
486     \includegraphics[width = 3.5in]{./figures/dualLinear.pdf}
487     \caption{Example least squares regressions of the configuration energy
488     differences for SPC/E water systems. The upper plot shows a data set
489     with a poor correlation coefficient ($R^2$), while the lower plot
490     shows a data set with a good correlation coefficient.}
491     \label{fig:linearFit}
492     \end{figure}
493 chrisfen 2987 The pairwise summation techniques (outlined in section
494     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
495     studying the energy differences between conformations. We took the
496     {\sc spme}-computed energy difference between two conformations to be the
497     correct behavior. An ideal performance by an alternative method would
498     reproduce these energy differences exactly (even if the absolute
499     energies calculated by the methods are different). Since none of the
500     methods provide exact energy differences, we used linear least squares
501     regressions of energy gap data to evaluate how closely the methods
502     mimicked the Ewald energy gaps. Unitary results for both the
503     correlation (slope) and correlation coefficient for these regressions
504     indicate perfect agreement between the alternative method and {\sc spme}.
505     Sample correlation plots for two alternate methods are shown in
506     Fig. \ref{fig:linearFit}.
507    
508     Each of the seven system types (detailed in section \ref{sec:RepSims})
509     were represented using 500 independent configurations. Thus, each of
510     the alternative (non-Ewald) electrostatic summation methods was
511     evaluated using an accumulated 873,250 configurational energy
512     differences.
513    
514     Results and discussion for the individual analysis of each of the
515 chrisfen 3001 system types appear in appendix \ref{app:IndividualResults}, while the
516 chrisfen 2987 cumulative results over all the investigated systems appear below in
517     sections \ref{sec:EnergyResults}.
518    
519     \subsection{Molecular Dynamics and the Force and Torque
520     Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
521     (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
522     comparing the force and torque vectors with those obtained using the
523     reference Ewald summation ({\sc spme}). Both the magnitude and the
524     direction of these vectors on each of the bodies in the system were
525     analyzed. For the magnitude of these vectors, linear least squares
526     regression analyses were performed as described previously for
527     comparing $\Delta E$ values. Instead of a single energy difference
528     between two system configurations, we compared the magnitudes of the
529     forces (and torques) on each molecule in each configuration. For a
530     system of 1000 water molecules and 40 ions, there are 1040 force
531     vectors and 1000 torque vectors. With 500 configurations, this
532     results in 520,000 force and 500,000 torque vector comparisons.
533     Additionally, data from seven different system types was aggregated
534     before the comparison was made.
535    
536     The {\it directionality} of the force and torque vectors was
537     investigated through measurement of the angle ($\theta$) formed
538     between those computed from the particular method and those from {\sc spme},
539     \begin{equation}
540     \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
541     \cdot \hat{F}_\textrm{M}\right),
542     \end{equation}
543     where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
544     vector computed using method M. Each of these $\theta$ values was
545     accumulated in a distribution function and weighted by the area on the
546     unit sphere. Since this distribution is a measure of angular error
547     between two different electrostatic summation methods, there is no
548     {\it a priori} reason for the profile to adhere to any specific
549     shape. Thus, gaussian fits were used to measure the width of the
550     resulting distributions. The variance ($\sigma^2$) was extracted from
551     each of these fits and was used to compare distribution widths.
552     Values of $\sigma^2$ near zero indicate vector directions
553     indistinguishable from those calculated when using the reference
554     method ({\sc spme}).
555    
556     \subsection{Short-time Dynamics}
557    
558     The effects of the alternative electrostatic summation methods on the
559     short-time dynamics of charged systems were evaluated by considering a
560     NaCl crystal at a temperature of 1000~K. A subset of the best
561     performing pairwise methods was used in this comparison. The NaCl
562     crystal was chosen to avoid possible complications from the treatment
563     of orientational motion in molecular systems. All systems were
564     started with the same initial positions and velocities. Simulations
565     were performed under the microcanonical ensemble, and velocity
566     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
567     of the trajectories,
568     \begin{equation}
569     C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
570     \label{eq:vCorr}
571     \end{equation}
572     Velocity autocorrelation functions require detailed short time data,
573     thus velocity information was saved every 2~fs over 10~ps
574     trajectories. Because the NaCl crystal is composed of two different
575     atom types, the average of the two resulting velocity autocorrelation
576     functions was used for comparisons.
577    
578     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
579    
580     The effects of the same subset of alternative electrostatic methods on
581     the {\it long-time} dynamics of charged systems were evaluated using
582     the same model system (NaCl crystals at 1000K). The power spectrum
583     ($I(\omega)$) was obtained via Fourier transform of the velocity
584     autocorrelation function,
585     \begin{equation} I(\omega) =
586     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
587     \label{eq:powerSpec}
588     \end{equation}
589     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
590     NaCl crystal is composed of two different atom types, the average of
591     the two resulting power spectra was used for comparisons. Simulations
592     were performed under the microcanonical ensemble, and velocity
593     information was saved every 5~fs over 100~ps trajectories.
594    
595     \subsection{Representative Simulations}\label{sec:RepSims}
596     A variety of representative molecular simulations were analyzed to
597     determine the relative effectiveness of the pairwise summation
598     techniques in reproducing the energetics and dynamics exhibited by
599     {\sc spme}. We wanted to span the space of typical molecular
600     simulations (i.e. from liquids of neutral molecules to ionic
601     crystals), so the systems studied were:
602    
603     \begin{enumerate}[itemsep=0pt]
604     \item liquid water (SPC/E),\cite{Berendsen87}
605     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
606     \item NaCl crystals,
607     \item NaCl melts,
608     \item a low ionic strength solution of NaCl in water (0.11 M),
609     \item a high ionic strength solution of NaCl in water (1.1 M), and
610     \item a 6~\AA\ radius sphere of Argon in water.
611     \end{enumerate}
612    
613     By utilizing the pairwise techniques (outlined in section
614     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
615     charged particles, and mixtures of the two, we hope to discern under
616     which conditions it will be possible to use one of the alternative
617     summation methodologies instead of the Ewald sum.
618    
619     For the solid and liquid water configurations, configurations were
620     taken at regular intervals from high temperature trajectories of 1000
621     SPC/E water molecules. Each configuration was equilibrated
622     independently at a lower temperature (300~K for the liquid, 200~K for
623     the crystal). The solid and liquid NaCl systems consisted of 500
624     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
625     these systems were selected and equilibrated in the same manner as the
626     water systems. In order to introduce measurable fluctuations in the
627     configuration energy differences, the crystalline simulations were
628     equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid
629     NaCl configurations needed to represent a fully disordered array of
630     point charges, so the high temperature of 7000~K was selected for
631     equilibration. The ionic solutions were made by solvating 4 (or 40)
632     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
633     water positions were then randomly swapped, and the resulting
634     configurations were again equilibrated individually. Finally, for the
635     Argon / Water ``charge void'' systems, the identities of all the SPC/E
636     waters within 6~\AA\ of the center of the equilibrated water
637     configurations were converted to argon.
638    
639     These procedures guaranteed us a set of representative configurations
640     from chemically-relevant systems sampled from appropriate
641     ensembles. Force field parameters for the ions and Argon were taken
642     from the force field utilized by {\sc oopse}.\cite{Meineke05}
643    
644     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
645     We compared the following alternative summation methods with results
646     from the reference method ({\sc spme}):
647    
648     \begin{enumerate}[itemsep=0pt]
649     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
650     and 0.3~\AA$^{-1}$,
651     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
652     and 0.3~\AA$^{-1}$,
653     \item reaction field with an infinite dielectric constant, and
654     \item an unmodified cutoff.
655     \end{enumerate}
656    
657     Group-based cutoffs with a fifth-order polynomial switching function
658     were utilized for the reaction field simulations. Additionally, we
659     investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
660     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
661     implementation of {\sc spme},\cite{Ponder87} while all other calculations
662     were performed using the {\sc oopse} molecular mechanics
663     package.\cite{Meineke05} All other portions of the energy calculation
664     (i.e. Lennard-Jones interactions) were handled in exactly the same
665     manner across all systems and configurations.
666    
667     The alternative methods were also evaluated with three different
668     cutoff radii (9, 12, and 15~\AA). As noted previously, the
669     convergence parameter ($\alpha$) plays a role in the balance of the
670     real-space and reciprocal-space portions of the Ewald calculation.
671     Typical molecular mechanics packages set this to a value dependent on
672     the cutoff radius and a tolerance (typically less than $1 \times
673     10^{-4}$~kcal/mol). Smaller tolerances are typically associated with
674     increasing accuracy at the expense of computational time spent on the
675     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
676     The default {\sc tinker} tolerance of $1 \times 10^{-8}$~kcal/mol was used
677     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
678     0.3119, and 0.2476~\AA$^{-1}$ for cutoff radii of 9, 12, and 15~\AA\
679     respectively.
680    
681     \section{Configuration Energy Difference Results}\label{sec:EnergyResults}
682     In order to evaluate the performance of the pairwise electrostatic
683     summation methods for Monte Carlo (MC) simulations, the energy
684     differences between configurations were compared to the values
685     obtained when using {\sc spme}. The results for the combined
686     regression analysis of all of the systems are shown in figure
687     \ref{fig:delE}.
688    
689     \begin{figure}
690     \centering
691     \includegraphics[width=4.75in]{./figures/delEplot.pdf}
692     \caption{Statistical analysis of the quality of configurational energy
693     differences for a given electrostatic method compared with the
694     reference Ewald sum. Results with a value equal to 1 (dashed line)
695     indicate $\Delta E$ values indistinguishable from those obtained using
696     {\sc spme}. Different values of the cutoff radius are indicated with
697     different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ =
698     inverted triangles).}
699     \label{fig:delE}
700     \end{figure}
701     The most striking feature of this plot is how well the Shifted Force
702     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
703     differences. For the undamped {\sc sf} method, and the
704     moderately-damped {\sc sp} methods, the results are nearly
705     indistinguishable from the Ewald results. The other common methods do
706     significantly less well.
707    
708     The unmodified cutoff method is essentially unusable. This is not
709     surprising since hard cutoffs give large energy fluctuations as atoms
710     or molecules move in and out of the cutoff
711     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
712     some degree by using group based cutoffs with a switching
713     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
714     significant improvement using the group-switched cutoff because the
715 chrisfen 3001 salt and salt solution systems contain non-neutral groups. Appendix
716     \ref{app:IndividualResults} includes results for systems comprised
717     entirely of neutral groups.
718 chrisfen 2987
719     For the {\sc sp} method, inclusion of electrostatic damping improves
720     the agreement with Ewald, and using an $\alpha$ of 0.2~\AA $^{-1}$
721     shows an excellent correlation and quality of fit with the {\sc spme}
722 chrisfen 3023 results, particularly with a cutoff radius greater than 12~\AA . Use
723 chrisfen 2987 of a larger damping parameter is more helpful for the shortest cutoff
724     shown, but it has a detrimental effect on simulations with larger
725     cutoffs.
726    
727     In the {\sc sf} sets, increasing damping results in progressively {\it
728     worse} correlation with Ewald. Overall, the undamped case is the best
729     performing set, as the correlation and quality of fits are
730     consistently superior regardless of the cutoff distance. The undamped
731     case is also less computationally demanding (because no evaluation of
732     the complementary error function is required).
733    
734     The reaction field results illustrates some of that method's
735     limitations, primarily that it was developed for use in homogeneous
736     systems; although it does provide results that are an improvement over
737     those from an unmodified cutoff.
738    
739     \section{Magnitude of the Force and Torque Vector Results}\label{sec:FTMagResults}
740    
741     Evaluation of pairwise methods for use in Molecular Dynamics
742     simulations requires consideration of effects on the forces and
743     torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
744     regression results for the force and torque vector magnitudes,
745     respectively. The data in these figures was generated from an
746     accumulation of the statistics from all of the system types.
747    
748     \begin{figure}
749     \centering
750     \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
751     \caption{Statistical analysis of the quality of the force vector
752     magnitudes for a given electrostatic method compared with the
753     reference Ewald sum. Results with a value equal to 1 (dashed line)
754     indicate force magnitude values indistinguishable from those obtained
755     using {\sc spme}. Different values of the cutoff radius are indicated with
756     different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ =
757     inverted triangles).}
758     \label{fig:frcMag}
759     \end{figure}
760     Again, it is striking how well the Shifted Potential and Shifted Force
761     methods are doing at reproducing the {\sc spme} forces. The undamped and
762     weakly-damped {\sc sf} method gives the best agreement with Ewald.
763     This is perhaps expected because this method explicitly incorporates a
764     smooth transition in the forces at the cutoff radius as well as the
765     neutralizing image charges.
766    
767     Figure \ref{fig:frcMag}, for the most part, parallels the results seen
768     in the previous $\Delta E$ section. The unmodified cutoff results are
769     poor, but using group based cutoffs and a switching function provides
770     an improvement much more significant than what was seen with $\Delta
771     E$.
772    
773     With moderate damping and a large enough cutoff radius, the {\sc sp}
774     method is generating usable forces. Further increases in damping,
775     while beneficial for simulations with a cutoff radius of 9~\AA\ , is
776     detrimental to simulations with larger cutoff radii.
777    
778     The reaction field results are surprisingly good, considering the poor
779     quality of the fits for the $\Delta E$ results. There is still a
780     considerable degree of scatter in the data, but the forces correlate
781     well with the Ewald forces in general. We note that the reaction
782     field calculations do not include the pure NaCl systems, so these
783     results are partly biased towards conditions in which the method
784     performs more favorably.
785    
786     \begin{figure}
787     \centering
788     \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
789     \caption{Statistical analysis of the quality of the torque vector
790     magnitudes for a given electrostatic method compared with the
791     reference Ewald sum. Results with a value equal to 1 (dashed line)
792     indicate torque magnitude values indistinguishable from those obtained
793     using {\sc spme}. Different values of the cutoff radius are indicated with
794     different symbols (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ =
795     inverted triangles).}
796     \label{fig:trqMag}
797     \end{figure}
798     Molecular torques were only available from the systems which contained
799     rigid molecules (i.e. the systems containing water). The data in
800     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
801    
802     Torques appear to be much more sensitive to charges at a longer
803     distance. The striking feature in comparing the new electrostatic
804     methods with {\sc spme} is how much the agreement improves with increasing
805     cutoff radius. Again, the weakly damped and undamped {\sc sf} method
806 chrisfen 3023 appears to reproduce the {\sc spme} torques most accurately.
807 chrisfen 2987
808     Water molecules are dipolar, and the reaction field method reproduces
809     the effect of the surrounding polarized medium on each of the
810     molecular bodies. Therefore it is not surprising that reaction field
811     performs best of all of the methods on molecular torques.
812    
813     \section{Directionality of the Force and Torque Vector Results}\label{sec:FTDirResults}
814    
815 chrisfen 3023 It is clearly important that a new electrostatic method should be able
816     to reproduce the magnitudes of the force and torque vectors obtained
817     via the Ewald sum. However, the {\it directionality} of these vectors
818     will also be vital in calculating dynamical quantities accurately.
819     Force and torque directionalities were investigated by measuring the
820     angles formed between these vectors and the same vectors calculated
821     using {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared
822     through the variance ($\sigma^2$) of the Gaussian fits of the angle
823     error distributions of the combined set over all system types.
824 chrisfen 2987
825     \begin{figure}
826     \centering
827     \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
828     \caption{Statistical analysis of the width of the angular distribution
829     that the force and torque vectors from a given electrostatic method
830     make with their counterparts obtained using the reference Ewald sum.
831     Results with a variance ($\sigma^2$) equal to zero (dashed line)
832     indicate force and torque directions indistinguishable from those
833     obtained using {\sc spme}. Different values of the cutoff radius are
834     indicated with different symbols (9~\AA\ = circles, 12~\AA\ = squares,
835     and 15~\AA\ = inverted triangles).}
836     \label{fig:frcTrqAng}
837     \end{figure}
838     Both the force and torque $\sigma^2$ results from the analysis of the
839     total accumulated system data are tabulated in figure
840     \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
841     sp}) method would be essentially unusable for molecular dynamics
842     unless the damping function is added. The Shifted Force ({\sc sf})
843     method, however, is generating force and torque vectors which are
844     within a few degrees of the Ewald results even with weak (or no)
845     damping.
846    
847     All of the sets (aside from the over-damped case) show the improvement
848     afforded by choosing a larger cutoff radius. Increasing the cutoff
849     from 9 to 12~\AA\ typically results in a halving of the width of the
850     distribution, with a similar improvement when going from 12 to
851     15~\AA .
852    
853     The undamped {\sc sf}, group-based cutoff, and reaction field methods
854     all do equivalently well at capturing the direction of both the force
855     and torque vectors. Using the electrostatic damping improves the
856     angular behavior significantly for the {\sc sp} and moderately for the
857     {\sc sf} methods. Over-damping is detrimental to both methods. Again
858     it is important to recognize that the force vectors cover all
859     particles in all seven systems, while torque vectors are only
860     available for neutral molecular groups. Damping is more beneficial to
861     charged bodies, and this observation is investigated further in
862 chrisfen 3001 appendix \ref{app:IndividualResults}.
863 chrisfen 2987
864     Although not discussed previously, group based cutoffs can be applied
865     to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
866     will reintroduce small discontinuities at the cutoff radius, but the
867     effects of these can be minimized by utilizing a switching function.
868     Though there are no significant benefits or drawbacks observed in
869     $\Delta E$ and the force and torque magnitudes when doing this, there
870     is a measurable improvement in the directionality of the forces and
871     torques. Table \ref{tab:groupAngle} shows the angular variances
872     obtained both without (N) and with (Y) group based cutoffs and a
873     switching function. Note that the $\alpha$ values have units of
874     \AA$^{-1}$ and the variance values have units of degrees$^2$. The
875     {\sc sp} (with an $\alpha$ of 0.2~\AA$^{-1}$ or smaller) shows much
876     narrower angular distributions when using group-based cutoffs. The
877     {\sc sf} method likewise shows improvement in the undamped and lightly
878     damped cases.
879    
880     \begin{table}
881     \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
882     THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
883     GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
884     THE REFERENCE EWALD SUMMATION}
885    
886     \footnotesize
887     \begin{center}
888     \begin{tabular}{@{} ccrrrrrrrr @{}}
889     \toprule
890     \toprule
891     & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
892     Force} \\
893     \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
894     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
895     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
896    
897     \midrule
898    
899     9~\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
900     & \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} \\
901     12~\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
902     & \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} \\
903     15~\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
904     & \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} \\
905    
906     \midrule
907    
908     9~\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
909     & \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} \\
910     12~\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
911     & \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} \\
912     15~\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
913     & \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} \\
914    
915     \bottomrule
916     \end{tabular}
917     \end{center}
918     \label{tab:groupAngle}
919     \end{table}
920    
921     One additional trend in table \ref{tab:groupAngle} is that the
922     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
923     increases, something that is more obvious with group-based cutoffs.
924     The complimentary error function inserted into the potential weakens
925     the electrostatic interaction as the value of $\alpha$ is increased.
926     However, at larger values of $\alpha$, it is possible to over-damp the
927     electrostatic interaction and to remove it completely. Kast
928     \textit{et al.} developed a method for choosing appropriate $\alpha$
929     values for these types of electrostatic summation methods by fitting
930     to $g(r)$ data, and their methods indicate optimal values of 0.34,
931     0.25, and 0.16~\AA$^{-1}$ for cutoff values of 9, 12, and 15~\AA\
932     respectively.\cite{Kast03} These appear to be reasonable choices to
933     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
934     these findings, choices this high would introduce error in the
935     molecular torques, particularly for the shorter cutoffs. Based on our
936     observations, empirical damping up to 0.2~\AA$^{-1}$ is beneficial,
937     but damping may be unnecessary when using the {\sc sf} method.
938    
939    
940     \section{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}\label{sec:ShortTimeDynamics}
941    
942     Zahn {\it et al.} investigated the structure and dynamics of water
943     using equations (\ref{eq:ZahnPot}) and
944     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
945     that a method similar (but not identical with) the damped {\sc sf}
946     method resulted in properties very similar to those obtained when
947     using the Ewald summation. The properties they studied (pair
948     distribution functions, diffusion constants, and velocity and
949     orientational correlation functions) may not be particularly sensitive
950     to the long-range and collective behavior that governs the
951     low-frequency behavior in crystalline systems. Additionally, the
952     ionic crystals are the worst case scenario for the pairwise methods
953     because they lack the reciprocal space contribution contained in the
954     Ewald summation.
955    
956     We are using two separate measures to probe the effects of these
957     alternative electrostatic methods on the dynamics in crystalline
958     materials. For short- and intermediate-time dynamics, we are
959     computing the velocity autocorrelation function, and for long-time
960     and large length-scale collective motions, we are looking at the
961     low-frequency portion of the power spectrum.
962    
963     \begin{figure}
964     \centering
965     \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
966     \caption{Velocity autocorrelation functions of NaCl crystals at
967     1000~K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \&
968     0.2~\AA$^{-1}$), and {\sc sp} ($\alpha$ = 0.2~\AA$^{-1}$). The inset is
969     a magnification of the area around the first minimum. The times to
970     first collision are nearly identical, but differences can be seen in
971     the peaks and troughs, where the undamped and weakly damped methods
972     are stiffer than the moderately damped and {\sc spme} methods.}
973     \label{fig:vCorrPlot}
974     \end{figure}
975     The short-time decay of the velocity autocorrelation function through
976     the first collision are nearly identical in figure
977     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
978     how the methods differ. The undamped {\sc sf} method has deeper
979     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
980     any of the other methods. As the damping parameter ($\alpha$) is
981     increased, these peaks are smoothed out, and the {\sc sf} method
982     approaches the {\sc spme} results. With $\alpha$ values of 0.2~\AA$^{-1}$,
983     the {\sc sf} and {\sc sp} functions are nearly identical and track the
984     {\sc spme} features quite well. This is not surprising because the {\sc sf}
985     and {\sc sp} potentials become nearly identical with increased
986     damping. However, this appears to indicate that once damping is
987     utilized, the details of the form of the potential (and forces)
988     constructed out of the damped electrostatic interaction are less
989     important.
990    
991     \section{Collective Motion: Power Spectra of NaCl Crystals}\label{sec:LongTimeDynamics}
992    
993     \begin{figure}
994     \centering
995     \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
996     \caption{Power spectra obtained from the velocity auto-correlation
997     functions of NaCl crystals at 1000~K while using {\sc spme}, {\sc sf}
998     ($\alpha$ = 0, 0.1, \& 0.2~\AA$^{-1}$), and {\sc sp} ($\alpha$ =
999     0.2~\AA$^{-1}$). The inset shows the frequency region below
1000     100~cm$^{-1}$ to highlight where the spectra differ.}
1001     \label{fig:methodPS}
1002     \end{figure}
1003 chrisfen 3023 To evaluate how the differences between the methods affect the
1004     collective long-time motion, we computed power spectra from long-time
1005     traces of the velocity autocorrelation function. The power spectra for
1006     the best-performing alternative methods are shown in
1007     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1008     a cubic switching function between 40 and 50~ps was used to reduce the
1009     ringing resulting from data truncation. This procedure had no
1010     noticeable effect on peak location or magnitude.
1011 chrisfen 2987
1012     While the high frequency regions of the power spectra for the
1013     alternative methods are quantitatively identical with Ewald spectrum,
1014     the low frequency region shows how the summation methods differ.
1015     Considering the low-frequency inset (expanded in the upper frame of
1016 chrisfen 3023 figure \ref{fig:methodPS}), at frequencies below 100~cm$^{-1}$, the
1017 chrisfen 2987 correlated motions are blue-shifted when using undamped or weakly
1018     damped {\sc sf}. When using moderate damping ($\alpha =
1019     0.2$~\AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly
1020     identical correlated motion to the Ewald method (which has a
1021     convergence parameter of 0.3119~\AA$^{-1}$). This weakening of the
1022     electrostatic interaction with increased damping explains why the
1023     long-ranged correlated motions are at lower frequencies for the
1024     moderately damped methods than for undamped or weakly damped methods.
1025    
1026 chrisfen 3023 \begin{figure}
1027     \centering
1028     \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1029     \caption{Effect of damping on the two lowest-frequency phonon modes in
1030     the NaCl crystal at 1000~K. The undamped shifted force ({\sc sf})
1031     method is off by less than 10~cm$^{-1}$, and increasing the
1032     electrostatic damping to 0.25~\AA$^{-1}$ gives quantitative agreement
1033     with the power spectrum obtained using the Ewald sum. Over-damping can
1034     result in underestimates of frequencies of the long-wavelength
1035     motions.}
1036     \label{fig:dampInc}
1037     \end{figure}
1038 chrisfen 2987 To isolate the role of the damping constant, we have computed the
1039     spectra for a single method ({\sc sf}) with a range of damping
1040     constants and compared this with the {\sc spme} spectrum.
1041     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1042     electrostatic damping red-shifts the lowest frequency phonon modes.
1043     However, even without any electrostatic damping, the {\sc sf} method
1044     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1045     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1046     would predict the lowest frequency peak near 325~cm$^{-1}$. {\it
1047     Most} of the collective behavior in the crystal is accurately captured
1048     using the {\sc sf} method. Quantitative agreement with Ewald can be
1049     obtained using moderate damping in addition to the shifting at the
1050     cutoff distance.
1051    
1052     \section{An Application: TIP5P-E Water}\label{sec:t5peApplied}
1053    
1054     The above sections focused on the energetics and dynamics of a variety
1055     of systems when utilizing the {\sc sp} and {\sc sf} pairwise
1056     techniques. A unitary correlation with results obtained using the
1057     Ewald summation should result in a successful reproduction of both the
1058     static and dynamic properties of any selected system. To test this,
1059     we decided to calculate a series of properties for the TIP5P-E water
1060     model when using the {\sc sf} technique.
1061    
1062     The TIP5P-E water model is a variant of Mahoney and Jorgensen's
1063     five-point transferable intermolecular potential (TIP5P) model for
1064     water.\cite{Mahoney00} TIP5P was developed to reproduce the density
1065     maximum anomaly present in liquid water near 4$^\circ$C. As with many
1066     previous point charge water models (such as ST2, TIP3P, TIP4P, SPC,
1067     and SPC/E), TIP5P was parametrized using a simple cutoff with no
1068     long-range electrostatic
1069     correction.\cite{Stillinger74,Jorgensen83,Berendsen81,Berendsen87}
1070     Without this correction, the pressure term on the central particle
1071     from the surroundings is missing. Because they expand to compensate
1072     for this added pressure term when this correction is included, systems
1073     composed of these particles tend to under-predict the density of water
1074     under standard conditions. When using any form of long-range
1075     electrostatic correction, it has become common practice to develop or
1076     utilize a reparametrized water model that corrects for this
1077     effect.\cite{vanderSpoel98,Fennell04,Horn04} The TIP5P-E model follows
1078     this practice and was optimized specifically for use with the Ewald
1079     summation.\cite{Rick04} In his publication, Rick preserved the
1080     geometry and point charge magnitudes in TIP5P and focused on altering
1081     the Lennard-Jones parameters to correct the density at
1082     298K.\cite{Rick04} With the density corrected, he compared common
1083     water properties for TIP5P-E using the Ewald sum with TIP5P using a
1084     9~\AA\ cutoff.
1085    
1086     In the following sections, we compared these same water properties
1087     calculated from TIP5P-E using the Ewald sum with TIP5P-E using the
1088     {\sc sf} technique. In the above evaluation of the pairwise
1089     techniques, we observed some flexibility in the choice of parameters.
1090     Because of this, the following comparisons include the {\sc sf}
1091     technique with a 12~\AA\ cutoff and an $\alpha$ = 0.0, 0.1, and
1092     0.2~\AA$^{-1}$, as well as a 9~\AA\ cutoff with an $\alpha$ =
1093     0.2~\AA$^{-1}$.
1094    
1095     \subsection{Density}\label{sec:t5peDensity}
1096    
1097     As stated previously, the property that prompted the development of
1098     TIP5P-E was the density at 1 atm. The density depends upon the
1099     internal pressure of the system in the $NPT$ ensemble, and the
1100     calculation of the pressure includes a components from both the
1101     kinetic energy and the virial. More specifically, the instantaneous
1102     molecular pressure ($p(t)$) is given by
1103     \begin{equation}
1104     p(t) = \frac{1}{\textrm{d}V}\sum_\mu
1105     \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
1106     + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
1107     \label{eq:MolecularPressure}
1108     \end{equation}
1109 chrisfen 3023 where d is the dimensionality of the system, $V$ is the volume,
1110     $\mathbf{P}_{\mu}$ is the momentum of molecule $\mu$, $\mathbf{R}_\mu$
1111     is the position of the center of mass ($M_\mu$) of molecule $\mu$, and
1112     $\mathbf{F}_{\mu i}$ is the force on atom $i$ of molecule
1113     $\mu$.\cite{Melchionna93} The virial term (the right term in the
1114     brackets of equation
1115     \ref{eq:MolecularPressure}) is directly dependent on the interatomic
1116     forces. Since the {\sc sp} method does not modify the forces (see
1117     section. \ref{sec:PairwiseDerivation}), the pressure using {\sc sp}
1118     will be identical to that obtained without an electrostatic
1119     correction. The {\sc sf} method does alter the virial component and,
1120     by way of the modified pressures, should provide densities more in
1121     line with those obtained using the Ewald summation.
1122 chrisfen 2987
1123     To compare densities, $NPT$ simulations were performed with the same
1124     temperatures as those selected by Rick in his Ewald summation
1125     simulations.\cite{Rick04} In order to improve statistics around the
1126     density maximum, 3~ns trajectories were accumulated at 0, 12.5, and
1127     25$^\circ$C, while 2~ns trajectories were obtained at all other
1128     temperatures. The average densities were calculated from the later
1129     three-fourths of each trajectory. Similar to Mahoney and Jorgensen's
1130     method for accumulating statistics, these sequences were spliced into
1131     200 segments to calculate the average density and standard deviation
1132     at each temperature.\cite{Mahoney00}
1133    
1134     \begin{figure}
1135     \includegraphics[width=\linewidth]{./figures/tip5peDensities.pdf}
1136     \caption{Density versus temperature for the TIP5P-E water model when
1137     using the Ewald summation (Ref. \cite{Rick04}) and the {\sc sf} method
1138     with various parameters. The pressure term from the image-charge shell
1139     is larger than that provided by the reciprocal-space portion of the
1140     Ewald summation, leading to slightly lower densities. This effect is
1141     more visible with the 9~\AA\ cutoff, where the image charges exert a
1142     greater force on the central particle. The error bars for the {\sc sf}
1143 chrisfen 3023 methods show the average one-sigma uncertainty of the density
1144     measurement, and this uncertainty is the same for all the {\sc sf}
1145     curves.}
1146 chrisfen 2987 \label{fig:t5peDensities}
1147     \end{figure}
1148     Figure \ref{fig:t5peDensities} shows the densities calculated for
1149     TIP5P-E using differing electrostatic corrections overlaid on the
1150     experimental values.\cite{CRC80} The densities when using the {\sc sf}
1151     technique are close to, though typically lower than, those calculated
1152     while using the Ewald summation. These slightly reduced densities
1153     indicate that the pressure component from the image charges at
1154     R$_\textrm{c}$ is larger than that exerted by the reciprocal-space
1155     portion of the Ewald summation. Bringing the image charges closer to
1156     the central particle by choosing a 9~\AA\ R$_\textrm{c}$ (rather than
1157     the preferred 12~\AA\ R$_\textrm{c}$) increases the strength of their
1158     interactions, resulting in a further reduction of the densities.
1159    
1160     Because the strength of the image charge interactions has a noticeable
1161     effect on the density, we would expect the use of electrostatic
1162     damping to also play a role in these calculations. Larger values of
1163     $\alpha$ weaken the pair-interactions; and since electrostatic damping
1164     is distance-dependent, force components from the image charges will be
1165     reduced more than those from particles close the the central
1166     charge. This effect is visible in figure \ref{fig:t5peDensities} with
1167     the damped {\sc sf} sums showing slightly higher densities; however,
1168     it is apparent that the choice of cutoff radius plays a much more
1169     important role in the resulting densities.
1170    
1171     As a final note, all of the above density calculations were performed
1172 chrisfen 3023 with systems of 512 water molecules. Rick observed a system size
1173 chrisfen 2987 dependence of the computed densities when using the Ewald summation,
1174     most likely due to his tying of the convergence parameter to the box
1175     dimensions.\cite{Rick04} For systems of 256 water molecules, the
1176     calculated densities were on average 0.002 to 0.003~g/cm$^3$ lower. A
1177     system size of 256 molecules would force the use of a shorter
1178     R$_\textrm{c}$ when using the {\sc sf} technique, and this would also
1179     lower the densities. Moving to larger systems, as long as the
1180     R$_\textrm{c}$ remains at a fixed value, we would expect the densities
1181     to remain constant.
1182    
1183     \subsection{Liquid Structure}\label{sec:t5peLiqStructure}
1184    
1185     A common function considered when developing and comparing water
1186     models is the oxygen-oxygen radial distribution function
1187     ($g_\textrm{OO}(r)$). The $g_\textrm{OO}(r)$ is the probability of
1188     finding a pair of oxygen atoms some distance ($r$) apart relative to a
1189     random distribution at the same density.\cite{Allen87} It is
1190     calculated via
1191     \begin{equation}
1192     g_\textrm{OO}(r) = \frac{V}{N^2}\left\langle\sum_i\sum_{j\ne i}
1193     \delta(\mathbf{r}-\mathbf{r}_{ij})\right\rangle,
1194     \label{eq:GOOofR}
1195     \end{equation}
1196     where the double sum is over all $i$ and $j$ pairs of $N$ oxygen
1197     atoms. The $g_\textrm{OO}(r)$ can be directly compared to X-ray or
1198     neutron scattering experiments through the oxygen-oxygen structure
1199     factor ($S_\textrm{OO}(k)$) by the following relationship:
1200     \begin{equation}
1201     S_\textrm{OO}(k) = 1 + 4\pi\rho
1202     \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
1203     \label{eq:SOOofK}
1204     \end{equation}
1205     Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
1206     of $g_\textrm{OO}(r)$.
1207    
1208     The experimentally determined $g_\textrm{OO}(r)$ for liquid water has
1209     been compared in great detail with the various common water models,
1210     and TIP5P was found to be in better agreement than other rigid,
1211     non-polarizable models.\cite{Sorenson00} This excellent agreement with
1212     experiment was maintained when Rick developed TIP5P-E.\cite{Rick04} To
1213     check whether the choice of using the Ewald summation or the {\sc sf}
1214     technique alters the liquid structure, the $g_\textrm{OO}(r)$s at 298K
1215     and 1atm were determined for the systems compared in the previous
1216     section.
1217    
1218     \begin{figure}
1219     \includegraphics[width=\linewidth]{./figures/tip5peGofR.pdf}
1220     \caption{The $g_\textrm{OO}(r)$s calculated for TIP5P-E at 298~K and
1221     1atm while using the Ewald summation (Ref. \cite{Rick04}) and the {\sc
1222     sf} technique with varying parameters. Even with the reduced densities
1223     using the {\sc sf} technique, the $g_\textrm{OO}(r)$s are essentially
1224     identical.}
1225     \label{fig:t5peGofRs}
1226     \end{figure}
1227     The $g_\textrm{OO}(r)$s calculated for TIP5P-E while using the {\sc
1228     sf} technique with a various parameters are overlaid on the
1229 chrisfen 3023 $g_\textrm{OO}(r)$ while using the Ewald summation in figure
1230     \ref{fig:t5peGofRs}. The differences in density do not appear to have
1231     any effect on the liquid structure as the $g_\textrm{OO}(r)$s are
1232     indistinguishable. These results indicate that the $g_\textrm{OO}(r)$
1233     is insensitive to the choice of electrostatic correction.
1234 chrisfen 2987
1235     \subsection{Thermodynamic Properties}\label{sec:t5peThermo}
1236    
1237     In addition to the density, there are a variety of thermodynamic
1238     quantities that can be calculated for water and compared directly to
1239     experimental values. Some of these additional quantities include the
1240     latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
1241     pressure heat capacity ($C_p$), the isothermal compressibility
1242     ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
1243     dielectric constant ($\epsilon$). All of these properties were
1244     calculated for TIP5P-E with the Ewald summation, so they provide a
1245     good set for comparisons involving the {\sc sf} technique.
1246    
1247     The $\Delta H_\textrm{vap}$ is the enthalpy change required to
1248     transform one mol of substance from the liquid phase to the gas
1249     phase.\cite{Berry00} In molecular simulations, this quantity can be
1250     determined via
1251     \begin{equation}
1252     \begin{split}
1253     \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
1254     &= E_\textrm{gas} - E_\textrm{liq.}
1255     + p(V_\textrm{gas} - V_\textrm{liq.}) \\
1256     &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
1257     \end{split}
1258     \label{eq:DeltaHVap}
1259     \end{equation}
1260     where $E$ is the total energy, $U$ is the potential energy, $p$ is the
1261     pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
1262     the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
1263     seen in the last line of equation (\ref{eq:DeltaHVap}), we can
1264     approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
1265     state. This allows us to cancel the kinetic energy terms, leaving only
1266     the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
1267     several orders of magnitude larger than that of the liquid, so we can
1268     neglect the liquid $pV$ term.
1269    
1270     The remaining thermodynamic properties can all be calculated from
1271     fluctuations of the enthalpy, volume, and system dipole
1272     moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
1273     enthalpy in constant pressure simulations via
1274     \begin{equation}
1275     \begin{split}
1276     C_p = \left(\frac{\partial H}{\partial T}\right)_{N,p}
1277     = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
1278     \end{split}
1279     \label{eq:Cp}
1280     \end{equation}
1281     where $k_B$ is Boltzmann's constant. $\kappa_T$ can be calculated via
1282     \begin{equation}
1283     \begin{split}
1284     \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
1285     = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
1286     {k_BT\langle V\rangle_{N,P,T}},
1287     \end{split}
1288     \label{eq:kappa}
1289     \end{equation}
1290     and $\alpha_p$ can be calculated via
1291     \begin{equation}
1292     \begin{split}
1293     \alpha_p = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
1294     = \frac{(\langle VH\rangle_{N,P,T}
1295     - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
1296     {k_BT^2\langle V\rangle_{N,P,T}}.
1297     \end{split}
1298     \label{eq:alpha}
1299     \end{equation}
1300     Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
1301     be calculated for systems of non-polarizable substances via
1302     \begin{equation}
1303     \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
1304     \label{eq:staticDielectric}
1305     \end{equation}
1306     where $\epsilon_0$ is the permittivity of free space and $\langle
1307     M^2\rangle$ is the fluctuation of the system dipole
1308     moment.\cite{Allen87} The numerator in the fractional term in equation
1309     (\ref{eq:staticDielectric}) is the fluctuation of the simulation-box
1310     dipole moment, identical to the quantity calculated in the
1311     finite-system Kirkwood $g$ factor ($G_k$):
1312     \begin{equation}
1313     G_k = \frac{\langle M^2\rangle}{N\mu^2},
1314     \label{eq:KirkwoodFactor}
1315     \end{equation}
1316     where $\mu$ is the dipole moment of a single molecule of the
1317     homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
1318     fluctuation term in both equation (\ref{eq:staticDielectric}) and
1319     \ref{eq:KirkwoodFactor} is calculated as follows,
1320     \begin{equation}
1321     \begin{split}
1322     \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
1323     - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
1324     &= \langle M_x^2+M_y^2+M_z^2\rangle
1325     - (\langle M_x\rangle^2 + \langle M_x\rangle^2
1326     + \langle M_x\rangle^2).
1327     \end{split}
1328     \label{eq:fluctBoxDipole}
1329     \end{equation}
1330     This fluctuation term can be accumulated during the simulation;
1331     however, it converges rather slowly, thus requiring multi-nanosecond
1332     simulation times.\cite{Horn04} In the case of tin-foil boundary
1333     conditions, the dielectric/surface term of equation (\ref{eq:EwaldSum})
1334     is equal to zero. Since the {\sc sf} method also lacks this
1335     dielectric/surface term, equation (\ref{eq:staticDielectric}) is still
1336     valid for determining static dielectric constants.
1337    
1338     All of the above properties were calculated from the same trajectories
1339     used to determine the densities in section \ref{sec:t5peDensity}
1340     except for the static dielectric constants. The $\epsilon$ values were
1341     accumulated from 2~ns $NVE$ ensemble trajectories with system densities
1342     fixed at the average values from the $NPT$ simulations at each of the
1343     temperatures. The resulting values are displayed in figure
1344     \ref{fig:t5peThermo}.
1345     \begin{figure}
1346     \centering
1347     \includegraphics[width=4.5in]{./figures/t5peThermo.pdf}
1348     \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
1349     and the {\sc sf} techniques along with the experimental values. Units
1350     for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
1351     cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
1352     and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
1353     reference \cite{Rick04}. Experimental values for $\Delta
1354     H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
1355     \cite{Kell75}. Experimental values for $C_p$ are from reference
1356     \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
1357     \cite{Malmberg56}.}
1358     \label{fig:t5peThermo}
1359     \end{figure}
1360    
1361     As observed for the density in section \ref{sec:t5peDensity}, the
1362     property trends with temperature seen when using the Ewald summation
1363 chrisfen 3023 are reproduced with the {\sc sf} technique. One noticable difference
1364     between the properties calculated using the two methods are the lower
1365     $\Delta H_\textrm{vap}$ values when using {\sc sf}. This is to be
1366     expected due to the direct weakening of the electrostatic interaction
1367     through forced neutralization. This results in an increase of the
1368     intermolecular potential producing lower values from equation
1369     (\ref{eq:DeltaHVap}). The slopes of these values with temperature are
1370     similar to that seen using the Ewald summation; however, they are both
1371     steeper than the experimental trend, indirectly resulting in the
1372     inflated $C_p$ values at all temperatures.
1373 chrisfen 2987
1374     Above the supercooled regime, $C_p$, $\kappa_T$, and $\alpha_p$
1375     values all overlap within error. As indicated for the $\Delta
1376     H_\textrm{vap}$ and $C_p$ results discussed in the previous paragraph,
1377     the deviations between experiment and simulation in this region are
1378     not the fault of the electrostatic summation methods but are due to
1379     the TIP5P class model itself. Like most rigid, non-polarizable,
1380     point-charge water models, the density decreases with temperature at a
1381     much faster rate than experiment (see figure
1382     \ref{fig:t5peDensities}). The reduced density leads to the inflated
1383     compressibility and expansivity values at higher temperatures seen
1384     here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
1385     and many-body effects are required in order for simulation to overcome
1386     these differences with experiment.\cite{Laasonen93,Donchev06}
1387    
1388     At temperatures below the freezing point for experimental water, the
1389     differences between {\sc sf} and the Ewald summation results are more
1390     apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
1391     indicate a more pronounced transition in the supercooled regime,
1392     particularly in the case of {\sc sf} without damping. This points to
1393     the onset of a more frustrated or glassy behavior for TIP5P-E at
1394 chrisfen 3023 temperatures below 250~K in the {\sc sf} simulations, indicating that
1395     disorder in the reciprical-space term of the Ewald summation might act
1396     to loosen up the local structure more than the image-charges in {\sc
1397     sf}. Because the systems are locked in different regions of
1398     phase-space, comparisons between properties at these temperatures are
1399     not exactly fair. This observation is explored in more detail in
1400     section \ref{sec:t5peDynamics}.
1401 chrisfen 2987
1402     The final thermodynamic property displayed in figure
1403     \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
1404     between the Ewald summation and the {\sc sf} technique (and experiment
1405     for that matter). It is known that the dielectric constant is
1406     dependent upon and quite sensitive to the imposed boundary
1407     conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
1408     converged $\epsilon$ values accumulated for the {\sc sf}
1409     simulations. Lack of a damping function results in dielectric
1410     constants significantly smaller than that obtained using the Ewald
1411     sum. Increasing the damping coefficient to 0.2~\AA$^{-1}$ improves the
1412     agreement considerably. It should be noted that the choice of the
1413     ``Ewald coefficient'' value also has a significant effect on the
1414     calculated value when using the Ewald summation. In the simulations of
1415     TIP5P-E with the Ewald sum, this screening parameter was tethered to
1416     the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04} In
1417     general, systems with larger screening parameters reported larger
1418     dielectric constant values, the same behavior we see here with {\sc
1419     sf}; however, the choice of cutoff radius also plays an important
1420     role. In section \ref{sec:dampingDielectric}, this connection is
1421     further explored as optimal damping coefficients for different choices
1422     of $R_\textrm{c}$ are determined for {\sc sf} for capturing the
1423     dielectric behavior.
1424    
1425     \subsection{Dynamic Properties}\label{sec:t5peDynamics}
1426    
1427     To look at the dynamic properties of TIP5P-E when using the {\sc sf}
1428     method, 200~ps $NVE$ simulations were performed for each temperature at
1429     the average density reported by the $NPT$ simulations. The
1430     self-diffusion constants ($D$) were calculated with the Einstein
1431     relation using the mean square displacement (MSD),
1432     \begin{equation}
1433 chrisfen 3023 D = \lim_{t\rightarrow\infty}
1434     \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
1435 chrisfen 2987 \label{eq:MSD}
1436     \end{equation}
1437     where $t$ is time, and $\mathbf{r}_i$ is the position of particle
1438     $i$.\cite{Allen87} Figure \ref{fig:ExampleMSD} shows an example MSD
1439     plot. As labeled in the figure, MSD plots consist of three distinct
1440     regions:
1441    
1442     \begin{enumerate}[itemsep=0pt]
1443     \item parabolic short-time ballistic motion,
1444     \item linear diffusive regime, and
1445 chrisfen 3023 \item a region with poor statistics.
1446 chrisfen 2987 \end{enumerate}
1447     The slope from the linear region (region 2) is used to calculate $D$.
1448     \begin{figure}
1449     \centering
1450     \includegraphics[width=3.5in]{./figures/ExampleMSD.pdf}
1451     \caption{Example plot of mean square displacement verses time. The
1452     left red region is the ballistic motion regime, the middle green
1453     region is the linear diffusive regime, and the right blue region is
1454     the region with poor statistics.}
1455     \label{fig:ExampleMSD}
1456     \end{figure}
1457    
1458     \begin{figure}
1459     \centering
1460     \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
1461     \caption{Body-fixed coordinate frame for a water molecule. The
1462     respective molecular principle axes point in the direction of the
1463     labeled frame axes.}
1464     \label{fig:waterFrame}
1465     \end{figure}
1466     In addition to translational diffusion, reorientational time constants
1467     were calculated for comparisons with the Ewald simulations and with
1468     experiments. These values were determined from 25~ps $NVE$ trajectories
1469     through calculation of the orientational time correlation function,
1470     \begin{equation}
1471     C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
1472     \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
1473     \label{eq:OrientCorr}
1474     \end{equation}
1475     where $P_l$ is the Legendre polynomial of order $l$ and
1476     $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
1477     principle axis $\alpha$. The principle axis frame for these water
1478     molecules is shown in figure \ref{fig:waterFrame}. As an example,
1479     $C_l^y$ is calculated from the time evolution of the unit vector
1480     connecting the two hydrogen atoms.
1481    
1482     \begin{figure}
1483     \centering
1484     \includegraphics[width=3.5in]{./figures/ExampleOrientCorr.pdf}
1485     \caption{Example plots of the orientational autocorrelation functions
1486     for the first and second Legendre polynomials. These curves show the
1487     time decay of the unit vector along the $y$ principle axis.}
1488     \label{fig:OrientCorr}
1489     \end{figure}
1490     From the orientation autocorrelation functions, we can obtain time
1491     constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
1492     some example plots of orientational autocorrelation functions for the
1493     first and second Legendre polynomials. The relatively short time
1494     portions (between 1 and 3~ps for water) of these curves can be fit to
1495     an exponential decay to obtain these constants, and they are directly
1496     comparable to water orientational relaxation times from nuclear
1497     magnetic resonance (NMR). The relaxation constant obtained from
1498     $C_2^y(t)$ is of particular interest because it describes the
1499     relaxation of the principle axis connecting the hydrogen atoms. Thus,
1500     $C_2^y(t)$ can be compared to the intermolecular portion of the
1501     dipole-dipole relaxation from a proton NMR signal and should provide
1502     the best estimate of the NMR relaxation time constant.\cite{Impey82}
1503    
1504     \begin{figure}
1505     \centering
1506     \includegraphics[width=3.5in]{./figures/t5peDynamics.pdf}
1507     \caption{Diffusion constants ({\it upper}) and reorientational time
1508     constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
1509     technique compared with experiment. Data at temperatures less that
1510     0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
1511     easier comparisons in the more relevant temperature regime.}
1512     \label{fig:t5peDynamics}
1513     \end{figure}
1514 chrisfen 3023 Results for the diffusion constants and orientational relaxation times
1515 chrisfen 2987 are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
1516     apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
1517     the Ewald sum are reproduced with the {\sc sf} technique. The enhanced
1518     diffusion at high temperatures are again the product of the lower
1519     densities in comparison with experiment and do not provide any special
1520     insight into differences between the electrostatic summation
1521     techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
1522     diffuse a little faster than with the Ewald sum; however, use of light
1523 chrisfen 3023 to moderate damping results in indistinguishable $D$ values. Though
1524     not apparent in this figure, {\sc sf} values at the lowest temperature
1525     are approximately an order of magnitude lower than with Ewald. These
1526 chrisfen 2987 values support the observation from section \ref{sec:t5peThermo} that
1527     there appeared to be a change to a more glassy-like phase with the
1528     {\sc sf} technique at these lower temperatures.
1529    
1530     The $\tau_2^y$ results in the lower frame of figure
1531     \ref{fig:t5peDynamics} show a much greater difference between the {\sc
1532     sf} results and the Ewald results. At all temperatures shown, TIP5P-E
1533     relaxes faster than experiment with the Ewald sum while tracking
1534     experiment fairly well when using the {\sc sf} technique, independent
1535     of the choice of damping constant. Their are several possible reasons
1536     for this deviation between techniques. The Ewald results were taken
1537     shorter (10ps) trajectories than the {\sc sf} results (25ps). A quick
1538     calculation from a 10~ps trajectory with {\sc sf} with an $\alpha$ of
1539 chrisfen 3023 0.2~\AA$^{-1}$ at 25$^\circ$C showed a 0.4~ps drop in $\tau_2^y$,
1540     placing the result more in line with that obtained using the Ewald
1541     sum. These results support this explanation; however, recomputing the
1542     results to meet a poorer statistical standard is
1543     counter-productive. Assuming the Ewald results are not the product of
1544     poor statistics, differences in techniques to integrate the
1545     orientational motion could also play a role. {\sc shake} is the most
1546     commonly used technique for approximating rigid-body orientational
1547     motion,\cite{Ryckaert77} where as in {\sc oopse}, we maintain and
1548     integrate the entire rotation matrix using the {\sc dlm}
1549     method.\cite{Meineke05} Since {\sc shake} is an iterative constraint
1550     technique, if the convergence tolerances are raised for increased
1551     performance, error will accumulate in the orientational
1552     motion. Finally, the Ewald results were calculated using the $NVT$
1553     ensemble, while the $NVE$ ensemble was used for {\sc sf}
1554 chrisfen 2987 calculations. The additional mode of motion due to the thermostat will
1555     alter the dynamics, resulting in differences between $NVT$ and $NVE$
1556     results. These differences are increasingly noticeable as the
1557     thermostat time constant decreases.
1558    
1559     \section{Damping of Point Multipoles}\label{sec:dampingMultipoles}
1560    
1561     As discussed above, the {\sc sp} and {\sc sf} methods operate by
1562     neutralizing the cutoff sphere with charge-charge interaction shifting
1563     and by damping the electrostatic interactions. Now we would like to
1564     consider an extension of these techniques to include point multipole
1565 chrisfen 3023 interactions. How will the shifting and damping need to be modified in
1566 chrisfen 2987 order to accommodate point multipoles?
1567    
1568 chrisfen 3023 Of the two techniques, the easiest to adapt is shifting. Shifting is
1569 chrisfen 2987 employed to neutralize the cutoff sphere; however, in a system
1570     composed purely of point multipoles, the cutoff sphere is already
1571     neutralized. This means that shifting is not necessary between point
1572     multipoles. In a mixed system of monopoles and multipoles, the
1573     undamped {\sc sf} potential needs only to shift the force terms of the
1574     monopole (and use the monopole potential of equation (\ref{eq:SFPot}))
1575     and smoothly cutoff the multipole interactions with a switching
1576     function. The switching function is required in order to conserve
1577     energy, because a discontinuity will exist at $R_\textrm{c}$ in the
1578     absence of shifting terms.
1579    
1580     If we consider damping the {\sc sf} potential (Eq. (\ref{eq:DSFPot})),
1581     then we need to incorporate the complimentary error function term into
1582     the multipole potentials. The most direct way to do this is by
1583     replacing $r^{-1}$ with erfc$(\alpha r)\cdot r^{-1}$ in the multipole
1584     expansion.\cite{Hirschfelder67} In the multipole expansion, rather
1585     than considering only the interactions between single point charges,
1586 chrisfen 3023 the electrostatic interaction is reformulated such that it describes
1587 chrisfen 2987 the interaction between charge distributions about central sites of
1588     the respective sets of charges. This procedure is what leads to the
1589     familiar charge-dipole,
1590     \begin{equation}
1591     V_\textrm{cd} = -q_i\frac{\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}{r^3_{ij}}
1592     = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}},
1593     \label{eq:chargeDipole}
1594     \end{equation}
1595     and dipole-dipole,
1596     \begin{equation}
1597     V_\textrm{dd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
1598     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}} -
1599     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}},
1600     \label{eq:dipoleDipole}
1601     \end{equation}
1602     interaction potentials.
1603    
1604     Using the charge-dipole interaction as an example, if we insert
1605     erfc$(\alpha r)\cdot r^{-1}$ in place of $r^{-1}$, a damped
1606     charge-dipole results,
1607     \begin{equation}
1608     V_\textrm{Dcd} = -q_i\mu_j\frac{\cos\theta}{r^2_{ij}} c_1(r_{ij}),
1609     \label{eq:dChargeDipole}
1610     \end{equation}
1611     where $c_1(r_{ij})$ is
1612     \begin{equation}
1613     c_1(r_{ij}) = \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
1614     + \textrm{erfc}(\alpha r_{ij}).
1615     \label{eq:c1Func}
1616     \end{equation}
1617     Thus, $c_1(r_{ij})$ is the resulting damping term that modifies the
1618     standard charge-dipole potential (Eq. (\ref{eq:chargeDipole})). Note
1619     that this damping term is dependent upon distance and not upon
1620     orientation, and that it is acting on what was originally an
1621     $r^{-3}$ function. By writing the damped form in this manner, we
1622     can collect the damping into one function and apply it to the original
1623     potential when damping is desired. This works well for potentials that
1624     have only one $r^{-n}$ term (where $n$ is an odd positive integer);
1625     but in the case of the dipole-dipole potential, there is one part
1626     dependent on $r^{-3}$ and another dependent on $r^{-5}$. In order to
1627     properly damping this potential, each of these parts is dampened with
1628     separate damping functions. We can determine the necessary damping
1629     functions by continuing with the multipole expansion; however, it
1630     quickly becomes more complex with ``two-center'' systems, like the
1631     dipole-dipole potential, and is typically approached with a spherical
1632     harmonic formalism.\cite{Hirschfelder67} A simpler method for
1633     determining these functions arises from adopting the tensor formalism
1634     for expressing the electrostatic interactions.\cite{Stone02}
1635    
1636     The tensor formalism for electrostatic interactions involves obtaining
1637     the multipole interactions from successive gradients of the monopole
1638     potential. Thus, tensors of rank one through three are
1639     \begin{equation}
1640     T = \frac{1}{4\pi\epsilon_0r_{ij}},
1641     \label{eq:tensorRank1}
1642     \end{equation}
1643     \begin{equation}
1644     T_\alpha = \frac{1}{4\pi\epsilon_0}\nabla_\alpha \frac{1}{r_{ij}},
1645     \label{eq:tensorRank2}
1646     \end{equation}
1647     \begin{equation}
1648     T_{\alpha\beta} = \frac{1}{4\pi\epsilon_0}
1649     \nabla_\alpha\nabla_\beta \frac{1}{r_{ij}},
1650     \label{eq:tensorRank3}
1651     \end{equation}
1652     where the form of the first tensor gives the monopole-monopole
1653     potential, the second gives the monopole-dipole potential, and the
1654     third gives the monopole-quadrupole and dipole-dipole
1655     potentials.\cite{Stone02} Since the force is $-\nabla V$, the forces
1656     for each potential come from the next higher tensor.
1657    
1658     To obtain the damped electrostatic forms, we replace $r^{-1}$ with
1659     erfc$(\alpha r)\cdot r^{-1}$. Equation \ref{eq:tensorRank2} generates
1660     $c_1(r_{ij})$, just like the multipole expansion, while equation
1661     \ref{eq:tensorRank3} generates $c_2(r_{ij})$, where
1662     \begin{equation}
1663     c_2(r_{ij}) = \frac{4\alpha^3r^3_{ij}e^{-\alpha^2r^2_{ij}}}{3\sqrt{\pi}}
1664     + \frac{2\alpha r_{ij}e^{-\alpha^2r^2_{ij}}}{\sqrt{\pi}}
1665     + \textrm{erfc}(\alpha r_{ij}).
1666     \end{equation}
1667     Note that $c_2(r_{ij})$ is equal to $c_1(r_{ij})$ plus an additional
1668     term. Continuing with higher rank tensors, we can obtain the damping
1669     functions for higher multipoles as well as the forces. Each subsequent
1670     damping function includes one additional term, and we can simplify the
1671     procedure for obtaining these terms by writing out the following
1672     generating function,
1673     \begin{equation}
1674     c_n(r_{ij}) = \frac{2^n(\alpha r_{ij})^{2n-1}e^{-\alpha^2r^2_{ij}}}
1675     {(2n-1)!!\sqrt{\pi}} + c_{n-1}(r_{ij}),
1676     \label{eq:dampingGeneratingFunc}
1677     \end{equation}
1678     where,
1679     \begin{equation}
1680     m!! \equiv \left\{ \begin{array}{l@{\quad\quad}l}
1681     m\cdot(m-2)\ldots 5\cdot3\cdot1 & m > 0\textrm{ odd} \\
1682     m\cdot(m-2)\ldots 6\cdot4\cdot2 & m > 0\textrm{ even} \\
1683     1 & m = -1\textrm{ or }0,
1684     \end{array}\right.
1685     \label{eq:doubleFactorial}
1686     \end{equation}
1687     and $c_0(r_{ij})$ is erfc$(\alpha r_{ij})$. This generating function
1688     is similar in form to those obtained by researchers for the
1689     application of the Ewald sum to
1690     multipoles.\cite{Smith82,Smith98,Aguado03}
1691    
1692     Returning to the dipole-dipole example, the potential consists of a
1693     portion dependent upon $r^{-5}$ and another dependent upon
1694     $r^{-3}$. In the damped dipole-dipole potential,
1695     \begin{equation}
1696     V_\textrm{Ddd} = 3\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
1697     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})}{r^5_{ij}}
1698     c_2(r_{ij}) -
1699     \frac{\boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j}{r^3_{ij}}
1700     c_1(r_{ij}),
1701     \label{eq:dampDipoleDipole}
1702     \end{equation}
1703     $c_2(r_{ij})$ and $c_1(r_{ij})$ respectively dampen these two
1704     parts. The forces for the damped dipole-dipole interaction,
1705     \begin{equation}
1706     \begin{split}
1707     F_\textrm{Ddd} = &15\frac{(\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij})
1708     (\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij})\mathbf{r}_{ij}}{r^7_{ij}}
1709     c_3(r_{ij})\\ &-
1710     3\frac{\boldsymbol{\mu}_i\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_j +
1711     \boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}\cdot\hat{\boldsymbol{\mu}}_i +
1712     \boldsymbol{\mu}_i\cdot\boldsymbol{\mu}_j\cdot\mathbf{r}_{ij}}
1713     {r^5_{ij}} c_2(r_{ij}),
1714     \end{split}
1715     \label{eq:dampDipoleDipoleForces}
1716     \end{equation}
1717     rely on higher order damping functions because we perform another
1718     gradient operation. In this manner, we can dampen higher order
1719     multipolar interactions along with the monopole interactions, allowing
1720     us to include multipoles in simulations involving damped electrostatic
1721     interactions.
1722    
1723    
1724     \section{Damping and Dielectric Constants}\label{sec:dampingDielectric}
1725    
1726     In section \ref{sec:t5peThermo}, we observed that the choice of
1727     damping coefficient plays a major role in the calculated dielectric
1728     constant. This is not too surprising given the results for damping
1729     parameter influence on the long-time correlated motions of the NaCl
1730     crystal in section \ref{sec:LongTimeDynamics}. The static dielectric
1731     constant is calculated from the long-time fluctuations of the system's
1732     accumulated dipole moment (Eq. (\ref{eq:staticDielectric})), so it is
1733     going to be quite sensitive to the choice of damping parameter. We
1734     would like to choose an optimal damping constant for any particular
1735     cutoff radius choice that would properly capture the dielectric
1736     behavior of the liquid.
1737    
1738     In order to find these optimal values, we mapped out the static
1739     dielectric constant as a function of both the damping parameter and
1740     cutoff radius for several different water models. To calculate the
1741     static dielectric constant, we performed 5~ns $NPT$ calculations on
1742     systems of 512 water molecules, using the TIP5P-E, TIP4P-Ew, SPC/E,
1743     and SSD/RF water models. TIP4P-Ew is a reparametrized version of the
1744     four-point transferable intermolecular potential (TIP4P) for water
1745     targeted for use with the Ewald summation.\cite{Horn04} SSD/RF is the
1746     reaction field modified variant of the soft sticky dipole (SSD) model
1747     for water\cite{Fennell04} This model is discussed in more detail in
1748     the next chapter. One thing to note about it, electrostatic
1749     interactions are handled via dipole-dipole interactions rather than
1750     charge-charge interactions like the other three models. Damping of the
1751     dipole-dipole interaction was handled as described in section
1752     \ref{sec:dampingMultipoles}. Each of these systems were studied with
1753     cutoff radii of 9, 10, 11, and 12~\AA\ and with damping parameter values
1754     ranging from 0 to 0.35~\AA$^{-1}$.
1755 chrisfen 3016
1756 chrisfen 2987 \begin{figure}
1757     \centering
1758     \includegraphics[width=\linewidth]{./figures/dielectricMap.pdf}
1759     \caption{The static dielectric constant for the TIP5P-E (A), TIP4P-Ew
1760     (B), SPC/E (C), and SSD/RF (D) water models as a function of cutoff
1761     radius ($R_\textrm{c}$) and damping coefficient ($\alpha$).}
1762     \label{fig:dielectricMap}
1763     \end{figure}
1764     The results of these calculations are displayed in figure
1765     \ref{fig:dielectricMap} in the form of shaded contour plots. An
1766     interesting aspect of all four contour plots is that the dielectric
1767     constant is effectively linear with respect to $\alpha$ and
1768     $R_\textrm{c}$ in the low to moderate damping regions, and the slope
1769     is 0.025~\AA$^{-1}\cdot R_\textrm{c}^{-1}$. Another point to note is
1770     that choosing $\alpha$ and $R_\textrm{c}$ identical to those used in
1771     studies with the Ewald summation results in the same calculated
1772     dielectric constant. As an example, in the paper outlining the
1773     development of TIP5P-E, the real-space cutoff and Ewald coefficient
1774     were tethered to the system size, and for a 512 molecule system are
1775     approximately 12~\AA\ and 0.25~\AA$^{-1}$ respectively.\cite{Rick04}
1776     These parameters resulted in a dielectric constant of 92$\pm$14, while
1777     with {\sc sf} these parameters give a dielectric constant of
1778     90.8$\pm$0.9. Another example comes from the TIP4P-Ew paper where
1779     $\alpha$ and $R_\textrm{c}$ were chosen to be 9.5~\AA\ and
1780     0.35~\AA$^{-1}$, and these parameters resulted in a $\epsilon_0$ equal
1781     to 63$\pm$1.\cite{Horn04} We did not perform calculations with these
1782     exact parameters, but interpolating between surrounding values gives a
1783     $\epsilon_0$ of 61$\pm$1. Seeing a dependence of the dielectric
1784     constant on $\alpha$ and $R_\textrm{c}$ with the {\sc sf} technique,
1785     it might be interesting to investigate the dielectric dependence of
1786     the real-space Ewald parameters.
1787    
1788     Although it is tempting to choose damping parameters equivalent to
1789     these Ewald examples, the results discussed in sections
1790 chrisfen 3001 \ref{sec:EnergyResults} through \ref{sec:FTDirResults} and appendix
1791     \ref{app:IndividualResults} indicate that values this high are
1792     destructive to both the energetics and dynamics. Ideally, $\alpha$
1793     should not exceed 0.3~\AA$^{-1}$ for any of the cutoff values in this
1794     range. If the optimal damping parameter is chosen to be midway between
1795     0.275 and 0.3~\AA$^{-1}$ (0.2875~\AA$^{-1}$) at the 9~\AA\ cutoff,
1796     then the linear trend with $R_\textrm{c}$ will always keep $\alpha$
1797     below 0.3~\AA$^{-1}$. This linear progression would give values of
1798     0.2875, 0.2625, 0.2375, and 0.2125~\AA$^{-1}$ for cutoff radii of 9,
1799     10, 11, and 12~\AA. Setting this to be the default behavior for the
1800     damped {\sc sf} technique will result in consistent dielectric
1801     behavior for these and other condensed molecular systems, regardless
1802     of the chosen cutoff radius. The static dielectric constants for
1803     TIP5P-E, TIP4P-Ew, SPC/E, and SSD/RF will be fixed at approximately
1804     74, 52, 58, and 89 respectively. These values are generally lower than
1805     the values reported in the literature; however, the relative
1806     dielectric behavior scales as expected when comparing the models to
1807     one another.
1808 chrisfen 2987
1809     \section{Conclusions}\label{sec:PairwiseConclusions}
1810    
1811     The above investigation of pairwise electrostatic summation techniques
1812     shows that there are viable and computationally efficient alternatives
1813     to the Ewald summation. These methods are derived from the damped and
1814     cutoff-neutralized Coulombic sum originally proposed by Wolf
1815     \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1816     method, reformulated above as equations (\ref{eq:DSFPot}) and
1817     (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1818     energetic and dynamic characteristics exhibited by simulations
1819     employing lattice summation techniques. The cumulative energy
1820     difference results showed the undamped {\sc sf} and moderately damped
1821     {\sc sp} methods produced results nearly identical to the Ewald
1822     summation. Similarly for the dynamic features, the undamped or
1823     moderately damped {\sc sf} and moderately damped {\sc sp} methods
1824     produce force and torque vector magnitude and directions very similar
1825     to the expected values. These results translate into long-time
1826     dynamic behavior equivalent to that produced in simulations using the
1827     Ewald summation. A detailed study of water simulations showed that
1828     liquid properties calculated when using {\sc sf} will also be
1829     equivalent to those obtained using the Ewald summation.
1830    
1831     As in all purely-pairwise cutoff methods, these methods are expected
1832     to scale approximately {\it linearly} with system size, and they are
1833     easily parallelizable. This should result in substantial reductions
1834     in the computational cost of performing large simulations.
1835    
1836     Aside from the computational cost benefit, these techniques have
1837     applicability in situations where the use of the Ewald sum can prove
1838     problematic. Of greatest interest is their potential use in
1839     interfacial systems, where the unmodified lattice sum techniques
1840     artificially accentuate the periodicity of the system in an
1841     undesirable manner. There have been alterations to the standard Ewald
1842     techniques, via corrections and reformulations, to compensate for
1843     these systems; but the pairwise techniques discussed here require no
1844     modifications, making them natural tools to tackle these problems.
1845     Additionally, this transferability gives them benefits over other
1846     pairwise methods, like reaction field, because estimations of physical
1847     properties (e.g. the dielectric constant) are unnecessary.
1848    
1849     If a researcher is using Monte Carlo simulations of large chemical
1850     systems containing point charges, most structural features will be
1851     accurately captured using the undamped {\sc sf} method or the {\sc sp}
1852     method with an electrostatic damping of 0.2~\AA$^{-1}$. These methods
1853     would also be appropriate for molecular dynamics simulations where the
1854     data of interest is either structural or short-time dynamical
1855     quantities. For long-time dynamics and collective motions, the safest
1856     pairwise method we have evaluated is the {\sc sf} method with an
1857     electrostatic damping between 0.2 and 0.25~\AA$^{-1}$. It is also
1858     important to note that the static dielectric constant in water
1859     simulations is highly dependent on both $\alpha$ and
1860     $R_\textrm{c}$. For consistent dielectric behavior, the damped {\sc
1861     sf} method should use an $\alpha$ of 0.2175~\AA$^{-1}$ for an
1862     $R_\textrm{c}$ of 12~\AA, and $\alpha$ should decrease by
1863     0.025~\AA$^{-1}$ for every 1~\AA\ increase in cutoff radius.
1864    
1865     We are not suggesting that there is any flaw with the Ewald sum; in
1866     fact, it is the standard by which these simple pairwise sums have been
1867     judged. However, these results do suggest that in the typical
1868     simulations performed today, the Ewald summation may no longer be
1869     required to obtain the level of accuracy most researchers have come to
1870     expect.