ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/fennellDissertation/dissertation.tex
Revision: 2918
Committed: Mon Jul 3 13:55:25 2006 UTC (18 years, 2 months ago) by chrisfen
Content type: application/x-tex
File size: 54673 byte(s)
Log Message:
Fennell Dissertation

File Contents

# User Rev Content
1 chrisfen 2918 \documentclass[12pt]{ndthesis}
2    
3     % some packages for things like equations and graphics
4     \usepackage{amsmath,bm}
5     \usepackage{amssymb}
6     \usepackage{mathrsfs}
7     \usepackage{tabularx}
8     \usepackage{graphicx}
9     \usepackage{booktabs}
10    
11     \begin{document}
12    
13     \frontmatter
14    
15     \title{APPLICATION AND DEVELOPMENT OF MOLECULAR DYNAMICS TECHNIQUES FOR THE
16     STUDY OF WATER}
17     \author{Christopher Joseph Fennell}
18     \work{Dissertation}
19     \degprior{B.Sc.}
20     \degaward{Doctor of Philosophy}
21     \advisor{J. Daniel Gezelter}
22     \department{Chemistry and Biochemistry}
23    
24     \maketitle
25    
26     \begin{abstract}
27     \end{abstract}
28    
29     \begin{dedication}
30     \end{dedication}
31    
32     \tableofcontents
33     \listoffigures
34     \listoftables
35    
36     \begin{acknowledge}
37     \end{acknowledge}
38    
39     \mainmatter
40    
41     \chapter{\label{chap:intro}INTRODUCTION AND BACKGROUND}
42    
43     \chapter{\label{chap:electrostatics}ELECTROSTATIC INTERACTION CORRECTION
44     TECHNIQUES}
45    
46     In molecular simulations, proper accumulation of the electrostatic
47     interactions is essential and is one of the most
48     computationally-demanding tasks. The common molecular mechanics force
49     fields represent atomic sites with full or partial charges protected
50     by Lennard-Jones (short range) interactions. This means that nearly
51     every pair interaction involves a calculation of charge-charge forces.
52     Coupled with relatively long-ranged $r^{-1}$ decay, the monopole
53     interactions quickly become the most expensive part of molecular
54     simulations. Historically, the electrostatic pair interaction would
55     not have decayed appreciably within the typical box lengths that could
56     be feasibly simulated. In the larger systems that are more typical of
57     modern simulations, large cutoffs should be used to incorporate
58     electrostatics correctly.
59    
60     There have been many efforts to address the proper and practical
61     handling of electrostatic interactions, and these have resulted in a
62     variety of techniques.\cite{Roux99,Sagui99,Tobias01} These are
63     typically classified as implicit methods (i.e., continuum dielectrics,
64     static dipolar fields),\cite{Born20,Grossfield00} explicit methods
65     (i.e., Ewald summations, interaction shifting or
66     truncation),\cite{Ewald21,Steinbach94} or a mixture of the two (i.e.,
67     reaction field type methods, fast multipole
68     methods).\cite{Onsager36,Rokhlin85} The explicit or mixed methods are
69     often preferred because they physically incorporate solvent molecules
70     in the system of interest, but these methods are sometimes difficult
71     to utilize because of their high computational cost.\cite{Roux99} In
72     addition to the computational cost, there have been some questions
73     regarding possible artifacts caused by the inherent periodicity of the
74     explicit Ewald summation.\cite{Tobias01}
75    
76     In this chapter, we focus on a new set of pairwise methods devised by
77     Wolf {\it et al.},\cite{Wolf99} which we further extend. These
78     methods along with a few other mixed methods (i.e. reaction field) are
79     compared with the smooth particle mesh Ewald
80     sum,\cite{Onsager36,Essmann99} which is our reference method for
81     handling long-range electrostatic interactions. The new methods for
82     handling electrostatics have the potential to scale linearly with
83     increasing system size since they involve only a simple modification
84     to the direct pairwise sum. They also lack the added periodicity of
85     the Ewald sum, so they can be used for systems which are non-periodic
86     or which have one- or two-dimensional periodicity. Below, these
87     methods are evaluated using a variety of model systems to
88     establish their usability in molecular simulations.
89    
90     \section{The Ewald Sum}
91    
92     The complete accumulation of the electrostatic interactions in a system with
93     periodic boundary conditions (PBC) requires the consideration of the
94     effect of all charges within a (cubic) simulation box as well as those
95     in the periodic replicas,
96     \begin{equation}
97     V_\textrm{elec} = \frac{1}{2} {\sum_{\mathbf{n}}}^\prime
98     \left[ \sum_{i=1}^N\sum_{j=1}^N \phi
99     \left( \mathbf{r}_{ij} + L\mathbf{n},\bm{\Omega}_i,\bm{\Omega}_j\right)
100     \right],
101     \label{eq:PBCSum}
102     \end{equation}
103     where the sum over $\mathbf{n}$ is a sum over all periodic box
104     replicas with integer coordinates $\mathbf{n} = (l,m,n)$, and the
105     prime indicates $i = j$ are neglected for $\mathbf{n} =
106     0$.\cite{deLeeuw80} Within the sum, $N$ is the number of electrostatic
107     particles, $\mathbf{r}_{ij}$ is $\mathbf{r}_j - \mathbf{r}_i$, $L$ is
108     the cell length, $\bm{\Omega}_{i,j}$ are the Euler angles for $i$ and
109     $j$, and $\phi$ is the solution to Poisson's equation
110     ($\phi(\mathbf{r}_{ij}) = q_i q_j |\mathbf{r}_{ij}|^{-1}$ for
111     charge-charge interactions). In the case of monopole electrostatics,
112     eq. (\ref{eq:PBCSum}) is conditionally convergent and is divergent for
113     non-neutral systems.
114    
115     The electrostatic summation problem was originally studied by Ewald
116     for the case of an infinite crystal.\cite{Ewald21}. The approach he
117     took was to convert this conditionally convergent sum into two
118     absolutely convergent summations: a short-ranged real-space summation
119     and a long-ranged reciprocal-space summation,
120     \begin{equation}
121     \begin{split}
122     V_\textrm{elec} = \frac{1}{2}&
123     \sum_{i=1}^N\sum_{j=1}^N \Biggr[ q_i q_j\Biggr( {\sum_{\mathbf{n}}}^\prime
124     \frac{\textrm{erfc}\left( \alpha|\mathbf{r}_{ij}+\mathbf{n}|\right)}
125     {|\mathbf{r}_{ij}+\mathbf{n}|} \\
126     &+ \frac{1}{\pi L^3}\sum_{\mathbf{k}\ne 0}\frac{4\pi^2}{|\mathbf{k}|^2}
127     \exp{\left(-\frac{\pi^2|\mathbf{k}|^2}{\alpha^2}\right)
128     \cos\left(\mathbf{k}\cdot\mathbf{r}_{ij}\right)}\Biggr)\Biggr] \\
129     &- \frac{\alpha}{\pi^{1/2}}\sum_{i=1}^N q_i^2
130     + \frac{2\pi}{(2\epsilon_\textrm{S}+1)L^3}
131     \left|\sum_{i=1}^N q_i\mathbf{r}_i\right|^2,
132     \end{split}
133     \label{eq:EwaldSum}
134     \end{equation}
135     where $\alpha$ is the damping or convergence parameter with units of
136     \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
137     $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
138     constant of the surrounding medium. The final two terms of
139     eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
140     for interacting with a surrounding dielectric.\cite{Allen87} This
141     dipolar term was neglected in early applications in molecular
142     simulations,\cite{Brush66,Woodcock71} until it was introduced by de
143     Leeuw {\it et al.} to address situations where the unit cell has a
144     dipole moment which is magnified through replication of the periodic
145     images.\cite{deLeeuw80,Smith81} If this term is taken to be zero, the
146     system is said to be using conducting (or ``tin-foil'') boundary
147     conditions, $\epsilon_{\rm S} = \infty$. Figure
148     \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
149     time. Initially, due to the small system sizes that could be
150     simulated feasibly, the entire simulation box was replicated to
151     convergence. In more modern simulations, the systems have grown large
152     enough that a real-space cutoff could potentially give convergent
153     behavior. Indeed, it has been observed that with the choice of a
154     small $\alpha$, the reciprocal-space portion of the Ewald sum can be
155     rapidly convergent and small relative to the real-space
156     portion.\cite{Karasawa89,Kolafa92}
157    
158     \begin{figure}
159     \includegraphics[width=\linewidth]{./figures/ewaldProgression.pdf}
160     \caption{The change in the need for the Ewald sum with
161     increasing computational power. A:~Initially, only small systems
162     could be studied, and the Ewald sum replicated the simulation box to
163     convergence. B:~Now, radial cutoff methods should be able to reach
164     convergence for the larger systems of charges that are common today.}
165     \label{fig:ewaldTime}
166     \end{figure}
167    
168     The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm. The
169     convergence parameter $(\alpha)$ plays an important role in balancing
170     the computational cost between the direct and reciprocal-space
171     portions of the summation. The choice of this value allows one to
172     select whether the real-space or reciprocal space portion of the
173     summation is an $\mathscr{O}(N^2)$ calculation (with the other being
174     $\mathscr{O}(N)$).\cite{Sagui99} With the appropriate choice of
175     $\alpha$ and thoughtful algorithm development, this cost can be
176     reduced to $\mathscr{O}(N^{3/2})$.\cite{Perram88} The typical route
177     taken to reduce the cost of the Ewald summation even further is to set
178     $\alpha$ such that the real-space interactions decay rapidly, allowing
179     for a short spherical cutoff. Then the reciprocal space summation is
180     optimized. These optimizations usually involve utilization of the
181     fast Fourier transform (FFT),\cite{Hockney81} leading to the
182     particle-particle particle-mesh (P3M) and particle mesh Ewald (PME)
183     methods.\cite{Shimada93,Luty94,Luty95,Darden93,Essmann95} In these
184     methods, the cost of the reciprocal-space portion of the Ewald
185     summation is reduced from $\mathscr{O}(N^2)$ down to $\mathscr{O}(N
186     \log N)$.
187    
188     These developments and optimizations have made the use of the Ewald
189     summation routine in simulations with periodic boundary
190     conditions. However, in certain systems, such as vapor-liquid
191     interfaces and membranes, the intrinsic three-dimensional periodicity
192     can prove problematic. The Ewald sum has been reformulated to handle
193     2-D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89} but these
194     methods are computationally expensive.\cite{Spohr97,Yeh99} More
195     recently, there have been several successful efforts toward reducing
196     the computational cost of 2-D lattice
197     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
198     bringing them more in line with the cost of the full 3-D summation.
199    
200     Several studies have recognized that the inherent periodicity in the
201     Ewald sum can also have an effect on three-dimensional
202     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00}
203     Solvated proteins are essentially kept at high concentration due to
204     the periodicity of the electrostatic summation method. In these
205     systems, the more compact folded states of a protein can be
206     artificially stabilized by the periodic replicas introduced by the
207     Ewald summation.\cite{Weber00} Thus, care must be taken when
208     considering the use of the Ewald summation where the assumed
209     periodicity would introduce spurious effects in the system dynamics.
210    
211    
212     \section{The Wolf and Zahn Methods}
213    
214     In a recent paper by Wolf \textit{et al.}, a procedure was outlined
215     for the accurate accumulation of electrostatic interactions in an
216     efficient pairwise fashion. This procedure lacks the inherent
217     periodicity of the Ewald summation.\cite{Wolf99} Wolf \textit{et al.}
218     observed that the electrostatic interaction is effectively
219     short-ranged in condensed phase systems and that neutralization of the
220     charge contained within the cutoff radius is crucial for potential
221     stability. They devised a pairwise summation method that ensures
222     charge neutrality and gives results similar to those obtained with the
223     Ewald summation. The resulting shifted Coulomb potential includes
224     image-charges subtracted out through placement on the cutoff sphere
225     and a distance-dependent damping function (identical to that seen in
226     the real-space portion of the Ewald sum) to aid convergence
227     \begin{equation}
228     V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}
229     - \lim_{r_{ij}\rightarrow R_\textrm{c}}
230     \left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
231     \label{eq:WolfPot}
232     \end{equation}
233     Eq. (\ref{eq:WolfPot}) is essentially the common form of a shifted
234     potential. However, neutralizing the charge contained within each
235     cutoff sphere requires the placement of a self-image charge on the
236     surface of the cutoff sphere. This additional self-term in the total
237     potential enabled Wolf {\it et al.} to obtain excellent estimates of
238     Madelung energies for many crystals.
239    
240     In order to use their charge-neutralized potential in molecular
241     dynamics simulations, Wolf \textit{et al.} suggested taking the
242     derivative of this potential prior to evaluation of the limit. This
243     procedure gives an expression for the forces,
244     \begin{equation}
245     \begin{split}
246     F_{\textrm{Wolf}}(r_{ij}) = q_i q_j \Biggr\{&
247     \Biggr[\frac{\textrm{erfc}\left(\alpha r_{ij}\right)}{r^2_{ij}}
248     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r_{ij}^2\right)}}{r_{ij}}
249     \Biggr]\\
250     &-\Biggr[
251     \frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}{R_{\textrm{c}}^2}
252     + \frac{2\alpha}{\pi^{1/2}}
253     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
254     \Biggr]\Biggr\},
255     \end{split}
256     \label{eq:WolfForces}
257     \end{equation}
258     that incorporates both image charges and damping of the electrostatic
259     interaction.
260    
261     More recently, Zahn \textit{et al.} investigated these potential and
262     force expressions for use in simulations involving water.\cite{Zahn02}
263     In their work, they pointed out that the forces and derivative of
264     the potential are not commensurate. Attempts to use both
265     eqs. (\ref{eq:WolfPot}) and (\ref{eq:WolfForces}) together will lead
266     to poor energy conservation. They correctly observed that taking the
267     limit shown in equation (\ref{eq:WolfPot}) {\it after} calculating the
268     derivatives gives forces for a different potential energy function
269     than the one shown in eq. (\ref{eq:WolfPot}).
270    
271     Zahn \textit{et al.} introduced a modified form of this summation
272     method as a way to use the technique in Molecular Dynamics
273     simulations. They proposed a new damped Coulomb potential,
274     \begin{equation}
275     \begin{split}
276     V_{\textrm{Zahn}}(r_{ij}) = q_iq_j\Biggr\{&
277     \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} \\
278     &- \left[\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
279     + \frac{2\alpha}{\pi^{1/2}}
280     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
281     \right]\left(r_{ij}-R_\mathrm{c}\right)\Biggr\},
282     \end{split}
283     \label{eq:ZahnPot}
284     \end{equation}
285     and showed that this potential does fairly well at capturing the
286     structural and dynamic properties of water compared the same
287     properties obtained using the Ewald sum.
288    
289     \section{Simple Forms for Pairwise Electrostatics}
290    
291     The potentials proposed by Wolf \textit{et al.} and Zahn \textit{et
292     al.} are constructed using two different (and separable) computational
293     tricks:
294    
295     \begin{enumerate}
296     \item shifting through the use of image charges, and
297     \item damping the electrostatic interaction.
298     \end{enumerate}
299     Wolf \textit{et al.} treated the
300     development of their summation method as a progressive application of
301     these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
302     their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
303     post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
304     both techniques. It is possible, however, to separate these
305     tricks and study their effects independently.
306    
307     Starting with the original observation that the effective range of the
308     electrostatic interaction in condensed phases is considerably less
309     than $r^{-1}$, either the cutoff sphere neutralization or the
310     distance-dependent damping technique could be used as a foundation for
311     a new pairwise summation method. Wolf \textit{et al.} made the
312     observation that charge neutralization within the cutoff sphere plays
313     a significant role in energy convergence; therefore we will begin our
314     analysis with the various shifted forms that maintain this charge
315     neutralization. We can evaluate the methods of Wolf
316     \textit{et al.} and Zahn \textit{et al.} by considering the standard
317     shifted potential,
318     \begin{equation}
319     V_\textrm{SP}(r) = \begin{cases}
320     v(r)-v_\textrm{c} &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r >
321     R_\textrm{c}
322     \end{cases},
323     \label{eq:shiftingPotForm}
324     \end{equation}
325     and shifted force,
326     \begin{equation}
327     V_\textrm{SF}(r) = \begin{cases}
328     v(r) - v_\textrm{c}
329     - \left(\frac{d v(r)}{d r}\right)_{r=R_\textrm{c}}(r-R_\textrm{c})
330     &\quad r\leqslant R_\textrm{c} \\ 0 &\quad r > R_\textrm{c}
331     \end{cases},
332     \label{eq:shiftingForm}
333     \end{equation}
334     functions where $v(r)$ is the unshifted form of the potential, and
335     $v_c$ is $v(R_\textrm{c})$. The Shifted Force ({\sc sf}) form ensures
336     that both the potential and the forces goes to zero at the cutoff
337     radius, while the Shifted Potential ({\sc sp}) form only ensures the
338     potential is smooth at the cutoff radius
339     ($R_\textrm{c}$).\cite{Allen87}
340    
341     The forces associated with the shifted potential are simply the forces
342     of the unshifted potential itself (when inside the cutoff sphere),
343     \begin{equation}
344     F_{\textrm{SP}} = -\left( \frac{d v(r)}{dr} \right),
345     \end{equation}
346     and are zero outside. Inside the cutoff sphere, the forces associated
347     with the shifted force form can be written,
348     \begin{equation}
349     F_{\textrm{SF}} = -\left( \frac{d v(r)}{dr} \right) + \left(\frac{d
350     v(r)}{dr} \right)_{r=R_\textrm{c}}.
351     \end{equation}
352    
353     If the potential, $v(r)$, is taken to be the normal Coulomb potential,
354     \begin{equation}
355     v(r) = \frac{q_i q_j}{r},
356     \label{eq:Coulomb}
357     \end{equation}
358     then the Shifted Potential ({\sc sp}) forms will give Wolf {\it et
359     al.}'s undamped prescription:
360     \begin{equation}
361     V_\textrm{SP}(r) =
362     q_iq_j\left(\frac{1}{r}-\frac{1}{R_\textrm{c}}\right) \quad
363     r\leqslant R_\textrm{c},
364     \label{eq:SPPot}
365     \end{equation}
366     with associated forces,
367     \begin{equation}
368     F_\textrm{SP}(r) = q_iq_j\left(\frac{1}{r^2}\right)
369     \quad r\leqslant R_\textrm{c}.
370     \label{eq:SPForces}
371     \end{equation}
372     These forces are identical to the forces of the standard Coulomb
373     interaction, and cutting these off at $R_c$ was addressed by Wolf
374     \textit{et al.} as undesirable. They pointed out that the effect of
375     the image charges is neglected in the forces when this form is
376     used,\cite{Wolf99} thereby eliminating any benefit from the method in
377     molecular dynamics. Additionally, there is a discontinuity in the
378     forces at the cutoff radius which results in energy drift during MD
379     simulations.
380    
381     The shifted force ({\sc sf}) form using the normal Coulomb potential
382     will give,
383     \begin{equation}
384     V_\textrm{SF}(r) = q_iq_j\left[\frac{1}{r}-\frac{1}{R_\textrm{c}}
385     + \left(\frac{1}{R_\textrm{c}^2}\right)(r-R_\textrm{c})\right]
386     \quad r\leqslant R_\textrm{c}.
387     \label{eq:SFPot}
388     \end{equation}
389     with associated forces,
390     \begin{equation}
391     F_\textrm{SF}(r) = q_iq_j\left(\frac{1}{r^2}-\frac{1}{R_\textrm{c}^2}\right)
392     \quad r\leqslant R_\textrm{c}.
393     \label{eq:SFForces}
394     \end{equation}
395     This formulation has the benefits that there are no discontinuities at
396     the cutoff radius, while the neutralizing image charges are present in
397     both the energy and force expressions. It would be simple to add the
398     self-neutralizing term back when computing the total energy of the
399     system, thereby maintaining the agreement with the Madelung energies.
400     A side effect of this treatment is the alteration in the shape of the
401     potential that comes from the derivative term. Thus, a degree of
402     clarity about agreement with the empirical potential is lost in order
403     to gain functionality in dynamics simulations.
404    
405     Wolf \textit{et al.} originally discussed the energetics of the
406     shifted Coulomb potential (Eq. \ref{eq:SPPot}) and found that it was
407     insufficient for accurate determination of the energy with reasonable
408     cutoff distances. The calculated Madelung energies fluctuated around
409     the expected value as the cutoff radius was increased, but the
410     oscillations converged toward the correct value.\cite{Wolf99} A
411     damping function was incorporated to accelerate the convergence; and
412     though alternative forms for the damping function could be
413     used,\cite{Jones56,Heyes81} the complimentary error function was
414     chosen to mirror the effective screening used in the Ewald summation.
415     Incorporating this error function damping into the simple Coulomb
416     potential,
417     \begin{equation}
418     v(r) = \frac{\mathrm{erfc}\left(\alpha r\right)}{r},
419     \label{eq:dampCoulomb}
420     \end{equation}
421     the shifted potential (eq. (\ref{eq:SPPot})) becomes
422     \begin{equation}
423     V_{\textrm{DSP}}(r) = q_iq_j\left(\frac{\textrm{erfc}\left(\alpha r\right)}{r}
424     - \frac{\textrm{erfc}\left(\alpha R_\textrm{c}\right)}{R_\textrm{c}}\right)
425     \quad r\leqslant R_\textrm{c},
426     \label{eq:DSPPot}
427     \end{equation}
428     with associated forces,
429     \begin{equation}
430     F_{\textrm{DSP}}(r) = q_iq_j
431     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
432     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right)
433     \quad r\leqslant R_\textrm{c}.
434     \label{eq:DSPForces}
435     \end{equation}
436     Again, this damped shifted potential suffers from a
437     force-discontinuity at the cutoff radius, and the image charges play
438     no role in the forces. To remedy these concerns, one may derive a
439     {\sc sf} variant by including the derivative term in
440     eq. (\ref{eq:shiftingForm}),
441     \begin{equation}
442     \begin{split}
443     V_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
444     \frac{\mathrm{erfc}\left(\alpha r\right)}{r}
445     - \frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}} \\
446     &+ \left(\frac{\mathrm{erfc}\left(\alpha R_\mathrm{c}\right)}{R_\mathrm{c}^2}
447     + \frac{2\alpha}{\pi^{1/2}}
448     \frac{\exp\left(-\alpha^2R_\mathrm{c}^2\right)}{R_\mathrm{c}}
449     \right)\left(r-R_\mathrm{c}\right)\ \Biggr{]}
450     \quad r\leqslant R_\textrm{c}.
451     \label{eq:DSFPot}
452     \end{split}
453     \end{equation}
454     The derivative of the above potential will lead to the following forces,
455     \begin{equation}
456     \begin{split}
457     F_\mathrm{DSF}(r) = q_iq_j\Biggr{[}&
458     \left(\frac{\textrm{erfc}\left(\alpha r\right)}{r^2}
459     + \frac{2\alpha}{\pi^{1/2}}\frac{\exp{\left(-\alpha^2r^2\right)}}{r}\right) \\
460     &- \left(\frac{\textrm{erfc}\left(\alpha R_{\textrm{c}}\right)}
461     {R_{\textrm{c}}^2}
462     + \frac{2\alpha}{\pi^{1/2}}
463     \frac{\exp{\left(-\alpha^2R_{\textrm{c}}^2\right)}}{R_{\textrm{c}}}
464     \right)\Biggr{]}
465     \quad r\leqslant R_\textrm{c}.
466     \label{eq:DSFForces}
467     \end{split}
468     \end{equation}
469     If the damping parameter $(\alpha)$ is set to zero, the undamped case,
470     eqs. (\ref{eq:SPPot} through \ref{eq:SFForces}) are correctly
471     recovered from eqs. (\ref{eq:DSPPot} through \ref{eq:DSFForces}).
472    
473     This new {\sc sf} potential is similar to equation \ref{eq:ZahnPot}
474     derived by Zahn \textit{et al.}; however, there are two important
475     differences.\cite{Zahn02} First, the $v_\textrm{c}$ term from
476     eq. (\ref{eq:shiftingForm}) is equal to eq. (\ref{eq:dampCoulomb})
477     with $r$ replaced by $R_\textrm{c}$. This term is {\it not} present
478     in the Zahn potential, resulting in a potential discontinuity as
479     particles cross $R_\textrm{c}$. Second, the sign of the derivative
480     portion is different. The missing $v_\textrm{c}$ term would not
481     affect molecular dynamics simulations (although the computed energy
482     would be expected to have sudden jumps as particle distances crossed
483     $R_c$). The sign problem is a potential source of errors, however.
484     In fact, it introduces a discontinuity in the forces at the cutoff,
485     because the force function is shifted in the wrong direction and
486     doesn't cross zero at $R_\textrm{c}$.
487    
488     Eqs. (\ref{eq:DSFPot}) and (\ref{eq:DSFForces}) result in an
489     electrostatic summation method in which the potential and forces are
490     continuous at the cutoff radius and which incorporates the damping
491     function proposed by Wolf \textit{et al.}\cite{Wolf99} In the rest of
492     this paper, we will evaluate exactly how good these methods ({\sc sp},
493     {\sc sf}, damping) are at reproducing the correct electrostatic
494     summation performed by the Ewald sum.
495    
496    
497     \section{Evaluating Pairwise Summation Techniques}
498    
499     In classical molecular mechanics simulations, there are two primary
500     techniques utilized to obtain information about the system of
501     interest: Monte Carlo (MC) and Molecular Dynamics (MD). Both of these
502     techniques utilize pairwise summations of interactions between
503     particle sites, but they use these summations in different ways.
504    
505     In MC, the potential energy difference between configurations dictates
506     the progression of MC sampling. Going back to the origins of this
507     method, the acceptance criterion for the canonical ensemble laid out
508     by Metropolis \textit{et al.} states that a subsequent configuration
509     is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
510     $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
511     Maintaining the correct $\Delta E$ when using an alternate method for
512     handling the long-range electrostatics will ensure proper sampling
513     from the ensemble.
514    
515     In MD, the derivative of the potential governs how the system will
516     progress in time. Consequently, the force and torque vectors on each
517     body in the system dictate how the system evolves. If the magnitude
518     and direction of these vectors are similar when using alternate
519     electrostatic summation techniques, the dynamics in the short term
520     will be indistinguishable. Because error in MD calculations is
521     cumulative, one should expect greater deviation at longer times,
522     although methods which have large differences in the force and torque
523     vectors will diverge from each other more rapidly.
524    
525     \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
526    
527     The pairwise summation techniques (outlined in section
528     \ref{sec:ESMethods}) were evaluated for use in MC simulations by
529     studying the energy differences between conformations. We took the
530     {\sc spme}-computed energy difference between two conformations to be the
531     correct behavior. An ideal performance by an alternative method would
532     reproduce these energy differences exactly (even if the absolute
533     energies calculated by the methods are different). Since none of the
534     methods provide exact energy differences, we used linear least squares
535     regressions of energy gap data to evaluate how closely the methods
536     mimicked the Ewald energy gaps. Unitary results for both the
537     correlation (slope) and correlation coefficient for these regressions
538     indicate perfect agreement between the alternative method and {\sc spme}.
539     Sample correlation plots for two alternate methods are shown in
540     Fig. \ref{fig:linearFit}.
541    
542     \begin{figure}
543     \centering
544     \includegraphics[width = \linewidth]{./figures/dualLinear.pdf}
545     \caption{Example least squares regressions of the configuration energy
546     differences for SPC/E water systems. The upper plot shows a data set
547     with a poor correlation coefficient ($R^2$), while the lower plot
548     shows a data set with a good correlation coefficient.}
549     \label{fig:linearFit}
550     \end{figure}
551    
552     Each of the seven system types (detailed in section \ref{sec:RepSims})
553     were represented using 500 independent configurations. Thus, each of
554     the alternative (non-Ewald) electrostatic summation methods was
555     evaluated using an accumulated 873,250 configurational energy
556     differences.
557    
558     Results and discussion for the individual analysis of each of the
559     system types appear in sections \ref{sec:SystemResults}, while the
560     cumulative results over all the investigated systems appear below in
561     sections \ref{sec:EnergyResults}.
562    
563     \subsection{Molecular Dynamics and the Force and Torque
564     Vectors}\label{sec:MDMethods} We evaluated the pairwise methods
565     (outlined in section \ref{sec:ESMethods}) for use in MD simulations by
566     comparing the force and torque vectors with those obtained using the
567     reference Ewald summation ({\sc spme}). Both the magnitude and the
568     direction of these vectors on each of the bodies in the system were
569     analyzed. For the magnitude of these vectors, linear least squares
570     regression analyses were performed as described previously for
571     comparing $\Delta E$ values. Instead of a single energy difference
572     between two system configurations, we compared the magnitudes of the
573     forces (and torques) on each molecule in each configuration. For a
574     system of 1000 water molecules and 40 ions, there are 1040 force
575     vectors and 1000 torque vectors. With 500 configurations, this
576     results in 520,000 force and 500,000 torque vector comparisons.
577     Additionally, data from seven different system types was aggregated
578     before the comparison was made.
579    
580     The {\it directionality} of the force and torque vectors was
581     investigated through measurement of the angle ($\theta$) formed
582     between those computed from the particular method and those from {\sc spme},
583     \begin{equation}
584     \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME}
585     \cdot \hat{F}_\textrm{M}\right),
586     \end{equation}
587     where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
588     vector computed using method M. Each of these $\theta$ values was
589     accumulated in a distribution function and weighted by the area on the
590     unit sphere. Since this distribution is a measure of angular error
591     between two different electrostatic summation methods, there is no
592     {\it a priori} reason for the profile to adhere to any specific
593     shape. Thus, gaussian fits were used to measure the width of the
594     resulting distributions. The variance ($\sigma^2$) was extracted from
595     each of these fits and was used to compare distribution widths.
596     Values of $\sigma^2$ near zero indicate vector directions
597     indistinguishable from those calculated when using the reference
598     method ({\sc spme}).
599    
600     \subsection{Short-time Dynamics}
601    
602     The effects of the alternative electrostatic summation methods on the
603     short-time dynamics of charged systems were evaluated by considering a
604     NaCl crystal at a temperature of 1000 K. A subset of the best
605     performing pairwise methods was used in this comparison. The NaCl
606     crystal was chosen to avoid possible complications from the treatment
607     of orientational motion in molecular systems. All systems were
608     started with the same initial positions and velocities. Simulations
609     were performed under the microcanonical ensemble, and velocity
610     autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
611     of the trajectories,
612     \begin{equation}
613     C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
614     \label{eq:vCorr}
615     \end{equation}
616     Velocity autocorrelation functions require detailed short time data,
617     thus velocity information was saved every 2 fs over 10 ps
618     trajectories. Because the NaCl crystal is composed of two different
619     atom types, the average of the two resulting velocity autocorrelation
620     functions was used for comparisons.
621    
622     \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
623    
624     The effects of the same subset of alternative electrostatic methods on
625     the {\it long-time} dynamics of charged systems were evaluated using
626     the same model system (NaCl crystals at 1000K). The power spectrum
627     ($I(\omega)$) was obtained via Fourier transform of the velocity
628     autocorrelation function, \begin{equation} I(\omega) =
629     \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
630     \label{eq:powerSpec}
631     \end{equation}
632     where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
633     NaCl crystal is composed of two different atom types, the average of
634     the two resulting power spectra was used for comparisons. Simulations
635     were performed under the microcanonical ensemble, and velocity
636     information was saved every 5~fs over 100~ps trajectories.
637    
638     \subsection{Representative Simulations}\label{sec:RepSims}
639     A variety of representative molecular simulations were analyzed to
640     determine the relative effectiveness of the pairwise summation
641     techniques in reproducing the energetics and dynamics exhibited by
642     {\sc spme}. We wanted to span the space of typical molecular
643     simulations (i.e. from liquids of neutral molecules to ionic
644     crystals), so the systems studied were:
645    
646     \begin{enumerate}
647     \item liquid water (SPC/E),\cite{Berendsen87}
648     \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
649     \item NaCl crystals,
650     \item NaCl melts,
651     \item a low ionic strength solution of NaCl in water (0.11 M),
652     \item a high ionic strength solution of NaCl in water (1.1 M), and
653     \item a 6\AA\ radius sphere of Argon in water.
654     \end{enumerate}
655    
656     By utilizing the pairwise techniques (outlined in section
657     \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
658     charged particles, and mixtures of the two, we hope to discern under
659     which conditions it will be possible to use one of the alternative
660     summation methodologies instead of the Ewald sum.
661    
662     For the solid and liquid water configurations, configurations were
663     taken at regular intervals from high temperature trajectories of 1000
664     SPC/E water molecules. Each configuration was equilibrated
665     independently at a lower temperature (300K for the liquid, 200K for
666     the crystal). The solid and liquid NaCl systems consisted of 500
667     $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions. Configurations for
668     these systems were selected and equilibrated in the same manner as the
669     water systems. In order to introduce measurable fluctuations in the
670     configuration energy differences, the crystalline simulations were
671     equilibrated at 1000K, near the $T_\textrm{m}$ for NaCl. The liquid
672     NaCl configurations needed to represent a fully disordered array of
673     point charges, so the high temperature of 7000K was selected for
674     equilibration. The ionic solutions were made by solvating 4 (or 40)
675     ions in a periodic box containing 1000 SPC/E water molecules. Ion and
676     water positions were then randomly swapped, and the resulting
677     configurations were again equilibrated individually. Finally, for the
678     Argon / Water ``charge void'' systems, the identities of all the SPC/E
679     waters within 6\AA\ of the center of the equilibrated water
680     configurations were converted to argon.
681    
682     These procedures guaranteed us a set of representative configurations
683     from chemically-relevant systems sampled from appropriate
684     ensembles. Force field parameters for the ions and Argon were taken
685     from the force field utilized by {\sc oopse}.\cite{Meineke05}
686    
687     \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
688     We compared the following alternative summation methods with results
689     from the reference method ({\sc spme}):
690    
691     \begin{enumerate}
692     \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
693     and 0.3\AA$^{-1}$,
694     \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
695     and 0.3\AA$^{-1}$,
696     \item reaction field with an infinite dielectric constant, and
697     \item an unmodified cutoff.
698     \end{enumerate}
699    
700     Group-based cutoffs with a fifth-order polynomial switching function
701     were utilized for the reaction field simulations. Additionally, we
702     investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
703     cutoff. The {\sc spme} electrostatics were performed using the {\sc tinker}
704     implementation of {\sc spme},\cite{Ponder87} while all other calculations
705     were performed using the {\sc oopse} molecular mechanics
706     package.\cite{Meineke05} All other portions of the energy calculation
707     (i.e. Lennard-Jones interactions) were handled in exactly the same
708     manner across all systems and configurations.
709    
710     The alternative methods were also evaluated with three different
711     cutoff radii (9, 12, and 15\AA). As noted previously, the
712     convergence parameter ($\alpha$) plays a role in the balance of the
713     real-space and reciprocal-space portions of the Ewald calculation.
714     Typical molecular mechanics packages set this to a value dependent on
715     the cutoff radius and a tolerance (typically less than $1 \times
716     10^{-4}$ kcal/mol). Smaller tolerances are typically associated with
717     increasing accuracy at the expense of computational time spent on the
718     reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
719     The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
720     in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
721     0.3119, and 0.2476\AA$^{-1}$ for cutoff radii of 9, 12, and 15\AA\
722     respectively.
723    
724    
725     \section{Discussion on the Pairwise Technique Evaluation}
726    
727     \subsection{Configuration Energy Differences}\label{sec:EnergyResults}
728     In order to evaluate the performance of the pairwise electrostatic
729     summation methods for Monte Carlo simulations, the energy differences
730     between configurations were compared to the values obtained when using
731     {\sc spme}. The results for the combined regression analysis of all
732     of the systems are shown in figure \ref{fig:delE}.
733    
734     \begin{figure}
735     \centering
736     \includegraphics[width=4.75in]{./figures/delEplot.pdf}
737     \caption{Statistical analysis of the quality of configurational energy
738     differences for a given electrostatic method compared with the
739     reference Ewald sum. Results with a value equal to 1 (dashed line)
740     indicate $\Delta E$ values indistinguishable from those obtained using
741     {\sc spme}. Different values of the cutoff radius are indicated with
742     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
743     inverted triangles).}
744     \label{fig:delE}
745     \end{figure}
746    
747     The most striking feature of this plot is how well the Shifted Force
748     ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
749     differences. For the undamped {\sc sf} method, and the
750     moderately-damped {\sc sp} methods, the results are nearly
751     indistinguishable from the Ewald results. The other common methods do
752     significantly less well.
753    
754     The unmodified cutoff method is essentially unusable. This is not
755     surprising since hard cutoffs give large energy fluctuations as atoms
756     or molecules move in and out of the cutoff
757     radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
758     some degree by using group based cutoffs with a switching
759     function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
760     significant improvement using the group-switched cutoff because the
761     salt and salt solution systems contain non-neutral groups. Section
762     \ref{sec:SystemResults} includes results for systems comprised entirely
763     of neutral groups.
764    
765     For the {\sc sp} method, inclusion of electrostatic damping improves
766     the agreement with Ewald, and using an $\alpha$ of 0.2\AA $^{-1}$
767     shows an excellent correlation and quality of fit with the {\sc spme}
768     results, particularly with a cutoff radius greater than 12
769     \AA . Use of a larger damping parameter is more helpful for the
770     shortest cutoff shown, but it has a detrimental effect on simulations
771     with larger cutoffs.
772    
773     In the {\sc sf} sets, increasing damping results in progressively {\it
774     worse} correlation with Ewald. Overall, the undamped case is the best
775     performing set, as the correlation and quality of fits are
776     consistently superior regardless of the cutoff distance. The undamped
777     case is also less computationally demanding (because no evaluation of
778     the complementary error function is required).
779    
780     The reaction field results illustrates some of that method's
781     limitations, primarily that it was developed for use in homogenous
782     systems; although it does provide results that are an improvement over
783     those from an unmodified cutoff.
784    
785     \sub
786    
787     \subsection{Magnitudes of the Force and Torque Vectors}
788    
789     Evaluation of pairwise methods for use in Molecular Dynamics
790     simulations requires consideration of effects on the forces and
791     torques. Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
792     regression results for the force and torque vector magnitudes,
793     respectively. The data in these figures was generated from an
794     accumulation of the statistics from all of the system types.
795    
796     \begin{figure}
797     \centering
798     \includegraphics[width=4.75in]{./figures/frcMagplot.pdf}
799     \caption{Statistical analysis of the quality of the force vector
800     magnitudes for a given electrostatic method compared with the
801     reference Ewald sum. Results with a value equal to 1 (dashed line)
802     indicate force magnitude values indistinguishable from those obtained
803     using {\sc spme}. Different values of the cutoff radius are indicated with
804     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
805     inverted triangles).}
806     \label{fig:frcMag}
807     \end{figure}
808    
809     Again, it is striking how well the Shifted Potential and Shifted Force
810     methods are doing at reproducing the {\sc spme} forces. The undamped and
811     weakly-damped {\sc sf} method gives the best agreement with Ewald.
812     This is perhaps expected because this method explicitly incorporates a
813     smooth transition in the forces at the cutoff radius as well as the
814     neutralizing image charges.
815    
816     Figure \ref{fig:frcMag}, for the most part, parallels the results seen
817     in the previous $\Delta E$ section. The unmodified cutoff results are
818     poor, but using group based cutoffs and a switching function provides
819     an improvement much more significant than what was seen with $\Delta
820     E$.
821    
822     With moderate damping and a large enough cutoff radius, the {\sc sp}
823     method is generating usable forces. Further increases in damping,
824     while beneficial for simulations with a cutoff radius of 9\AA\ , is
825     detrimental to simulations with larger cutoff radii.
826    
827     The reaction field results are surprisingly good, considering the poor
828     quality of the fits for the $\Delta E$ results. There is still a
829     considerable degree of scatter in the data, but the forces correlate
830     well with the Ewald forces in general. We note that the reaction
831     field calculations do not include the pure NaCl systems, so these
832     results are partly biased towards conditions in which the method
833     performs more favorably.
834    
835     \begin{figure}
836     \centering
837     \includegraphics[width=4.75in]{./figures/trqMagplot.pdf}
838     \caption{Statistical analysis of the quality of the torque vector
839     magnitudes for a given electrostatic method compared with the
840     reference Ewald sum. Results with a value equal to 1 (dashed line)
841     indicate torque magnitude values indistinguishable from those obtained
842     using {\sc spme}. Different values of the cutoff radius are indicated with
843     different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
844     inverted triangles).}
845     \label{fig:trqMag}
846     \end{figure}
847    
848     Molecular torques were only available from the systems which contained
849     rigid molecules (i.e. the systems containing water). The data in
850     fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
851    
852     Torques appear to be much more sensitive to charges at a longer
853     distance. The striking feature in comparing the new electrostatic
854     methods with {\sc spme} is how much the agreement improves with increasing
855     cutoff radius. Again, the weakly damped and undamped {\sc sf} method
856     appears to be reproducing the {\sc spme} torques most accurately.
857    
858     Water molecules are dipolar, and the reaction field method reproduces
859     the effect of the surrounding polarized medium on each of the
860     molecular bodies. Therefore it is not surprising that reaction field
861     performs best of all of the methods on molecular torques.
862    
863     \subsection{Directionality of the Force and Torque Vectors}
864    
865     It is clearly important that a new electrostatic method can reproduce
866     the magnitudes of the force and torque vectors obtained via the Ewald
867     sum. However, the {\it directionality} of these vectors will also be
868     vital in calculating dynamical quantities accurately. Force and
869     torque directionalities were investigated by measuring the angles
870     formed between these vectors and the same vectors calculated using
871     {\sc spme}. The results (Fig. \ref{fig:frcTrqAng}) are compared through the
872     variance ($\sigma^2$) of the Gaussian fits of the angle error
873     distributions of the combined set over all system types.
874    
875     \begin{figure}
876     \centering
877     \includegraphics[width=4.75in]{./figures/frcTrqAngplot.pdf}
878     \caption{Statistical analysis of the width of the angular distribution
879     that the force and torque vectors from a given electrostatic method
880     make with their counterparts obtained using the reference Ewald sum.
881     Results with a variance ($\sigma^2$) equal to zero (dashed line)
882     indicate force and torque directions indistinguishable from those
883     obtained using {\sc spme}. Different values of the cutoff radius are
884     indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
885     and 15\AA\ = inverted triangles).}
886     \label{fig:frcTrqAng}
887     \end{figure}
888    
889     Both the force and torque $\sigma^2$ results from the analysis of the
890     total accumulated system data are tabulated in figure
891     \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
892     sp}) method would be essentially unusable for molecular dynamics
893     unless the damping function is added. The Shifted Force ({\sc sf})
894     method, however, is generating force and torque vectors which are
895     within a few degrees of the Ewald results even with weak (or no)
896     damping.
897    
898     All of the sets (aside from the over-damped case) show the improvement
899     afforded by choosing a larger cutoff radius. Increasing the cutoff
900     from 9 to 12\AA\ typically results in a halving of the width of the
901     distribution, with a similar improvement when going from 12 to 15
902     \AA .
903    
904     The undamped {\sc sf}, group-based cutoff, and reaction field methods
905     all do equivalently well at capturing the direction of both the force
906     and torque vectors. Using the electrostatic damping improves the
907     angular behavior significantly for the {\sc sp} and moderately for the
908     {\sc sf} methods. Overdamping is detrimental to both methods. Again
909     it is important to recognize that the force vectors cover all
910     particles in all seven systems, while torque vectors are only
911     available for neutral molecular groups. Damping is more beneficial to
912     charged bodies, and this observation is investigated further in
913     section \ref{SystemResults}.
914    
915     Although not discussed previously, group based cutoffs can be applied
916     to both the {\sc sp} and {\sc sf} methods. The group-based cutoffs
917     will reintroduce small discontinuities at the cutoff radius, but the
918     effects of these can be minimized by utilizing a switching function.
919     Though there are no significant benefits or drawbacks observed in
920     $\Delta E$ and the force and torque magnitudes when doing this, there
921     is a measurable improvement in the directionality of the forces and
922     torques. Table \ref{tab:groupAngle} shows the angular variances
923     obtained both without (N) and with (Y) group based cutoffs and a
924     switching function. Note that the $\alpha$ values have units of
925     \AA$^{-1}$ and the variance values have units of degrees$^2$. The
926     {\sc sp} (with an $\alpha$ of 0.2\AA$^{-1}$ or smaller) shows much
927     narrower angular distributions when using group-based cutoffs. The
928     {\sc sf} method likewise shows improvement in the undamped and lightly
929     damped cases.
930    
931     \begin{table}
932     \caption{STATISTICAL ANALYSIS OF THE ANGULAR DISTRIBUTIONS ($\sigma^2$)
933     THAT THE FORCE ({\it upper}) AND TORQUE ({\it lower}) VECTORS FROM A
934     GIVEN ELECTROSTATIC METHOD MAKE WITH THEIR COUNTERPARTS OBTAINED USING
935     THE REFERENCE EWALD SUMMATION}
936    
937     \footnotesize
938     \begin{center}
939     \begin{tabular}{@{} ccrrrrrrrr @{}} \\
940     \toprule
941     \toprule
942    
943     & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
944     Force} \\
945     \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
946     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$ &
947     $\alpha = 0$ & $\alpha = 0.1$ & $\alpha = 0.2$ & $\alpha = 0.3$\\
948    
949     \midrule
950    
951     9\AA & N & 29.545 & 12.003 & 5.489 & 0.610 & 2.323 & 2.321 & 0.429 & 0.603 \\
952     & \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} \\
953     12\AA & N & 19.381 & 3.097 & 0.190 & 0.608 & 0.920 & 0.736 & 0.133 & 0.612 \\
954     & \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} \\
955     15\AA & N & 12.700 & 1.196 & 0.123 & 0.601 & 0.339 & 0.160 & 0.123 & 0.601 \\
956     & \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} \\
957    
958     \midrule
959    
960     9\AA & N & 262.716 & 116.585 & 5.234 & 5.103 & 2.392 & 2.350 & 1.770 & 5.122 \\
961     & \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} \\
962     12\AA & N & 129.576 & 25.560 & 1.369 & 5.080 & 0.913 & 0.790 & 1.362 & 5.124 \\
963     & \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} \\
964     15\AA & N & 87.275 & 4.473 & 1.271 & 5.000 & 0.372 & 0.312 & 1.271 & 5.000 \\
965     & \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} \\
966    
967     \bottomrule
968     \end{tabular}
969     \end{center}
970     \label{tab:groupAngle}
971     \end{table}
972    
973     One additional trend in table \ref{tab:groupAngle} is that the
974     $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
975     increases, something that is more obvious with group-based cutoffs.
976     The complimentary error function inserted into the potential weakens
977     the electrostatic interaction as the value of $\alpha$ is increased.
978     However, at larger values of $\alpha$, it is possible to overdamp the
979     electrostatic interaction and to remove it completely. Kast
980     \textit{et al.} developed a method for choosing appropriate $\alpha$
981     values for these types of electrostatic summation methods by fitting
982     to $g(r)$ data, and their methods indicate optimal values of 0.34,
983     0.25, and 0.16\AA$^{-1}$ for cutoff values of 9, 12, and 15\AA\
984     respectively.\cite{Kast03} These appear to be reasonable choices to
985     obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
986     these findings, choices this high would introduce error in the
987     molecular torques, particularly for the shorter cutoffs. Based on our
988     observations, empirical damping up to 0.2\AA$^{-1}$ is beneficial,
989     but damping may be unnecessary when using the {\sc sf} method.
990    
991     \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
992    
993     Zahn {\it et al.} investigated the structure and dynamics of water
994     using eqs. (\ref{eq:ZahnPot}) and
995     (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
996     that a method similar (but not identical with) the damped {\sc sf}
997     method resulted in properties very similar to those obtained when
998     using the Ewald summation. The properties they studied (pair
999     distribution functions, diffusion constants, and velocity and
1000     orientational correlation functions) may not be particularly sensitive
1001     to the long-range and collective behavior that governs the
1002     low-frequency behavior in crystalline systems. Additionally, the
1003     ionic crystals are the worst case scenario for the pairwise methods
1004     because they lack the reciprocal space contribution contained in the
1005     Ewald summation.
1006    
1007     We are using two separate measures to probe the effects of these
1008     alternative electrostatic methods on the dynamics in crystalline
1009     materials. For short- and intermediate-time dynamics, we are
1010     computing the velocity autocorrelation function, and for long-time
1011     and large length-scale collective motions, we are looking at the
1012     low-frequency portion of the power spectrum.
1013    
1014     \begin{figure}
1015     \centering
1016     \includegraphics[width = \linewidth]{./figures/vCorrPlot.pdf}
1017     \caption{Velocity autocorrelation functions of NaCl crystals at
1018     1000K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1019     sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
1020     the first minimum. The times to first collision are nearly identical,
1021     but differences can be seen in the peaks and troughs, where the
1022     undamped and weakly damped methods are stiffer than the moderately
1023     damped and {\sc spme} methods.}
1024     \label{fig:vCorrPlot}
1025     \end{figure}
1026    
1027     The short-time decay of the velocity autocorrelation function through
1028     the first collision are nearly identical in figure
1029     \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1030     how the methods differ. The undamped {\sc sf} method has deeper
1031     troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1032     any of the other methods. As the damping parameter ($\alpha$) is
1033     increased, these peaks are smoothed out, and the {\sc sf} method
1034     approaches the {\sc spme} results. With $\alpha$ values of 0.2\AA$^{-1}$,
1035     the {\sc sf} and {\sc sp} functions are nearly identical and track the
1036     {\sc spme} features quite well. This is not surprising because the {\sc sf}
1037     and {\sc sp} potentials become nearly identical with increased
1038     damping. However, this appears to indicate that once damping is
1039     utilized, the details of the form of the potential (and forces)
1040     constructed out of the damped electrostatic interaction are less
1041     important.
1042    
1043     \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1044    
1045     To evaluate how the differences between the methods affect the
1046     collective long-time motion, we computed power spectra from long-time
1047     traces of the velocity autocorrelation function. The power spectra for
1048     the best-performing alternative methods are shown in
1049     fig. \ref{fig:methodPS}. Apodization of the correlation functions via
1050     a cubic switching function between 40 and 50 ps was used to reduce the
1051     ringing resulting from data truncation. This procedure had no
1052     noticeable effect on peak location or magnitude.
1053    
1054     \begin{figure}
1055     \centering
1056     \includegraphics[width = \linewidth]{./figures/spectraSquare.pdf}
1057     \caption{Power spectra obtained from the velocity auto-correlation
1058     functions of NaCl crystals at 1000K while using {\sc spme}, {\sc sf}
1059     ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset
1060     shows the frequency region below 100 cm$^{-1}$ to highlight where the
1061     spectra differ.}
1062     \label{fig:methodPS}
1063     \end{figure}
1064    
1065     While the high frequency regions of the power spectra for the
1066     alternative methods are quantitatively identical with Ewald spectrum,
1067     the low frequency region shows how the summation methods differ.
1068     Considering the low-frequency inset (expanded in the upper frame of
1069     figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1070     correlated motions are blue-shifted when using undamped or weakly
1071     damped {\sc sf}. When using moderate damping ($\alpha = 0.2$
1072     \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1073     correlated motion to the Ewald method (which has a convergence
1074     parameter of 0.3119\AA$^{-1}$). This weakening of the electrostatic
1075     interaction with increased damping explains why the long-ranged
1076     correlated motions are at lower frequencies for the moderately damped
1077     methods than for undamped or weakly damped methods.
1078    
1079     To isolate the role of the damping constant, we have computed the
1080     spectra for a single method ({\sc sf}) with a range of damping
1081     constants and compared this with the {\sc spme} spectrum.
1082     Fig. \ref{fig:dampInc} shows more clearly that increasing the
1083     electrostatic damping red-shifts the lowest frequency phonon modes.
1084     However, even without any electrostatic damping, the {\sc sf} method
1085     has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1086     Without the {\sc sf} modifications, an undamped (pure cutoff) method
1087     would predict the lowest frequency peak near 325 cm$^{-1}$. {\it
1088     Most} of the collective behavior in the crystal is accurately captured
1089     using the {\sc sf} method. Quantitative agreement with Ewald can be
1090     obtained using moderate damping in addition to the shifting at the
1091     cutoff distance.
1092    
1093     \begin{figure}
1094     \centering
1095     \includegraphics[width = \linewidth]{./figures/increasedDamping.pdf}
1096     \caption{Effect of damping on the two lowest-frequency phonon modes in
1097     the NaCl crystal at 1000K. The undamped shifted force ({\sc sf})
1098     method is off by less than 10 cm$^{-1}$, and increasing the
1099     electrostatic damping to 0.25\AA$^{-1}$ gives quantitative agreement
1100     with the power spectrum obtained using the Ewald sum. Overdamping can
1101     result in underestimates of frequencies of the long-wavelength
1102     motions.}
1103     \label{fig:dampInc}
1104     \end{figure}
1105    
1106    
1107     \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
1108    
1109     \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}
1110    
1111     \chapter{\label{chap:shapes}SPHERICAL HARMONIC APPROXIMATIONS FOR MOLECULAR
1112     SIMULATIONS}
1113    
1114     \chapter{\label{chap:conclusion}CONCLUSION}
1115    
1116     \backmatter
1117    
1118     \bibliographystyle{ndthesis}
1119     \bibliography{dissertation}
1120    
1121     \end{document}
1122    
1123    
1124     \endinput