ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/electrostaticMethodsPaper/electrostaticMethods.tex
(Generate patch)

Comparing trunk/electrostaticMethodsPaper/electrostaticMethods.tex (file contents):
Revision 2643 by gezelter, Mon Mar 20 17:32:33 2006 UTC vs.
Revision 2741 by chrisfen, Wed Apr 26 15:22:09 2006 UTC

# Line 25 | Line 25
25  
26   \begin{document}
27  
28 < \title{Is the Ewald Summation necessary? Pairwise alternatives to the accepted standard for long-range electrostatics}
28 > \title{Is the Ewald summation still necessary? \\
29 > Pairwise alternatives to the accepted standard for \\
30 > long-range electrostatics in molecular simulations}
31  
32   \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
33   gezelter@nd.edu} \\
# Line 38 | Line 40 | Notre Dame, Indiana 46556}
40   \maketitle
41   \doublespacing
42  
41 \nobibliography{}
43   \begin{abstract}
44 < A new method for accumulating electrostatic interactions was derived
45 < from the previous efforts described in \bibentry{Wolf99} and
46 < \bibentry{Zahn02} as a possible replacement for lattice sum methods in
47 < molecular simulations.  Comparisons were performed with this and other
48 < pairwise electrostatic summation techniques against the smooth
49 < particle mesh Ewald (SPME) summation to see how well they reproduce
50 < the energetics and dynamics of a variety of simulation types.  The
51 < newly derived Shifted-Force technique shows a remarkable ability to
52 < reproduce the behavior exhibited in simulations using SPME with an
53 < $\mathscr{O}(N)$ computational cost, equivalent to merely the
54 < real-space portion of the lattice summation.
55 <
44 > We investigate pairwise electrostatic interaction methods and show
45 > that there are viable and computationally efficient $(\mathscr{O}(N))$
46 > alternatives to the Ewald summation for typical modern molecular
47 > simulations.  These methods are extended from the damped and
48 > cutoff-neutralized Coulombic sum originally proposed by
49 > [D. Wolf, P. Keblinski, S.~R. Phillpot, and J. Eggebrecht, {\it J. Chem. Phys.} {\bf 110}, 8255 (1999)] One of these, the damped shifted force method, shows
50 > a remarkable ability to reproduce the energetic and dynamic
51 > characteristics exhibited by simulations employing lattice summation
52 > techniques.  Comparisons were performed with this and other pairwise
53 > methods against the smooth particle mesh Ewald ({\sc spme}) summation
54 > to see how well they reproduce the energetics and dynamics of a
55 > variety of simulation types.
56   \end{abstract}
57  
58   \newpage
# Line 94 | Line 95 | explicit Ewald summation.\cite{Tobias01}
95   regarding possible artifacts caused by the inherent periodicity of the
96   explicit Ewald summation.\cite{Tobias01}
97  
98 < In this paper, we focus on a new set of shifted methods devised by
98 > In this paper, we focus on a new set of pairwise methods devised by
99   Wolf {\it et al.},\cite{Wolf99} which we further extend.  These
100   methods along with a few other mixed methods (i.e. reaction field) are
101   compared with the smooth particle mesh Ewald
# Line 105 | Line 106 | or which have one- or two-dimensional periodicity.  Be
106   to the direct pairwise sum.  They also lack the added periodicity of
107   the Ewald sum, so they can be used for systems which are non-periodic
108   or which have one- or two-dimensional periodicity.  Below, these
109 < methods are evaluated using a variety of model systems to establish
110 < their usability in molecular simulations.
109 > methods are evaluated using a variety of model systems to
110 > establish their usability in molecular simulations.
111  
112   \subsection{The Ewald Sum}
113 < The complete accumulation electrostatic interactions in a system with
113 > The complete accumulation of the electrostatic interactions in a system with
114   periodic boundary conditions (PBC) requires the consideration of the
115   effect of all charges within a (cubic) simulation box as well as those
116   in the periodic replicas,
# Line 140 | Line 141 | V_\textrm{elec} = \frac{1}{2}& \sum_{i=1}^N\sum_{j=1}^
141   \end{split}
142   \label{eq:EwaldSum}
143   \end{equation}
144 < where $\alpha$ is a damping parameter, or separation constant, with
145 < units of \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are
146 < equal to $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the
147 < dielectric constant of the surrounding medium. The final two terms of
144 > where $\alpha$ is the damping or convergence parameter with units of
145 > \AA$^{-1}$, $\mathbf{k}$ are the reciprocal vectors and are equal to
146 > $2\pi\mathbf{n}/L^2$, and $\epsilon_\textrm{S}$ is the dielectric
147 > constant of the surrounding medium. The final two terms of
148   eq. (\ref{eq:EwaldSum}) are a particle-self term and a dipolar term
149   for interacting with a surrounding dielectric.\cite{Allen87} This
150   dipolar term was neglected in early applications in molecular
# Line 154 | Line 155 | conditions, $\epsilon_{\rm S} = \infty$. Figure
155   system is said to be using conducting (or ``tin-foil'') boundary
156   conditions, $\epsilon_{\rm S} = \infty$. Figure
157   \ref{fig:ewaldTime} shows how the Ewald sum has been applied over
158 < time.  Initially, due to the small sizes of the systems that could be
159 < feasibly simulated, the entire simulation box was replicated to
160 < convergence.  In more modern simulations, the simulation boxes have
161 < grown large enough that a real-space cutoff could potentially give
162 < convergent behavior.  Indeed, it has often been observed that the
163 < reciprocal-space portion of the Ewald sum can be vanishingly
164 < small compared to the real-space portion.\cite{XXX}
158 > time.  Initially, due to the small system sizes that could be
159 > simulated feasibly, the entire simulation box was replicated to
160 > convergence.  In more modern simulations, the systems have grown large
161 > enough that a real-space cutoff could potentially give convergent
162 > behavior.  Indeed, it has been observed that with the choice of a
163 > small $\alpha$, the reciprocal-space portion of the Ewald sum can be
164 > rapidly convergent and small relative to the real-space
165 > portion.\cite{Karasawa89,Kolafa92}
166  
167   \begin{figure}
168   \centering
169   \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
170 < \caption{How the application of the Ewald summation has changed with
171 < the increase in computer power.  Initially, only small numbers of
172 < particles could be studied, and the Ewald sum acted to replicate the
173 < unit cell charge distribution out to convergence.  Now, much larger
174 < systems of charges are investigated with fixed distance cutoffs.  The
173 < calculated structure factor is used to sum out to great distance, and
174 < a surrounding dielectric term is included.}
170 > \caption{The change in the need for the Ewald sum with
171 > increasing computational power.  A:~Initially, only small systems
172 > could be studied, and the Ewald sum replicated the simulation box to
173 > convergence.  B:~Now, radial cutoff methods should be able to reach
174 > convergence for the larger systems of charges that are common today.}
175   \label{fig:ewaldTime}
176   \end{figure}
177  
178   The original Ewald summation is an $\mathscr{O}(N^2)$ algorithm.  The
179 < separation constant $(\alpha)$ plays an important role in balancing
179 > convergence parameter $(\alpha)$ plays an important role in balancing
180   the computational cost between the direct and reciprocal-space
181   portions of the summation.  The choice of this value allows one to
182   select whether the real-space or reciprocal space portion of the
# Line 201 | Line 201 | can prove problematic.  The Ewald sum has been reformu
201   interfaces and membranes, the intrinsic three-dimensional periodicity
202   can prove problematic.  The Ewald sum has been reformulated to handle
203   2D systems,\cite{Parry75,Parry76,Heyes77,deLeeuw79,Rhee89}, but the
204 < new methods are computationally expensive.\cite{Spohr97,Yeh99}
205 < Inclusion of a correction term in the Ewald summation is a possible
206 < direction for handling 2D systems while still enabling the use of the
207 < modern optimizations.\cite{Yeh99}
204 > new methods are computationally expensive.\cite{Spohr97,Yeh99} More
205 > recently, there have been several successful efforts toward reducing
206 > the computational cost of 2D lattice summations, often enabling the
207 > use of the mentioned
208 > optimizations.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
209  
210   Several studies have recognized that the inherent periodicity in the
211   Ewald sum can also have an effect on three-dimensional
# Line 227 | Line 228 | charge neutrality and gives results similar to those o
228   charge contained within the cutoff radius is crucial for potential
229   stability. They devised a pairwise summation method that ensures
230   charge neutrality and gives results similar to those obtained with the
231 < Ewald summation.  The resulting shifted Coulomb potential
232 < (Eq. \ref{eq:WolfPot}) includes image-charges subtracted out through
233 < placement on the cutoff sphere and a distance-dependent damping
234 < function (identical to that seen in the real-space portion of the
234 < Ewald sum) to aid convergence
231 > Ewald summation.  The resulting shifted Coulomb potential includes
232 > image-charges subtracted out through placement on the cutoff sphere
233 > and a distance-dependent damping function (identical to that seen in
234 > the real-space portion of the Ewald sum) to aid convergence
235   \begin{equation}
236   V_{\textrm{Wolf}}(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha r_{ij})}{r_{ij}}\right\}.
237   \label{eq:WolfPot}
# Line 492 | Line 492 | particle sites, but they use these summations in diffe
492   techniques utilize pairwise summations of interactions between
493   particle sites, but they use these summations in different ways.
494  
495 < In MC, the potential energy difference between two subsequent
496 < configurations dictates the progression of MC sampling.  Going back to
497 < the origins of this method, the acceptance criterion for the canonical
498 < ensemble laid out by Metropolis \textit{et al.} states that a
499 < subsequent configuration is accepted if $\Delta E < 0$ or if $\xi <
500 < \exp(-\Delta E/kT)$, where $\xi$ is a random number between 0 and
501 < 1.\cite{Metropolis53} Maintaining the correct $\Delta E$ when using an
502 < alternate method for handling the long-range electrostatics will
503 < ensure proper sampling from the ensemble.
495 > In MC, the potential energy difference between configurations dictates
496 > the progression of MC sampling.  Going back to the origins of this
497 > method, the acceptance criterion for the canonical ensemble laid out
498 > by Metropolis \textit{et al.} states that a subsequent configuration
499 > is accepted if $\Delta E < 0$ or if $\xi < \exp(-\Delta E/kT)$, where
500 > $\xi$ is a random number between 0 and 1.\cite{Metropolis53}
501 > Maintaining the correct $\Delta E$ when using an alternate method for
502 > handling the long-range electrostatics will ensure proper sampling
503 > from the ensemble.
504  
505   In MD, the derivative of the potential governs how the system will
506   progress in time.  Consequently, the force and torque vectors on each
# Line 513 | Line 513 | vectors will diverge from each other more rapidly.
513   vectors will diverge from each other more rapidly.
514  
515   \subsection{Monte Carlo and the Energy Gap}\label{sec:MCMethods}
516 +
517   The pairwise summation techniques (outlined in section
518   \ref{sec:ESMethods}) were evaluated for use in MC simulations by
519   studying the energy differences between conformations.  We took the
520 < SPME-computed energy difference between two conformations to be the
520 > {\sc spme}-computed energy difference between two conformations to be the
521   correct behavior. An ideal performance by an alternative method would
522 < reproduce these energy differences exactly.  Since none of the methods
523 < provide exact energy differences, we used linear least squares
524 < regressions of the $\Delta E$ values between configurations using SPME
525 < against $\Delta E$ values using tested methods provides a quantitative
526 < comparison of this agreement.  Unitary results for both the
527 < correlation and correlation coefficient for these regressions indicate
528 < equivalent energetic results between the method under consideration
529 < and electrostatics handled using SPME.  Sample correlation plots for
530 < two alternate methods are shown in Fig. \ref{fig:linearFit}.
522 > reproduce these energy differences exactly (even if the absolute
523 > energies calculated by the methods are different).  Since none of the
524 > methods provide exact energy differences, we used linear least squares
525 > regressions of energy gap data to evaluate how closely the methods
526 > mimicked the Ewald energy gaps.  Unitary results for both the
527 > correlation (slope) and correlation coefficient for these regressions
528 > indicate perfect agreement between the alternative method and {\sc spme}.
529 > Sample correlation plots for two alternate methods are shown in
530 > Fig. \ref{fig:linearFit}.
531  
532   \begin{figure}
533   \centering
534   \includegraphics[width = \linewidth]{./dualLinear.pdf}
535 < \caption{Example least squares regressions of the configuration energy differences for SPC/E water systems. The upper plot shows a data set with a poor correlation coefficient ($R^2$), while the lower plot shows a data set with a good correlation coefficient.}
536 < \label{fig:linearFit}
535 > \caption{Example least squares regressions of the configuration energy
536 > differences for SPC/E water systems. The upper plot shows a data set
537 > with a poor correlation coefficient ($R^2$), while the lower plot
538 > shows a data set with a good correlation coefficient.}
539 > \label{fig:linearFit}
540   \end{figure}
541  
542 < Each system type (detailed in section \ref{sec:RepSims}) was
543 < represented using 500 independent configurations.  Additionally, we
544 < used seven different system types, so each of the alternate
545 < (non-Ewald) electrostatic summation methods was evaluated using
546 < 873,250 configurational energy differences.
542 > Each of the seven system types (detailed in section \ref{sec:RepSims})
543 > were represented using 500 independent configurations.  Thus, each of
544 > the alternative (non-Ewald) electrostatic summation methods was
545 > evaluated using an accumulated 873,250 configurational energy
546 > differences.
547  
548   Results and discussion for the individual analysis of each of the
549   system types appear in the supporting information, while the
# Line 550 | Line 554 | comparing the force and torque vectors with those obta
554   We evaluated the pairwise methods (outlined in section
555   \ref{sec:ESMethods}) for use in MD simulations by
556   comparing the force and torque vectors with those obtained using the
557 < reference Ewald summation (SPME).  Both the magnitude and the
557 > reference Ewald summation ({\sc spme}).  Both the magnitude and the
558   direction of these vectors on each of the bodies in the system were
559   analyzed.  For the magnitude of these vectors, linear least squares
560   regression analyses were performed as described previously for
# Line 565 | Line 569 | investigated through measurement of the angle ($\theta
569  
570   The {\it directionality} of the force and torque vectors was
571   investigated through measurement of the angle ($\theta$) formed
572 < between those computed from the particular method and those from SPME,
572 > between those computed from the particular method and those from {\sc spme},
573   \begin{equation}
574 < \theta_f = \cos^{-1} \left(\hat{f}_\textrm{SPME} \cdot \hat{f}_\textrm{Method}\right),
574 > \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
575   \end{equation}
576 < where $\hat{f}_\textrm{M}$ is the unit vector pointing along the
577 < force vector computed using method $M$.  
578 <
579 < Each of these $\theta$ values was accumulated in a distribution
576 < function, weighted by the area on the unit sphere.  Non-linear
577 < Gaussian fits were used to measure the width of the resulting
578 < distributions.
579 <
580 < \begin{figure}
581 < \centering
582 < \includegraphics[width = \linewidth]{./gaussFit.pdf}
583 < \caption{Sample fit of the angular distribution of the force vectors over all of the studied systems.  Gaussian fits were used to obtain values for the variance in force and torque vectors used in the following figure.}
584 < \label{fig:gaussian}
585 < \end{figure}
586 <
587 < Figure \ref{fig:gaussian} shows an example distribution with applied
588 < non-linear fits.  The solid line is a Gaussian profile, while the
589 < dotted line is a Voigt profile, a convolution of a Gaussian and a
590 < Lorentzian.  Since this distribution is a measure of angular error
576 > where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
577 > vector computed using method M.  Each of these $\theta$ values was
578 > accumulated in a distribution function and weighted by the area on the
579 > unit sphere.  Since this distribution is a measure of angular error
580   between two different electrostatic summation methods, there is no
581 < {\it a priori} reason for the profile to adhere to any specific shape.
582 < Gaussian fits was used to compare all the tested methods.  The
583 < variance ($\sigma^2$) was extracted from each of these fits and was
584 < used to compare distribution widths.  Values of $\sigma^2$ near zero
585 < indicate vector directions indistinguishable from those calculated
586 < when using the reference method (SPME).
581 > {\it a priori} reason for the profile to adhere to any specific
582 > shape. Thus, gaussian fits were used to measure the width of the
583 > resulting distributions. The variance ($\sigma^2$) was extracted from
584 > each of these fits and was used to compare distribution widths.
585 > Values of $\sigma^2$ near zero indicate vector directions
586 > indistinguishable from those calculated when using the reference
587 > method ({\sc spme}).
588  
589   \subsection{Short-time Dynamics}
590 < Evaluation of the short-time dynamics of charged systems was performed
591 < by considering the 1000 K NaCl crystal system while using a subset of the
592 < best performing pairwise methods.  The NaCl crystal was chosen to
593 < avoid possible complications involving the propagation techniques of
594 < orientational motion in molecular systems.  All systems were started
595 < with the same initial positions and velocities.  Simulations were
596 < performed under the microcanonical ensemble, and velocity
590 >
591 > The effects of the alternative electrostatic summation methods on the
592 > short-time dynamics of charged systems were evaluated by considering a
593 > NaCl crystal at a temperature of 1000 K.  A subset of the best
594 > performing pairwise methods was used in this comparison.  The NaCl
595 > crystal was chosen to avoid possible complications from the treatment
596 > of orientational motion in molecular systems.  All systems were
597 > started with the same initial positions and velocities.  Simulations
598 > were performed under the microcanonical ensemble, and velocity
599   autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
600   of the trajectories,
601   \begin{equation}
602 < C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
602 > C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
603   \label{eq:vCorr}
604   \end{equation}
605   Velocity autocorrelation functions require detailed short time data,
# Line 617 | Line 609 | functions was used for comparisons.
609   functions was used for comparisons.
610  
611   \subsection{Long-Time and Collective Motion}\label{sec:LongTimeMethods}
612 < Evaluation of the long-time dynamics of charged systems was performed
613 < by considering the NaCl crystal system, again while using a subset of
614 < the best performing pairwise methods.  To enhance the atomic motion,
615 < these crystals were equilibrated at 1000 K, near the experimental
616 < $T_m$ for NaCl.  Simulations were performed under the microcanonical
617 < ensemble, and velocity information was saved every 5 fs over 100 ps
618 < trajectories.  The power spectrum ($I(\omega)$) was obtained via
627 < Fourier transform of the velocity autocorrelation function
628 < \begin{equation}
629 < I(\omega) = \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
612 >
613 > The effects of the same subset of alternative electrostatic methods on
614 > the {\it long-time} dynamics of charged systems were evaluated using
615 > the same model system (NaCl crystals at 1000~K).  The power spectrum
616 > ($I(\omega)$) was obtained via Fourier transform of the velocity
617 > autocorrelation function, \begin{equation} I(\omega) =
618 > \frac{1}{2\pi}\int^{\infty}_{-\infty}C_v(t)e^{-i\omega t}dt,
619   \label{eq:powerSpec}
620   \end{equation}
621   where the frequency, $\omega=0,\ 1,\ ...,\ N-1$. Again, because the
622   NaCl crystal is composed of two different atom types, the average of
623 < the two resulting power spectra was used for comparisons.
623 > the two resulting power spectra was used for comparisons. Simulations
624 > were performed under the microcanonical ensemble, and velocity
625 > information was saved every 5~fs over 100~ps trajectories.
626  
627   \subsection{Representative Simulations}\label{sec:RepSims}
628 < A variety of common and representative simulations were analyzed to
628 > A variety of representative molecular simulations were analyzed to
629   determine the relative effectiveness of the pairwise summation
630   techniques in reproducing the energetics and dynamics exhibited by
631 < SPME.  The studied systems were as follows:
631 > {\sc spme}.  We wanted to span the space of typical molecular
632 > simulations (i.e. from liquids of neutral molecules to ionic
633 > crystals), so the systems studied were:
634   \begin{enumerate}
635 < \item Liquid Water
636 < \item Crystalline Water (Ice I$_\textrm{c}$)
637 < \item NaCl Crystal
638 < \item NaCl Melt
639 < \item Low Ionic Strength Solution of NaCl in Water
640 < \item High Ionic Strength Solution of NaCl in Water
641 < \item 6 \AA\  Radius Sphere of Argon in Water
635 > \item liquid water (SPC/E),\cite{Berendsen87}
636 > \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
637 > \item NaCl crystals,
638 > \item NaCl melts,
639 > \item a low ionic strength solution of NaCl in water (0.11 M),
640 > \item a high ionic strength solution of NaCl in water (1.1 M), and
641 > \item a 6 \AA\  radius sphere of Argon in water.
642   \end{enumerate}
643   By utilizing the pairwise techniques (outlined in section
644   \ref{sec:ESMethods}) in systems composed entirely of neutral groups,
645 < charged particles, and mixtures of the two, we can comment on possible
646 < system dependence and/or universal applicability of the techniques.
645 > charged particles, and mixtures of the two, we hope to discern under
646 > which conditions it will be possible to use one of the alternative
647 > summation methodologies instead of the Ewald sum.
648  
649 < Generation of the system configurations was dependent on the system
650 < type.  For the solid and liquid water configurations, configuration
651 < snapshots were taken at regular intervals from higher temperature 1000
652 < SPC/E water molecule trajectories and each equilibrated
653 < individually.\cite{Berendsen87} The solid and liquid NaCl systems
654 < consisted of 500 Na+ and 500 Cl- ions and were selected and
655 < equilibrated in the same fashion as the water systems.  For the low
656 < and high ionic strength NaCl solutions, 4 and 40 ions were first
657 < solvated in a 1000 water molecule boxes respectively.  Ion and water
658 < positions were then randomly swapped, and the resulting configurations
659 < were again equilibrated individually.  Finally, for the Argon/Water
660 < "charge void" systems, the identities of all the SPC/E waters within 6
661 < \AA\ of the center of the equilibrated water configurations were
662 < converted to argon (Fig. \ref{fig:argonSlice}).
649 > For the solid and liquid water configurations, configurations were
650 > taken at regular intervals from high temperature trajectories of 1000
651 > SPC/E water molecules.  Each configuration was equilibrated
652 > independently at a lower temperature (300~K for the liquid, 200~K for
653 > the crystal).  The solid and liquid NaCl systems consisted of 500
654 > $\textrm{Na}^{+}$ and 500 $\textrm{Cl}^{-}$ ions.  Configurations for
655 > these systems were selected and equilibrated in the same manner as the
656 > water systems. In order to introduce measurable fluctuations in the
657 > configuration energy differences, the crystalline simulations were
658 > equilibrated at 1000~K, near the $T_\textrm{m}$ for NaCl. The liquid
659 > NaCl configurations needed to represent a fully disordered array of
660 > point charges, so the high temperature of 7000~K was selected for
661 > equilibration. The ionic solutions were made by solvating 4 (or 40)
662 > ions in a periodic box containing 1000 SPC/E water molecules.  Ion and
663 > water positions were then randomly swapped, and the resulting
664 > configurations were again equilibrated individually.  Finally, for the
665 > Argon / Water ``charge void'' systems, the identities of all the SPC/E
666 > waters within 6 \AA\ of the center of the equilibrated water
667 > configurations were converted to argon.
668  
669 < \begin{figure}
670 < \centering
671 < \includegraphics[width = \linewidth]{./slice.pdf}
672 < \caption{A slice from the center of a water box used in a charge void simulation.  The darkened region represents the boundary sphere within which the water molecules were converted to argon atoms.}
674 < \label{fig:argonSlice}
675 < \end{figure}
669 > These procedures guaranteed us a set of representative configurations
670 > from chemically-relevant systems sampled from appropriate
671 > ensembles. Force field parameters for the ions and Argon were taken
672 > from the force field utilized by {\sc oopse}.\cite{Meineke05}
673  
674 < \subsection{Electrostatic Summation Methods}\label{sec:ESMethods}
675 < Electrostatic summation method comparisons were performed using SPME,
676 < the {\sc sp} and {\sc sf} methods - both with damping
677 < parameters ($\alpha$) of 0.0, 0.1, 0.2, and 0.3 \AA$^{-1}$ (no, weak,
678 < moderate, and strong damping respectively), reaction field with an
679 < infinite dielectric constant, and an unmodified cutoff.  Group-based
680 < cutoffs with a fifth-order polynomial switching function were
681 < necessary for the reaction field simulations and were utilized in the
682 < SP, SF, and pure cutoff methods for comparison to the standard lack of
683 < group-based cutoffs with a hard truncation.  The SPME calculations
684 < were performed using the TINKER implementation of SPME,\cite{Ponder87}
685 < while all other method calculations were performed using the OOPSE
686 < molecular mechanics package.\cite{Meineke05}
674 > \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
675 > We compared the following alternative summation methods with results
676 > from the reference method ({\sc spme}):
677 > \begin{itemize}
678 > \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
679 > and 0.3 \AA$^{-1}$,
680 > \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
681 > and 0.3 \AA$^{-1}$,
682 > \item reaction field with an infinite dielectric constant, and
683 > \item an unmodified cutoff.
684 > \end{itemize}
685 > Group-based cutoffs with a fifth-order polynomial switching function
686 > were utilized for the reaction field simulations.  Additionally, we
687 > investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
688 > cutoff.  The {\sc spme} electrostatics were performed using the {\sc tinker}
689 > implementation of {\sc spme},\cite{Ponder87} while all other calculations
690 > were performed using the {\sc oopse} molecular mechanics
691 > package.\cite{Meineke05} All other portions of the energy calculation
692 > (i.e. Lennard-Jones interactions) were handled in exactly the same
693 > manner across all systems and configurations.
694  
695 < These methods were additionally evaluated with three different cutoff
696 < radii (9, 12, and 15 \AA) to investigate possible cutoff radius
697 < dependence.  It should be noted that the damping parameter chosen in
698 < SPME, or so called ``Ewald Coefficient", has a significant effect on
699 < the energies and forces calculated.  Typical molecular mechanics
700 < packages default this to a value dependent on the cutoff radius and a
701 < tolerance (typically less than $1 \times 10^{-4}$ kcal/mol).  Smaller
702 < tolerances are typically associated with increased accuracy, but this
703 < usually means more time spent calculating the reciprocal-space portion
704 < of the summation.\cite{Perram88,Essmann95} The default TINKER
705 < tolerance of $1 \times 10^{-8}$ kcal/mol was used in all SPME
706 < calculations, resulting in Ewald Coefficients of 0.4200, 0.3119, and
707 < 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\ respectively.
695 > The alternative methods were also evaluated with three different
696 > cutoff radii (9, 12, and 15 \AA).  As noted previously, the
697 > convergence parameter ($\alpha$) plays a role in the balance of the
698 > real-space and reciprocal-space portions of the Ewald calculation.
699 > Typical molecular mechanics packages set this to a value dependent on
700 > the cutoff radius and a tolerance (typically less than $1 \times
701 > 10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with
702 > increasing accuracy at the expense of computational time spent on the
703 > reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
704 > The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
705 > in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
706 > 0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
707 > respectively.
708  
709   \section{Results and Discussion}
710  
# Line 708 | Line 712 | between configurations were compared to the values obt
712   In order to evaluate the performance of the pairwise electrostatic
713   summation methods for Monte Carlo simulations, the energy differences
714   between configurations were compared to the values obtained when using
715 < SPME.  The results for the subsequent regression analysis are shown in
715 > {\sc spme}.  The results for the subsequent regression analysis are shown in
716   figure \ref{fig:delE}.
717  
718   \begin{figure}
719   \centering
720   \includegraphics[width=5.5in]{./delEplot.pdf}
721 < \caption{Statistical analysis of the quality of configurational energy differences for a given electrostatic method compared with the reference Ewald sum.  Results with a value equal to 1 (dashed line) indicate $\Delta E$ values indistinguishable from those obtained using SPME.  Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
721 > \caption{Statistical analysis of the quality of configurational energy
722 > differences for a given electrostatic method compared with the
723 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
724 > indicate $\Delta E$ values indistinguishable from those obtained using
725 > {\sc spme}.  Different values of the cutoff radius are indicated with
726 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
727 > inverted triangles).}
728   \label{fig:delE}
729   \end{figure}
730  
731 < In this figure, it is apparent that it is unreasonable to expect
732 < realistic results using an unmodified cutoff.  This is not all that
733 < surprising since this results in large energy fluctuations as atoms or
734 < molecules move in and out of the cutoff radius.\cite{Rahman71,Adams79}
735 < These fluctuations can be alleviated to some degree by using group
736 < based cutoffs with a switching
727 < function.\cite{Adams79,Steinbach94,Leach01} The Group Switch Cutoff
728 < row doesn't show a significant improvement in this plot because the
729 < salt and salt solution systems contain non-neutral groups, see the
730 < accompanying supporting information for a comparison where all groups
731 < are neutral.
731 > The most striking feature of this plot is how well the Shifted Force
732 > ({\sc sf}) and Shifted Potential ({\sc sp}) methods capture the energy
733 > differences.  For the undamped {\sc sf} method, and the
734 > moderately-damped {\sc sp} methods, the results are nearly
735 > indistinguishable from the Ewald results.  The other common methods do
736 > significantly less well.  
737  
738 < Correcting the resulting charged cutoff sphere is one of the purposes
739 < of the damped Coulomb summation proposed by Wolf \textit{et
740 < al.},\cite{Wolf99} and this correction indeed improves the results as
741 < seen in the {\sc sp} rows.  While the undamped case of this
742 < method is a significant improvement over the pure cutoff, it still
743 < doesn't correlate that well with SPME.  Inclusion of potential damping
744 < improves the results, and using an $\alpha$ of 0.2 \AA $^{-1}$ shows
745 < an excellent correlation and quality of fit with the SPME results,
746 < particularly with a cutoff radius greater than 12 \AA .  Use of a
747 < larger damping parameter is more helpful for the shortest cutoff
743 < shown, but it has a detrimental effect on simulations with larger
744 < cutoffs.  In the {\sc sf} sets, increasing damping results in
745 < progressively poorer correlation.  Overall, the undamped case is the
746 < best performing set, as the correlation and quality of fits are
747 < consistently superior regardless of the cutoff distance.  This result
748 < is beneficial in that the undamped case is less computationally
749 < prohibitive do to the lack of complimentary error function calculation
750 < when performing the electrostatic pair interaction.  The reaction
751 < field results illustrates some of that method's limitations, primarily
752 < that it was developed for use in homogenous systems; although it does
753 < provide results that are an improvement over those from an unmodified
754 < cutoff.
738 > The unmodified cutoff method is essentially unusable.  This is not
739 > surprising since hard cutoffs give large energy fluctuations as atoms
740 > or molecules move in and out of the cutoff
741 > radius.\cite{Rahman71,Adams79} These fluctuations can be alleviated to
742 > some degree by using group based cutoffs with a switching
743 > function.\cite{Adams79,Steinbach94,Leach01} However, we do not see
744 > significant improvement using the group-switched cutoff because the
745 > salt and salt solution systems contain non-neutral groups.  Interested
746 > readers can consult the accompanying supporting information for a
747 > comparison where all groups are neutral.
748  
749 + For the {\sc sp} method, inclusion of electrostatic damping improves
750 + the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
751 + shows an excellent correlation and quality of fit with the {\sc spme}
752 + results, particularly with a cutoff radius greater than 12
753 + \AA .  Use of a larger damping parameter is more helpful for the
754 + shortest cutoff shown, but it has a detrimental effect on simulations
755 + with larger cutoffs.  
756 +
757 + In the {\sc sf} sets, increasing damping results in progressively {\it
758 + worse} correlation with Ewald.  Overall, the undamped case is the best
759 + performing set, as the correlation and quality of fits are
760 + consistently superior regardless of the cutoff distance.  The undamped
761 + case is also less computationally demanding (because no evaluation of
762 + the complementary error function is required).
763 +
764 + The reaction field results illustrates some of that method's
765 + limitations, primarily that it was developed for use in homogenous
766 + systems; although it does provide results that are an improvement over
767 + those from an unmodified cutoff.
768 +
769   \subsection{Magnitudes of the Force and Torque Vectors}
770  
771   Evaluation of pairwise methods for use in Molecular Dynamics
772   simulations requires consideration of effects on the forces and
773 < torques.  Investigation of the force and torque vector magnitudes
774 < provides a measure of the strength of these values relative to SPME.
775 < Figures \ref{fig:frcMag} and \ref{fig:trqMag} respectively show the
776 < force and torque vector magnitude regression results for the
764 < accumulated analysis over all the system types.
773 > torques.  Figures \ref{fig:frcMag} and \ref{fig:trqMag} show the
774 > regression results for the force and torque vector magnitudes,
775 > respectively.  The data in these figures was generated from an
776 > accumulation of the statistics from all of the system types.
777  
778   \begin{figure}
779   \centering
780   \includegraphics[width=5.5in]{./frcMagplot.pdf}
781 < \caption{Statistical analysis of the quality of the force vector magnitudes for a given electrostatic method compared with the reference Ewald sum.  Results with a value equal to 1 (dashed line) indicate force magnitude values indistinguishable from those obtained using SPME.  Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
781 > \caption{Statistical analysis of the quality of the force vector
782 > magnitudes for a given electrostatic method compared with the
783 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
784 > indicate force magnitude values indistinguishable from those obtained
785 > using {\sc spme}.  Different values of the cutoff radius are indicated with
786 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
787 > inverted triangles).}
788   \label{fig:frcMag}
789   \end{figure}
790  
791 + Again, it is striking how well the Shifted Potential and Shifted Force
792 + methods are doing at reproducing the {\sc spme} forces.  The undamped and
793 + weakly-damped {\sc sf} method gives the best agreement with Ewald.
794 + This is perhaps expected because this method explicitly incorporates a
795 + smooth transition in the forces at the cutoff radius as well as the
796 + neutralizing image charges.
797 +
798   Figure \ref{fig:frcMag}, for the most part, parallels the results seen
799   in the previous $\Delta E$ section.  The unmodified cutoff results are
800   poor, but using group based cutoffs and a switching function provides
801 < a improvement much more significant than what was seen with $\Delta
802 < E$.  Looking at the {\sc sp} sets, the slope and $R^2$
803 < improve with the use of damping to an optimal result of 0.2 \AA
804 < $^{-1}$ for the 12 and 15 \AA\ cutoffs.  Further increases in damping,
801 > an improvement much more significant than what was seen with $\Delta
802 > E$.
803 >
804 > With moderate damping and a large enough cutoff radius, the {\sc sp}
805 > method is generating usable forces.  Further increases in damping,
806   while beneficial for simulations with a cutoff radius of 9 \AA\ , is
807 < detrimental to simulations with larger cutoff radii.  The undamped
808 < {\sc sf} method gives forces in line with those obtained using
809 < SPME, and use of a damping function results in minor improvement.  The
784 < reaction field results are surprisingly good, considering the poor
807 > detrimental to simulations with larger cutoff radii.
808 >
809 > The reaction field results are surprisingly good, considering the poor
810   quality of the fits for the $\Delta E$ results.  There is still a
811 < considerable degree of scatter in the data, but it correlates well in
812 < general.  To be fair, we again note that the reaction field
813 < calculations do not encompass NaCl crystal and melt systems, so these
811 > considerable degree of scatter in the data, but the forces correlate
812 > well with the Ewald forces in general.  We note that the reaction
813 > field calculations do not include the pure NaCl systems, so these
814   results are partly biased towards conditions in which the method
815   performs more favorably.
816  
817   \begin{figure}
818   \centering
819   \includegraphics[width=5.5in]{./trqMagplot.pdf}
820 < \caption{Statistical analysis of the quality of the torque vector magnitudes for a given electrostatic method compared with the reference Ewald sum.  Results with a value equal to 1 (dashed line) indicate torque magnitude values indistinguishable from those obtained using SPME.  Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
820 > \caption{Statistical analysis of the quality of the torque vector
821 > magnitudes for a given electrostatic method compared with the
822 > reference Ewald sum.  Results with a value equal to 1 (dashed line)
823 > indicate torque magnitude values indistinguishable from those obtained
824 > using {\sc spme}.  Different values of the cutoff radius are indicated with
825 > different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
826 > inverted triangles).}
827   \label{fig:trqMag}
828   \end{figure}
829  
830 < To evaluate the torque vector magnitudes, the data set from which
831 < values are drawn is limited to rigid molecules in the systems
832 < (i.e. water molecules).  In spite of this smaller sampling pool, the
802 < torque vector magnitude results in figure \ref{fig:trqMag} are still
803 < similar to those seen for the forces; however, they more clearly show
804 < the improved behavior that comes with increasing the cutoff radius.
805 < Moderate damping is beneficial to the {\sc sp} and helpful
806 < yet possibly unnecessary with the {\sc sf} method, and they also
807 < show that over-damping adversely effects all cutoff radii rather than
808 < showing an improvement for systems with short cutoffs.  The reaction
809 < field method performs well when calculating the torques, better than
810 < the Shifted Force method over this limited data set.
830 > Molecular torques were only available from the systems which contained
831 > rigid molecules (i.e. the systems containing water).  The data in
832 > fig. \ref{fig:trqMag} is taken from this smaller sampling pool.
833  
834 + Torques appear to be much more sensitive to charges at a longer
835 + distance.   The striking feature in comparing the new electrostatic
836 + methods with {\sc spme} is how much the agreement improves with increasing
837 + cutoff radius.  Again, the weakly damped and undamped {\sc sf} method
838 + appears to be reproducing the {\sc spme} torques most accurately.  
839 +
840 + Water molecules are dipolar, and the reaction field method reproduces
841 + the effect of the surrounding polarized medium on each of the
842 + molecular bodies. Therefore it is not surprising that reaction field
843 + performs best of all of the methods on molecular torques.
844 +
845   \subsection{Directionality of the Force and Torque Vectors}
846  
847 < Having force and torque vectors with magnitudes that are well
848 < correlated to SPME is good, but if they are not pointing in the proper
849 < direction the results will be incorrect.  These vector directions were
850 < investigated through measurement of the angle formed between them and
851 < those from SPME.  The results (Fig. \ref{fig:frcTrqAng}) are compared
852 < through the variance ($\sigma^2$) of the Gaussian fits of the angle
853 < error distributions of the combined set over all system types.
847 > It is clearly important that a new electrostatic method can reproduce
848 > the magnitudes of the force and torque vectors obtained via the Ewald
849 > sum. However, the {\it directionality} of these vectors will also be
850 > vital in calculating dynamical quantities accurately.  Force and
851 > torque directionalities were investigated by measuring the angles
852 > formed between these vectors and the same vectors calculated using
853 > {\sc spme}.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the
854 > variance ($\sigma^2$) of the Gaussian fits of the angle error
855 > distributions of the combined set over all system types.
856  
857   \begin{figure}
858   \centering
859   \includegraphics[width=5.5in]{./frcTrqAngplot.pdf}
860 < \caption{Statistical analysis of the quality of the Gaussian fit of the force and torque vector angular distributions for a given electrostatic method compared with the reference Ewald sum.  Results with a variance ($\sigma^2$) equal to zero (dashed line) indicate force and torque directions indistinguishable from those obtained using SPME.  Different values of the cutoff radius are indicated with different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
860 > \caption{Statistical analysis of the width of the angular distribution
861 > that the force and torque vectors from a given electrostatic method
862 > make with their counterparts obtained using the reference Ewald sum.
863 > Results with a variance ($\sigma^2$) equal to zero (dashed line)
864 > indicate force and torque directions indistinguishable from those
865 > obtained using {\sc spme}.  Different values of the cutoff radius are
866 > indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
867 > and 15\AA\ = inverted triangles).}
868   \label{fig:frcTrqAng}
869   \end{figure}
870  
871   Both the force and torque $\sigma^2$ results from the analysis of the
872   total accumulated system data are tabulated in figure
873 < \ref{fig:frcTrqAng}.  All of the sets, aside from the over-damped case
874 < show the improvement afforded by choosing a longer simulation cutoff.
875 < Increasing the cutoff from 9 to 12 \AA\ typically results in a halving
876 < of the distribution widths, with a similar improvement going from 12
877 < to 15 \AA .  The undamped {\sc sf}, Group Based Cutoff, and
878 < Reaction Field methods all do equivalently well at capturing the
837 < direction of both the force and torque vectors.  Using damping
838 < improves the angular behavior significantly for the {\sc sp}
839 < and moderately for the {\sc sf} methods.  Increasing the damping
840 < too far is destructive for both methods, particularly to the torque
841 < vectors.  Again it is important to recognize that the force vectors
842 < cover all particles in the systems, while torque vectors are only
843 < available for neutral molecular groups.  Damping appears to have a
844 < more beneficial effect on non-neutral bodies, and this observation is
845 < investigated further in the accompanying supporting information.
873 > \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
874 > sp}) method would be essentially unusable for molecular dynamics
875 > unless the damping function is added.  The Shifted Force ({\sc sf})
876 > method, however, is generating force and torque vectors which are
877 > within a few degrees of the Ewald results even with weak (or no)
878 > damping.
879  
880 + All of the sets (aside from the over-damped case) show the improvement
881 + afforded by choosing a larger cutoff radius.  Increasing the cutoff
882 + from 9 to 12 \AA\ typically results in a halving of the width of the
883 + distribution, with a similar improvement when going from 12 to 15
884 + \AA .
885 +
886 + The undamped {\sc sf}, group-based cutoff, and reaction field methods
887 + all do equivalently well at capturing the direction of both the force
888 + and torque vectors.  Using the electrostatic damping improves the
889 + angular behavior significantly for the {\sc sp} and moderately for the
890 + {\sc sf} methods.  Overdamping is detrimental to both methods.  Again
891 + it is important to recognize that the force vectors cover all
892 + particles in all seven systems, while torque vectors are only
893 + available for neutral molecular groups.  Damping is more beneficial to
894 + charged bodies, and this observation is investigated further in the
895 + accompanying supporting information.
896 +
897 + Although not discussed previously, group based cutoffs can be applied
898 + to both the {\sc sp} and {\sc sf} methods.  The group-based cutoffs
899 + will reintroduce small discontinuities at the cutoff radius, but the
900 + effects of these can be minimized by utilizing a switching function.
901 + Though there are no significant benefits or drawbacks observed in
902 + $\Delta E$ and the force and torque magnitudes when doing this, there
903 + is a measurable improvement in the directionality of the forces and
904 + torques. Table \ref{tab:groupAngle} shows the angular variances
905 + obtained using group based cutoffs along with the results seen in
906 + figure \ref{fig:frcTrqAng}.  The {\sc sp} (with an $\alpha$ of 0.2
907 + \AA$^{-1}$ or smaller) shows much narrower angular distributions when
908 + using group-based cutoffs. The {\sc sf} method likewise shows
909 + improvement in the undamped and lightly damped cases.
910 +
911   \begin{table}[htbp]
912 <   \centering
913 <   \caption{Variance ($\sigma^2$) of the force (top set) and torque (bottom set) vector angle difference distributions for the Shifted Potential and Shifted Force methods.  Calculations were performed both with (Y) and without (N) group based cutoffs and a switching function.  The $\alpha$ values have units of \AA$^{-1}$ and the variance values have units of degrees$^2$.}  
912 >   \centering
913 >   \caption{Statistical analysis of the angular
914 >   distributions that the force (upper) and torque (lower) vectors
915 >   from a given electrostatic method make with their counterparts
916 >   obtained using the reference Ewald sum.  Calculations were
917 >   performed both with (Y) and without (N) group based cutoffs and a
918 >   switching function.  The $\alpha$ values have units of \AA$^{-1}$
919 >   and the variance values have units of degrees$^2$.}
920 >
921     \begin{tabular}{@{} ccrrrrrrrr @{}}
922        \\
923        \toprule
# Line 877 | Line 948 | investigated further in the accompanying supporting in
948     \label{tab:groupAngle}
949   \end{table}
950  
951 < Although not discussed previously, group based cutoffs can be applied
952 < to both the {\sc sp} and {\sc sf} methods.  Use off a
953 < switching function corrects for the discontinuities that arise when
954 < atoms of a group exit the cutoff before the group's center of mass.
955 < Though there are no significant benefit or drawbacks observed in
956 < $\Delta E$ and vector magnitude results when doing this, there is a
957 < measurable improvement in the vector angle results.  Table
958 < \ref{tab:groupAngle} shows the angular variance values obtained using
959 < group based cutoffs and a switching function alongside the standard
960 < results seen in figure \ref{fig:frcTrqAng} for comparison purposes.
961 < The {\sc sp} shows much narrower angular distributions for
962 < both the force and torque vectors when using an $\alpha$ of 0.2
963 < \AA$^{-1}$ or less, while {\sc sf} shows improvements in the
964 < undamped and lightly damped cases.  Thus, by calculating the
965 < electrostatic interactions in terms of molecular pairs rather than
966 < atomic pairs, the direction of the force and torque vectors are
967 < determined more accurately.
951 > One additional trend in table \ref{tab:groupAngle} is that the
952 > $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
953 > increases, something that is more obvious with group-based cutoffs.
954 > The complimentary error function inserted into the potential weakens
955 > the electrostatic interaction as the value of $\alpha$ is increased.
956 > However, at larger values of $\alpha$, it is possible to overdamp the
957 > electrostatic interaction and to remove it completely.  Kast
958 > \textit{et al.}  developed a method for choosing appropriate $\alpha$
959 > values for these types of electrostatic summation methods by fitting
960 > to $g(r)$ data, and their methods indicate optimal values of 0.34,
961 > 0.25, and 0.16 \AA$^{-1}$ for cutoff values of 9, 12, and 15 \AA\
962 > respectively.\cite{Kast03} These appear to be reasonable choices to
963 > obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
964 > these findings, choices this high would introduce error in the
965 > molecular torques, particularly for the shorter cutoffs.  Based on our
966 > observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
967 > but damping may be unnecessary when using the {\sc sf} method.
968  
898 One additional trend to recognize in table \ref{tab:groupAngle} is
899 that the $\sigma^2$ values for both {\sc sp} and
900 {\sc sf} converge as $\alpha$ increases, something that is easier
901 to see when using group based cutoffs.  Looking back on figures
902 \ref{fig:delE}, \ref{fig:frcMag}, and \ref{fig:trqMag}, show this
903 behavior clearly at large $\alpha$ and cutoff values.  The reason for
904 this is that the complimentary error function inserted into the
905 potential weakens the electrostatic interaction as $\alpha$ increases.
906 Thus, at larger values of $\alpha$, both the summation method types
907 progress toward non-interacting functions, so care is required in
908 choosing large damping functions lest one generate an undesirable loss
909 in the pair interaction.  Kast \textit{et al.}  developed a method for
910 choosing appropriate $\alpha$ values for these types of electrostatic
911 summation methods by fitting to $g(r)$ data, and their methods
912 indicate optimal values of 0.34, 0.25, and 0.16 \AA$^{-1}$ for cutoff
913 values of 9, 12, and 15 \AA\ respectively.\cite{Kast03} These appear
914 to be reasonable choices to obtain proper MC behavior
915 (Fig. \ref{fig:delE}); however, based on these findings, choices this
916 high would introduce error in the molecular torques, particularly for
917 the shorter cutoffs.  Based on the above findings, empirical damping
918 up to 0.2 \AA$^{-1}$ proves to be beneficial, but damping is arguably
919 unnecessary when using the {\sc sf} method.
920
969   \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
970  
971 < In the previous studies using a {\sc sf} variant of the damped
972 < Wolf coulomb potential, the structure and dynamics of water were
973 < investigated rather extensively.\cite{Zahn02,Kast03} Their results
974 < indicated that the damped {\sc sf} method results in properties
975 < very similar to those obtained when using the Ewald summation.
976 < Considering the statistical results shown above, the good performance
977 < of this method is not that surprising.  Rather than consider the same
978 < systems and simply recapitulate their results, we decided to look at
979 < the solid state dynamical behavior obtained using the best performing
980 < summation methods from the above results.
971 > Zahn {\it et al.} investigated the structure and dynamics of water
972 > using eqs. (\ref{eq:ZahnPot}) and
973 > (\ref{eq:WolfForces}).\cite{Zahn02,Kast03} Their results indicated
974 > that a method similar (but not identical with) the damped {\sc sf}
975 > method resulted in properties very similar to those obtained when
976 > using the Ewald summation.  The properties they studied (pair
977 > distribution functions, diffusion constants, and velocity and
978 > orientational correlation functions) may not be particularly sensitive
979 > to the long-range and collective behavior that governs the
980 > low-frequency behavior in crystalline systems.  Additionally, the
981 > ionic crystals are the worst case scenario for the pairwise methods
982 > because they lack the reciprocal space contribution contained in the
983 > Ewald summation.  
984  
985 + We are using two separate measures to probe the effects of these
986 + alternative electrostatic methods on the dynamics in crystalline
987 + materials.  For short- and intermediate-time dynamics, we are
988 + computing the velocity autocorrelation function, and for long-time
989 + and large length-scale collective motions, we are looking at the
990 + low-frequency portion of the power spectrum.
991 +
992   \begin{figure}
993   \centering
994   \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
995 < \caption{Velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2). The inset is a magnification of the first trough. The times to first collision are nearly identical, but the differences can be seen in the peaks and troughs, where the undamped to weakly damped methods are stiffer than the moderately damped and SPME methods.}
995 > \caption{Velocity autocorrelation functions of NaCl crystals at
996 > 1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
997 > sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
998 > the first minimum.  The times to first collision are nearly identical,
999 > but differences can be seen in the peaks and troughs, where the
1000 > undamped and weakly damped methods are stiffer than the moderately
1001 > damped and {\sc spme} methods.}
1002   \label{fig:vCorrPlot}
1003   \end{figure}
1004  
1005 < The short-time decays through the first collision are nearly identical
1006 < in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
1007 < functions show how the methods differ.  The undamped {\sc sf} method
1008 < has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
1009 < peaks than any of the other methods.  As the damping function is
1010 < increased, these peaks are smoothed out, and approach the SPME
1011 < curve. The damping acts as a distance dependent Gaussian screening of
1012 < the point charges for the pairwise summation methods; thus, the
1013 < collisions are more elastic in the undamped {\sc sf} potential, and the
1014 < stiffness of the potential is diminished as the electrostatic
1015 < interactions are softened by the damping function.  With $\alpha$
1016 < values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
1017 < nearly identical and track the SPME features quite well.  This is not
1018 < too surprising in that the differences between the {\sc sf} and {\sc
1019 < sp} potentials are mitigated with increased damping.  However, this
956 < appears to indicate that once damping is utilized, the form of the
957 < potential seems to play a lesser role in the crystal dynamics.
1005 > The short-time decay of the velocity autocorrelation function through
1006 > the first collision are nearly identical in figure
1007 > \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1008 > how the methods differ.  The undamped {\sc sf} method has deeper
1009 > troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1010 > any of the other methods.  As the damping parameter ($\alpha$) is
1011 > increased, these peaks are smoothed out, and the {\sc sf} method
1012 > approaches the {\sc spme} results.  With $\alpha$ values of 0.2 \AA$^{-1}$,
1013 > the {\sc sf} and {\sc sp} functions are nearly identical and track the
1014 > {\sc spme} features quite well.  This is not surprising because the {\sc sf}
1015 > and {\sc sp} potentials become nearly identical with increased
1016 > damping.  However, this appears to indicate that once damping is
1017 > utilized, the details of the form of the potential (and forces)
1018 > constructed out of the damped electrostatic interaction are less
1019 > important.
1020  
1021   \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1022  
1023 < The short time dynamics were extended to evaluate how the differences
1024 < between the methods affect the collective long-time motion.  The same
1025 < electrostatic summation methods were used as in the short time
1026 < velocity autocorrelation function evaluation, but the trajectories
1027 < were sampled over a much longer time. The power spectra of the
1028 < resulting velocity autocorrelation functions were calculated and are
1029 < displayed in figure \ref{fig:methodPS}.
1023 > To evaluate how the differences between the methods affect the
1024 > collective long-time motion, we computed power spectra from long-time
1025 > traces of the velocity autocorrelation function. The power spectra for
1026 > the best-performing alternative methods are shown in
1027 > fig. \ref{fig:methodPS}.  Apodization of the correlation functions via
1028 > a cubic switching function between 40 and 50 ps was used to reduce the
1029 > ringing resulting from data truncation.  This procedure had no
1030 > noticeable effect on peak location or magnitude.
1031  
1032   \begin{figure}
1033   \centering
1034   \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1035 < \caption{Power spectra obtained from the velocity auto-correlation functions of NaCl crystals at 1000 K while using SPME, {\sc sf} ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).  Apodization of the correlation functions via a cubic switching function between 40 and 50 ps was used to clear up the spectral noise resulting from data truncation, and had no noticeable effect on peak location or magnitude.  The inset shows the frequency region below 100 cm$^{-1}$ to highlight where the spectra begin to differ.}
1035 > \caption{Power spectra obtained from the velocity auto-correlation
1036 > functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf}
1037 > ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).  The inset
1038 > shows the frequency region below 100 cm$^{-1}$ to highlight where the
1039 > spectra differ.}
1040   \label{fig:methodPS}
1041   \end{figure}
1042  
1043 < While high frequency peaks of the spectra in this figure overlap,
1044 < showing the same general features, the low frequency region shows how
1045 < the summation methods differ.  Considering the low-frequency inset
1046 < (expanded in the upper frame of figure \ref{fig:dampInc}), at
1047 < frequencies below 100 cm$^{-1}$, the correlated motions are
1048 < blue-shifted when using undamped or weakly damped {\sc sf}.  When
1049 < using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
1050 < and {\sc sp} methods give near identical correlated motion behavior as
1051 < the Ewald method (which has a damping value of 0.3119).  This
1052 < weakening of the electrostatic interaction with increased damping
1053 < explains why the long-ranged correlated motions are at lower
1054 < frequencies for the moderately damped methods than for undamped or
1055 < weakly damped methods.  To see this effect more clearly, we show how
1056 < damping strength alone affects a simple real-space electrostatic
1057 < potential,
1058 < \begin{equation}
1059 < V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
1060 < \end{equation}
1061 < where $S(r)$ is a switching function that smoothly zeroes the
1062 < potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how
1063 < the low frequency motions are dependent on the damping used in the
1064 < direct electrostatic sum.  As the damping increases, the peaks drop to
1065 < lower frequencies.  Incidentally, use of an $\alpha$ of 0.25
1066 < \AA$^{-1}$ on a simple electrostatic summation results in low
1067 < frequency correlated dynamics equivalent to a simulation using SPME.
1068 < When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1069 < shift to higher frequency in exponential fashion.  Though not shown,
1070 < the spectrum for the simple undamped electrostatic potential is
1004 < blue-shifted such that the lowest frequency peak resides near 325
1005 < cm$^{-1}$.  In light of these results, the undamped {\sc sf} method
1006 < producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1007 < respectable and shows that the shifted force procedure accounts for
1008 < most of the effect afforded through use of the Ewald summation.
1009 < However, it appears as though moderate damping is required for
1010 < accurate reproduction of crystal dynamics.
1043 > While the high frequency regions of the power spectra for the
1044 > alternative methods are quantitatively identical with Ewald spectrum,
1045 > the low frequency region shows how the summation methods differ.
1046 > Considering the low-frequency inset (expanded in the upper frame of
1047 > figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1048 > correlated motions are blue-shifted when using undamped or weakly
1049 > damped {\sc sf}.  When using moderate damping ($\alpha = 0.2$
1050 > \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1051 > correlated motion to the Ewald method (which has a convergence
1052 > parameter of 0.3119 \AA$^{-1}$).  This weakening of the electrostatic
1053 > interaction with increased damping explains why the long-ranged
1054 > correlated motions are at lower frequencies for the moderately damped
1055 > methods than for undamped or weakly damped methods.
1056 >
1057 > To isolate the role of the damping constant, we have computed the
1058 > spectra for a single method ({\sc sf}) with a range of damping
1059 > constants and compared this with the {\sc spme} spectrum.
1060 > Fig. \ref{fig:dampInc} shows more clearly that increasing the
1061 > electrostatic damping red-shifts the lowest frequency phonon modes.
1062 > However, even without any electrostatic damping, the {\sc sf} method
1063 > has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1064 > Without the {\sc sf} modifications, an undamped (pure cutoff) method
1065 > would predict the lowest frequency peak near 325 cm$^{-1}$.  {\it
1066 > Most} of the collective behavior in the crystal is accurately captured
1067 > using the {\sc sf} method.  Quantitative agreement with Ewald can be
1068 > obtained using moderate damping in addition to the shifting at the
1069 > cutoff distance.
1070 >
1071   \begin{figure}
1072   \centering
1073 < \includegraphics[width = \linewidth]{./comboSquare.pdf}
1074 < \caption{Regions of spectra showing the low-frequency correlated motions for NaCl crystals at 1000 K using various electrostatic summation methods.  The upper plot is a zoomed inset from figure \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential increases, the low-frequency peaks red-shift.  The lower plot is of spectra when using SPME and a simple damped Coulombic sum with damping coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$.  As $\alpha$ increases, the peaks are red-shifted toward and eventually beyond the values given by SPME.  The larger $\alpha$ values weaken the real-space electrostatics, explaining this shift towards less strongly correlated motions in the crystal.}
1073 > \includegraphics[width = \linewidth]{./increasedDamping.pdf}
1074 > \caption{Effect of damping on the two lowest-frequency phonon modes in
1075 > the NaCl crystal at 1000~K.  The undamped shifted force ({\sc sf})
1076 > method is off by less than 10 cm$^{-1}$, and increasing the
1077 > electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1078 > with the power spectrum obtained using the Ewald sum.  Overdamping can
1079 > result in underestimates of frequencies of the long-wavelength
1080 > motions.}
1081   \label{fig:dampInc}
1082   \end{figure}
1083  
1084   \section{Conclusions}
1085  
1086   This investigation of pairwise electrostatic summation techniques
1087 < shows that there are viable and more computationally efficient
1088 < electrostatic summation techniques than the Ewald summation, chiefly
1089 < methods derived from the damped Coulombic sum originally proposed by
1090 < Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1091 < {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1092 < shows a remarkable ability to reproduce the energetic and dynamic
1093 < characteristics exhibited by simulations employing lattice summation
1094 < techniques.  The cumulative energy difference results showed the
1095 < undamped {\sc sf} and moderately damped {\sc sp} methods
1096 < produced results nearly identical to SPME.  Similarly for the dynamic
1097 < features, the undamped or moderately damped {\sc sf} and
1098 < moderately damped {\sc sp} methods produce force and torque
1099 < vector magnitude and directions very similar to the expected values.
1100 < These results translate into long-time dynamic behavior equivalent to
1101 < that produced in simulations using SPME.
1087 > shows that there are viable and computationally efficient alternatives
1088 > to the Ewald summation.  These methods are derived from the damped and
1089 > cutoff-neutralized Coulombic sum originally proposed by Wolf
1090 > \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1091 > method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1092 > (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1093 > energetic and dynamic characteristics exhibited by simulations
1094 > employing lattice summation techniques.  The cumulative energy
1095 > difference results showed the undamped {\sc sf} and moderately damped
1096 > {\sc sp} methods produced results nearly identical to {\sc spme}.  Similarly
1097 > for the dynamic features, the undamped or moderately damped {\sc sf}
1098 > and moderately damped {\sc sp} methods produce force and torque vector
1099 > magnitude and directions very similar to the expected values.  These
1100 > results translate into long-time dynamic behavior equivalent to that
1101 > produced in simulations using {\sc spme}.
1102  
1103 + As in all purely-pairwise cutoff methods, these methods are expected
1104 + to scale approximately {\it linearly} with system size, and they are
1105 + easily parallelizable.  This should result in substantial reductions
1106 + in the computational cost of performing large simulations.
1107 +
1108   Aside from the computational cost benefit, these techniques have
1109   applicability in situations where the use of the Ewald sum can prove
1110 < problematic.  Primary among them is their use in interfacial systems,
1111 < where the unmodified lattice sum techniques artificially accentuate
1112 < the periodicity of the system in an undesirable manner.  There have
1113 < been alterations to the standard Ewald techniques, via corrections and
1114 < reformulations, to compensate for these systems; but the pairwise
1115 < techniques discussed here require no modifications, making them
1116 < natural tools to tackle these problems.  Additionally, this
1117 < transferability gives them benefits over other pairwise methods, like
1118 < reaction field, because estimations of physical properties (e.g. the
1119 < dielectric constant) are unnecessary.
1110 > problematic.  Of greatest interest is their potential use in
1111 > interfacial systems, where the unmodified lattice sum techniques
1112 > artificially accentuate the periodicity of the system in an
1113 > undesirable manner.  There have been alterations to the standard Ewald
1114 > techniques, via corrections and reformulations, to compensate for
1115 > these systems; but the pairwise techniques discussed here require no
1116 > modifications, making them natural tools to tackle these problems.
1117 > Additionally, this transferability gives them benefits over other
1118 > pairwise methods, like reaction field, because estimations of physical
1119 > properties (e.g. the dielectric constant) are unnecessary.
1120  
1121 < We are not suggesting any flaw with the Ewald sum; in fact, it is the
1122 < standard by which these simple pairwise sums are judged.  However,
1123 < these results do suggest that in the typical simulations performed
1124 < today, the Ewald summation may no longer be required to obtain the
1125 < level of accuracy most researchers have come to expect
1121 > If a researcher is using Monte Carlo simulations of large chemical
1122 > systems containing point charges, most structural features will be
1123 > accurately captured using the undamped {\sc sf} method or the {\sc sp}
1124 > method with an electrostatic damping of 0.2 \AA$^{-1}$.  These methods
1125 > would also be appropriate for molecular dynamics simulations where the
1126 > data of interest is either structural or short-time dynamical
1127 > quantities.  For long-time dynamics and collective motions, the safest
1128 > pairwise method we have evaluated is the {\sc sf} method with an
1129 > electrostatic damping between 0.2 and 0.25
1130 > \AA$^{-1}$.
1131  
1132 + We are not suggesting that there is any flaw with the Ewald sum; in
1133 + fact, it is the standard by which these simple pairwise sums have been
1134 + judged.  However, these results do suggest that in the typical
1135 + simulations performed today, the Ewald summation may no longer be
1136 + required to obtain the level of accuracy most researchers have come to
1137 + expect.
1138 +
1139   \section{Acknowledgments}
1140 + Support for this project was provided by the National Science
1141 + Foundation under grant CHE-0134881.  The authors would like to thank
1142 + Steve Corcelli and Ed Maginn for helpful discussions and comments.
1143 +
1144   \newpage
1145  
1146   \bibliographystyle{jcp2}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines