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 2655 by gezelter, Tue Mar 21 20:46:55 2006 UTC vs.
Revision 2656 by gezelter, Wed Mar 22 20:49:53 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 necessary? \\
29 > Pairwise alternatives to the accepted standard for \\
30 > long-range electrostatics}
31  
32   \author{Christopher J. Fennell and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail:
33   gezelter@nd.edu} \\
# Line 40 | Line 42 | A new method for accumulating electrostatic interactio
42  
43   \nobibliography{}
44   \begin{abstract}
45 < A new method for accumulating electrostatic interactions was derived
46 < from the previous efforts described in \bibentry{Wolf99} and
47 < \bibentry{Zahn02} as a possible replacement for lattice sum methods in
48 < molecular simulations.  Comparisons were performed with this and other
49 < pairwise electrostatic summation techniques against the smooth
50 < particle mesh Ewald (SPME) summation to see how well they reproduce
51 < the energetics and dynamics of a variety of simulation types.  The
52 < newly derived Shifted-Force technique shows a remarkable ability to
53 < reproduce the behavior exhibited in simulations using SPME with an
54 < $\mathscr{O}(N)$ computational cost, equivalent to merely the
55 < real-space portion of the lattice summation.
56 <
45 > We investigate pairwise electrostatic interaction methods and show
46 > that there are viable and computationally efficient $(\mathscr{O}(N))$
47 > alternatives to the Ewald summation for typical modern molecular
48 > simulations.  These methods are extended from the damped and
49 > cutoff-neutralized Coulombic sum originally proposed by Wolf
50 > \textit{et al.}  One of these, the damped shifted force method, shows
51 > a remarkable ability to reproduce the energetic and dynamic
52 > characteristics exhibited by simulations employing lattice summation
53 > techniques.  Comparisons were performed with this and other pairwise
54 > methods against the smooth particle mesh Ewald ({\sc spme}) summation to see
55 > how well they reproduce the energetics and dynamics of a variety of
56 > simulation types.
57   \end{abstract}
58  
59   \newpage
# Line 165 | Line 167 | portion.\cite{Karasawa89,Kolafa92}
167  
168   \begin{figure}
169   \centering
170 < \includegraphics[width = \linewidth]{./ewaldProgression2.pdf}
170 > \includegraphics[width = \linewidth]{./ewaldProgression.pdf}
171   \caption{The change in the application of the Ewald sum with
172   increasing computational power.  Initially, only small systems could
173   be studied, and the Ewald sum replicated the simulation box to
# Line 516 | Line 518 | SPME-computed energy difference between two conformati
518   The pairwise summation techniques (outlined in section
519   \ref{sec:ESMethods}) were evaluated for use in MC simulations by
520   studying the energy differences between conformations.  We took the
521 < SPME-computed energy difference between two conformations to be the
521 > {\sc spme}-computed energy difference between two conformations to be the
522   correct behavior. An ideal performance by an alternative method would
523   reproduce these energy differences exactly (even if the absolute
524   energies calculated by the methods are different).  Since none of the
# Line 524 | Line 526 | indicate perfect agreement between the alternative met
526   regressions of energy gap data to evaluate how closely the methods
527   mimicked the Ewald energy gaps.  Unitary results for both the
528   correlation (slope) and correlation coefficient for these regressions
529 < indicate perfect agreement between the alternative method and SPME.
529 > indicate perfect agreement between the alternative method and {\sc spme}.
530   Sample correlation plots for two alternate methods are shown in
531   Fig. \ref{fig:linearFit}.
532  
# Line 553 | Line 555 | reference Ewald summation (SPME).  Both the magnitude
555   We evaluated the pairwise methods (outlined in section
556   \ref{sec:ESMethods}) for use in MD simulations by
557   comparing the force and torque vectors with those obtained using the
558 < reference Ewald summation (SPME).  Both the magnitude and the
558 > reference Ewald summation ({\sc spme}).  Both the magnitude and the
559   direction of these vectors on each of the bodies in the system were
560   analyzed.  For the magnitude of these vectors, linear least squares
561   regression analyses were performed as described previously for
# Line 568 | Line 570 | between those computed from the particular method and
570  
571   The {\it directionality} of the force and torque vectors was
572   investigated through measurement of the angle ($\theta$) formed
573 < between those computed from the particular method and those from SPME,
573 > between those computed from the particular method and those from {\sc spme},
574   \begin{equation}
575   \theta_f = \cos^{-1} \left(\hat{F}_\textrm{SPME} \cdot \hat{F}_\textrm{M}\right),
576   \end{equation}
577 < where $\hat{f}_\textrm{M}$ is the unit vector pointing along the force
577 > where $\hat{F}_\textrm{M}$ is the unit vector pointing along the force
578   vector computed using method M.  Each of these $\theta$ values was
579   accumulated in a distribution function and weighted by the area on the
580   unit sphere.  Since this distribution is a measure of angular error
# Line 601 | Line 603 | calculated when using the reference method (SPME).
603   The variance ($\sigma^2$) was extracted from each of these fits and
604   was used to compare distribution widths.  Values of $\sigma^2$ near
605   zero indicate vector directions indistinguishable from those
606 < calculated when using the reference method (SPME).
606 > calculated when using the reference method ({\sc spme}).
607  
608   \subsection{Short-time Dynamics}
609  
# Line 616 | Line 618 | C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\lan
618   autocorrelation functions (Eq. \ref{eq:vCorr}) were computed for each
619   of the trajectories,
620   \begin{equation}
621 < C_v(t) = \frac{\langle v_i(0)\cdot v_i(t)\rangle}{\langle v_i(0)\cdot v_i(0)\rangle}.
621 > C_v(t) = \frac{\langle v(0)\cdot v(t)\rangle}{\langle v^2\rangle}.
622   \label{eq:vCorr}
623   \end{equation}
624   Velocity autocorrelation functions require detailed short time data,
# Line 644 | Line 646 | reproducing the energetics and dynamics exhibited by S
646   \subsection{Representative Simulations}\label{sec:RepSims}
647   A variety of representative simulations were analyzed to determine the
648   relative effectiveness of the pairwise summation techniques in
649 < reproducing the energetics and dynamics exhibited by SPME.  We wanted
649 > reproducing the energetics and dynamics exhibited by {\sc spme}.  We wanted
650   to span the space of modern simulations (i.e. from liquids of neutral
651   molecules to ionic crystals), so the systems studied were:
652   \begin{enumerate}
# Line 695 | Line 697 | from the reference method (SPME):
697  
698   \subsection{Comparison of Summation Methods}\label{sec:ESMethods}
699   We compared the following alternative summation methods with results
700 < from the reference method (SPME):
700 > from the reference method ({\sc spme}):
701   \begin{itemize}
702   \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
703   and 0.3 \AA$^{-1}$,
# Line 706 | Line 708 | investigated the use of these cutoffs with the SP, SF,
708   \end{itemize}
709   Group-based cutoffs with a fifth-order polynomial switching function
710   were utilized for the reaction field simulations.  Additionally, we
711 < investigated the use of these cutoffs with the SP, SF, and pure
712 < cutoff.  The SPME electrostatics were performed using the TINKER
713 < implementation of SPME,\cite{Ponder87} while all other calculations
711 > investigated the use of these cutoffs with the {\sc sp}, {\sc sf}, and pure
712 > cutoff.  The {\sc spme} electrostatics were performed using the {\sc tinker}
713 > implementation of {\sc spme},\cite{Ponder87} while all other calculations
714   were performed using the {\sc oopse} molecular mechanics
715   package.\cite{Meineke05} All other portions of the energy calculation
716   (i.e. Lennard-Jones interactions) were handled in exactly the same
# Line 723 | Line 725 | The default TINKER tolerance of $1 \times 10^{-8}$ kca
725   10^{-4}$ kcal/mol).  Smaller tolerances are typically associated with
726   increasing accuracy at the expense of computational time spent on the
727   reciprocal-space portion of the summation.\cite{Perram88,Essmann95}
728 < The default TINKER tolerance of $1 \times 10^{-8}$ kcal/mol was used
729 < in all SPME calculations, resulting in Ewald coefficients of 0.4200,
728 > The default {\sc tinker} tolerance of $1 \times 10^{-8}$ kcal/mol was used
729 > in all {\sc spme} calculations, resulting in Ewald coefficients of 0.4200,
730   0.3119, and 0.2476 \AA$^{-1}$ for cutoff radii of 9, 12, and 15 \AA\
731   respectively.
732  
# Line 734 | Line 736 | SPME.  The results for the subsequent regression analy
736   In order to evaluate the performance of the pairwise electrostatic
737   summation methods for Monte Carlo simulations, the energy differences
738   between configurations were compared to the values obtained when using
739 < SPME.  The results for the subsequent regression analysis are shown in
739 > {\sc spme}.  The results for the subsequent regression analysis are shown in
740   figure \ref{fig:delE}.
741  
742   \begin{figure}
# Line 744 | Line 746 | SPME.  Different values of the cutoff radius are indic
746   differences for a given electrostatic method compared with the
747   reference Ewald sum.  Results with a value equal to 1 (dashed line)
748   indicate $\Delta E$ values indistinguishable from those obtained using
749 < SPME.  Different values of the cutoff radius are indicated with
749 > {\sc spme}.  Different values of the cutoff radius are indicated with
750   different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
751   inverted triangles).}
752   \label{fig:delE}
# Line 770 | Line 772 | shows an excellent correlation and quality of fit with
772  
773   For the {\sc sp} method, inclusion of electrostatic damping improves
774   the agreement with Ewald, and using an $\alpha$ of 0.2 \AA $^{-1}$
775 < shows an excellent correlation and quality of fit with the SPME
775 > shows an excellent correlation and quality of fit with the {\sc spme}
776   results, particularly with a cutoff radius greater than 12
777   \AA .  Use of a larger damping parameter is more helpful for the
778   shortest cutoff shown, but it has a detrimental effect on simulations
# Line 804 | Line 806 | using SPME.  Different values of the cutoff radius are
806   magnitudes for a given electrostatic method compared with the
807   reference Ewald sum.  Results with a value equal to 1 (dashed line)
808   indicate force magnitude values indistinguishable from those obtained
809 < using SPME.  Different values of the cutoff radius are indicated with
809 > using {\sc spme}.  Different values of the cutoff radius are indicated with
810   different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
811   inverted triangles).}
812   \label{fig:frcMag}
813   \end{figure}
814  
815   Again, it is striking how well the Shifted Potential and Shifted Force
816 < methods are doing at reproducing the SPME forces.  The undamped and
816 > methods are doing at reproducing the {\sc spme} forces.  The undamped and
817   weakly-damped {\sc sf} method gives the best agreement with Ewald.
818   This is perhaps expected because this method explicitly incorporates a
819   smooth transition in the forces at the cutoff radius as well as the
# Line 843 | Line 845 | using SPME.  Different values of the cutoff radius are
845   magnitudes for a given electrostatic method compared with the
846   reference Ewald sum.  Results with a value equal to 1 (dashed line)
847   indicate torque magnitude values indistinguishable from those obtained
848 < using SPME.  Different values of the cutoff radius are indicated with
848 > using {\sc spme}.  Different values of the cutoff radius are indicated with
849   different symbols (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ =
850   inverted triangles).}
851   \label{fig:trqMag}
# Line 855 | Line 857 | methods with SPME is how much the agreement improves w
857  
858   Torques appear to be much more sensitive to charges at a longer
859   distance.   The striking feature in comparing the new electrostatic
860 < methods with SPME is how much the agreement improves with increasing
860 > methods with {\sc spme} is how much the agreement improves with increasing
861   cutoff radius.  Again, the weakly damped and undamped {\sc sf} method
862 < appears to be reproducing the SPME torques most accurately.  
862 > appears to be reproducing the {\sc spme} torques most accurately.  
863  
864   Water molecules are dipolar, and the reaction field method reproduces
865   the effect of the surrounding polarized medium on each of the
# Line 872 | Line 874 | SPME.  The results (Fig. \ref{fig:frcTrqAng}) are comp
874   vital in calculating dynamical quantities accurately.  Force and
875   torque directionalities were investigated by measuring the angles
876   formed between these vectors and the same vectors calculated using
877 < SPME.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the
877 > {\sc spme}.  The results (Fig. \ref{fig:frcTrqAng}) are compared through the
878   variance ($\sigma^2$) of the Gaussian fits of the angle error
879   distributions of the combined set over all system types.
880  
# Line 884 | Line 886 | obtained using SPME.  Different values of the cutoff r
886   make with their counterparts obtained using the reference Ewald sum.
887   Results with a variance ($\sigma^2$) equal to zero (dashed line)
888   indicate force and torque directions indistinguishable from those
889 < obtained using SPME.  Different values of the cutoff radius are
889 > obtained using {\sc spme}.  Different values of the cutoff radius are
890   indicated with different symbols (9\AA\ = circles, 12\AA\ = squares,
891   and 15\AA\ = inverted triangles).}
892   \label{fig:frcTrqAng}
# Line 893 | Line 895 | sp}) method would be essentially unusable for molecula
895   Both the force and torque $\sigma^2$ results from the analysis of the
896   total accumulated system data are tabulated in figure
897   \ref{fig:frcTrqAng}. Here it is clear that the Shifted Potential ({\sc
898 < sp}) method would be essentially unusable for molecular dynamics until
899 < the damping function is added.  The Shifted Force ({\sc sf}) method,
900 < however, is generating force and torque vectors which are within a few
901 < degrees of the Ewald results even with weak (or no) damping.
898 > sp}) method would be essentially unusable for molecular dynamics
899 > unless the damping function is added.  The Shifted Force ({\sc sf})
900 > method, however, is generating force and torque vectors which are
901 > within a few degrees of the Ewald results even with weak (or no)
902 > damping.
903  
904   All of the sets (aside from the over-damped case) show the improvement
905   afforded by choosing a larger cutoff radius.  Increasing the cutoff
906   from 9 to 12 \AA\ typically results in a halving of the width of the
907 < distribution, with a similar improvement going from 12 to 15
907 > distribution, with a similar improvement when going from 12 to 15
908   \AA .
909  
910   The undamped {\sc sf}, group-based cutoff, and reaction field methods
911   all do equivalently well at capturing the direction of both the force
912 < and torque vectors.  Using damping improves the angular behavior
913 < significantly for the {\sc sp} and moderately for the {\sc sf}
914 < methods.  Overdamping is detrimental to both methods.  Again it is
915 < important to recognize that the force vectors cover all particles in
916 < the systems, while torque vectors are only available for neutral
917 < molecular groups.  Damping appears to have a more beneficial effect on
912 > and torque vectors.  Using the electrostatic damping improves the
913 > angular behavior significantly for the {\sc sp} and moderately for the
914 > {\sc sf} methods.  Overdamping is detrimental to both methods.  Again
915 > it is important to recognize that the force vectors cover all
916 > particles in all seven systems, while torque vectors are only
917 > available for neutral molecular groups.  Damping is more beneficial to
918   charged bodies, and this observation is investigated further in the
919   accompanying supporting information.
920  
921   Although not discussed previously, group based cutoffs can be applied
922 < to both the {\sc sp} and {\sc sf} methods.  Use of a switching
923 < function corrects for the discontinuities that arise when atoms of the
924 < two groups exit the cutoff radius before the group centers leave each
925 < other's cutoff. Though there are no significant benefits or drawbacks
926 < observed in $\Delta E$ and vector magnitude results when doing this,
927 < there is a measurable improvement in the vector angle results.  Table
928 < \ref{tab:groupAngle} shows the angular variance values obtained using
929 < group based cutoffs and a switching function alongside the results
930 < seen in figure \ref{fig:frcTrqAng}.  The {\sc sp} shows much narrower
931 < angular distributions for both the force and torque vectors when using
932 < an $\alpha$ of 0.2 \AA$^{-1}$ or less, while {\sc sf} shows
933 < improvements in the undamped and lightly damped cases.  Thus, by
931 < calculating the electrostatic interactions in terms of molecular pairs
932 < rather than atomic pairs, the direction of the force and torque
933 < vectors can be determined more accurately.
922 > to both the {\sc sp} and {\sc sf} methods.  The group-based cutoffs
923 > will reintroduce small discontinuities at the cutoff radius, but the
924 > effects of these can be minimized by utilizing a switching function.
925 > Though there are no significant benefits or drawbacks observed in
926 > $\Delta E$ and the force and torque magnitudes when doing this, there
927 > is a measurable improvement in the directionality of the forces and
928 > torques. Table \ref{tab:groupAngle} shows the angular variances
929 > obtained using group based cutoffs along with the results seen in
930 > figure \ref{fig:frcTrqAng}.  The {\sc sp} (with an $\alpha$ of 0.2
931 > \AA$^{-1}$ or smaller) shows much narrower angular distributions when
932 > using group-based cutoffs. The {\sc sf} method likewise shows
933 > improvement in the undamped and lightly damped cases.
934  
935   \begin{table}[htbp]
936 <   \centering
937 <   \caption{Variance ($\sigma^2$) of the force (top set) and torque
938 < (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$^2$.}      
936 >   \centering
937 >   \caption{Statistical analysis of the angular
938 >   distributions that the force (upper) and torque (lower) vectors
939 >   from a given electrostatic method make with their counterparts
940 >   obtained using the reference Ewald sum.  Calculations were
941 >   performed both with (Y) and without (N) group based cutoffs and a
942 >   switching function.  The $\alpha$ values have units of \AA$^{-1}$
943 >   and the variance values have units of degrees$^2$.}
944 >
945     \begin{tabular}{@{} ccrrrrrrrr @{}}
946        \\
947        \toprule
# Line 966 | Line 972 | One additional trend to recognize in table \ref{tab:gr
972     \label{tab:groupAngle}
973   \end{table}
974  
975 < One additional trend to recognize in table \ref{tab:groupAngle} is
976 < that the $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as
977 < $\alpha$ increases, something that is easier to see when using group
978 < based cutoffs.  The reason for this is that the complimentary error
979 < function inserted into the potential weakens the electrostatic
980 < interaction as $\alpha$ increases.  Thus, at larger values of
981 < $\alpha$, both summation methods progress toward non-interacting
976 < functions, so care is required in choosing large damping functions
977 < lest one generate an undesirable loss in the pair interaction.  Kast
975 > One additional trend in table \ref{tab:groupAngle} is that the
976 > $\sigma^2$ values for both {\sc sp} and {\sc sf} converge as $\alpha$
977 > increases, something that is more obvious with group-based cutoffs.
978 > The complimentary error function inserted into the potential weakens
979 > the electrostatic interaction as the value of $\alpha$ is increased.
980 > However, at larger values of $\alpha$, it is possible to overdamp the
981 > electrostatic interaction and to remove it completely.  Kast
982   \textit{et al.}  developed a method for choosing appropriate $\alpha$
983   values for these types of electrostatic summation methods by fitting
984   to $g(r)$ data, and their methods indicate optimal values of 0.34,
# Line 982 | Line 986 | molecular torques, particularly for the shorter cutoff
986   respectively.\cite{Kast03} These appear to be reasonable choices to
987   obtain proper MC behavior (Fig. \ref{fig:delE}); however, based on
988   these findings, choices this high would introduce error in the
989 < molecular torques, particularly for the shorter cutoffs.  Based on the
990 < above findings, empirical damping up to 0.2 \AA$^{-1}$ proves to be
991 < beneficial, but damping may be unnecessary when using the {\sc sf}
988 < method.
989 > molecular torques, particularly for the shorter cutoffs.  Based on our
990 > observations, empirical damping up to 0.2 \AA$^{-1}$ is beneficial,
991 > but damping may be unnecessary when using the {\sc sf} method.
992  
993   \subsection{Short-Time Dynamics: Velocity Autocorrelation Functions of NaCl Crystals}
994  
# Line 998 | Line 1001 | low-frequency behavior in crystalline systems.
1001   distribution functions, diffusion constants, and velocity and
1002   orientational correlation functions) may not be particularly sensitive
1003   to the long-range and collective behavior that governs the
1004 < low-frequency behavior in crystalline systems.
1004 > low-frequency behavior in crystalline systems.  Additionally, the
1005 > ionic crystals are the worst case scenario for the pairwise methods
1006 > because they lack the reciprocal space contribution contained in the
1007 > Ewald summation.  
1008  
1009   We are using two separate measures to probe the effects of these
1010   alternative electrostatic methods on the dynamics in crystalline
# Line 1010 | Line 1016 | low-frequency portion of the power spectrum.
1016   \begin{figure}
1017   \centering
1018   \includegraphics[width = \linewidth]{./vCorrPlot.pdf}
1019 < \caption{Velocity auto-correlation functions of NaCl crystals at
1020 < 1000 K using SPME, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1019 > \caption{Velocity autocorrelation functions of NaCl crystals at
1020 > 1000 K using {\sc spme}, {\sc sf} ($\alpha$ = 0.0, 0.1, \& 0.2), and {\sc
1021   sp} ($\alpha$ = 0.2). The inset is a magnification of the area around
1022   the first minimum.  The times to first collision are nearly identical,
1023   but differences can be seen in the peaks and troughs, where the
1024   undamped and weakly damped methods are stiffer than the moderately
1025 < damped and SPME methods.}
1025 > damped and {\sc spme} methods.}
1026   \label{fig:vCorrPlot}
1027   \end{figure}
1028  
1029 < The short-time decays through the first collision are nearly identical
1030 < in figure \ref{fig:vCorrPlot}, but the peaks and troughs of the
1031 < functions show how the methods differ.  The undamped {\sc sf} method
1032 < has deeper troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher
1033 < peaks than any of the other methods.  As the damping function is
1034 < increased, these peaks are smoothed out, and approach the SPME
1035 < curve. The damping acts as a distance dependent Gaussian screening of
1036 < the point charges for the pairwise summation methods; thus, the
1037 < collisions are more elastic in the undamped {\sc sf} potential, and the
1038 < stiffness of the potential is diminished as the electrostatic
1039 < interactions are softened by the damping function.  With $\alpha$
1040 < values of 0.2 \AA$^{-1}$, the {\sc sf} and {\sc sp} functions are
1041 < nearly identical and track the SPME features quite well.  This is not
1042 < too surprising in that the differences between the {\sc sf} and {\sc
1043 < sp} potentials are mitigated with increased damping.  However, this
1038 < appears to indicate that once damping is utilized, the form of the
1039 < potential seems to play a lesser role in the crystal dynamics.
1029 > The short-time decay of the velocity autocorrelation function through
1030 > the first collision are nearly identical in figure
1031 > \ref{fig:vCorrPlot}, but the peaks and troughs of the functions show
1032 > how the methods differ.  The undamped {\sc sf} method has deeper
1033 > troughs (see inset in Fig. \ref{fig:vCorrPlot}) and higher peaks than
1034 > any of the other methods.  As the damping parameter ($\alpha$) is
1035 > increased, these peaks are smoothed out, and the {\sc sf} method
1036 > approaches the {\sc spme} results.  With $\alpha$ values of 0.2 \AA$^{-1}$,
1037 > the {\sc sf} and {\sc sp} functions are nearly identical and track the
1038 > {\sc spme} features quite well.  This is not surprising because the {\sc sf}
1039 > and {\sc sp} potentials become nearly identical with increased
1040 > damping.  However, this appears to indicate that once damping is
1041 > utilized, the details of the form of the potential (and forces)
1042 > constructed out of the damped electrostatic interaction are less
1043 > important.
1044  
1045   \subsection{Collective Motion: Power Spectra of NaCl Crystals}
1046  
1047 < The short time dynamics were extended to evaluate how the differences
1048 < between the methods affect the collective long-time motion.  The same
1049 < electrostatic summation methods were used as in the short time
1050 < velocity autocorrelation function evaluation, but the trajectories
1051 < were sampled over a much longer time. The power spectra of the
1052 < resulting velocity autocorrelation functions were calculated and are
1053 < displayed in figure \ref{fig:methodPS}.
1047 > To evaluate how the differences between the methods affect the
1048 > collective long-time motion, we computed power spectra from long-time
1049 > traces of the velocity autocorrelation function. The power spectra for
1050 > the best-performing alternative methods are shown in
1051 > fig. \ref{fig:methodPS}.  Apodization of the correlation functions via
1052 > a cubic switching function between 40 and 50 ps was used to reduce the
1053 > ringing resulting from data truncation.  This procedure had no
1054 > noticeable effect on peak location or magnitude.
1055  
1056   \begin{figure}
1057   \centering
1058   \includegraphics[width = \linewidth]{./spectraSquare.pdf}
1059   \caption{Power spectra obtained from the velocity auto-correlation
1060 < functions of NaCl crystals at 1000 K while using SPME, {\sc sf}
1061 < ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).
1062 < Apodization of the correlation functions via a cubic switching
1063 < function between 40 and 50 ps was used to clear up the spectral noise
1059 < resulting from data truncation, and had no noticeable effect on peak
1060 < location or magnitude.  The inset shows the frequency region below 100
1061 < cm$^{-1}$ to highlight where the spectra begin to differ.}
1060 > functions of NaCl crystals at 1000 K while using {\sc spme}, {\sc sf}
1061 > ($\alpha$ = 0, 0.1, \& 0.2), and {\sc sp} ($\alpha$ = 0.2).  The inset
1062 > shows the frequency region below 100 cm$^{-1}$ to highlight where the
1063 > spectra differ.}
1064   \label{fig:methodPS}
1065   \end{figure}
1066  
1067 < While high frequency peaks of the spectra in this figure overlap,
1068 < showing the same general features, the low frequency region shows how
1069 < the summation methods differ.  Considering the low-frequency inset
1070 < (expanded in the upper frame of figure \ref{fig:dampInc}), at
1071 < frequencies below 100 cm$^{-1}$, the correlated motions are
1072 < blue-shifted when using undamped or weakly damped {\sc sf}.  When
1073 < using moderate damping ($\alpha = 0.2$ \AA$^{-1}$) both the {\sc sf}
1074 < and {\sc sp} methods give near identical correlated motion behavior as
1075 < the Ewald method (which has a damping value of 0.3119).  This
1076 < weakening of the electrostatic interaction with increased damping
1077 < explains why the long-ranged correlated motions are at lower
1078 < frequencies for the moderately damped methods than for undamped or
1079 < weakly damped methods.  To see this effect more clearly, we show how
1080 < damping strength alone affects a simple real-space electrostatic
1081 < potential,
1082 < \begin{equation}
1083 < V_\textrm{damped}=\sum^N_i\sum^N_{j\ne i}q_iq_j\left[\frac{\textrm{erfc}({\alpha r})}{r}\right]S(r),
1084 < \end{equation}
1085 < where $S(r)$ is a switching function that smoothly zeroes the
1086 < potential at the cutoff radius.  Figure \ref{fig:dampInc} shows how
1087 < the low frequency motions are dependent on the damping used in the
1088 < direct electrostatic sum.  As the damping increases, the peaks drop to
1089 < lower frequencies.  Incidentally, use of an $\alpha$ of 0.25
1090 < \AA$^{-1}$ on a simple electrostatic summation results in low
1091 < frequency correlated dynamics equivalent to a simulation using SPME.
1092 < When the coefficient lowers to 0.15 \AA$^{-1}$ and below, these peaks
1093 < shift to higher frequency in exponential fashion.  Though not shown,
1094 < the spectrum for the simple undamped electrostatic potential is
1093 < blue-shifted such that the lowest frequency peak resides near 325
1094 < cm$^{-1}$.  In light of these results, the undamped {\sc sf} method
1095 < producing low-lying motion peaks within 10 cm$^{-1}$ of SPME is quite
1096 < respectable and shows that the shifted force procedure accounts for
1097 < most of the effect afforded through use of the Ewald summation.
1098 < However, it appears as though moderate damping is required for
1099 < accurate reproduction of crystal dynamics.
1067 > While the high frequency regions of the power spectra for the
1068 > alternative methods are quantitatively identical with Ewald spectrum,
1069 > the low frequency region shows how the summation methods differ.
1070 > Considering the low-frequency inset (expanded in the upper frame of
1071 > figure \ref{fig:dampInc}), at frequencies below 100 cm$^{-1}$, the
1072 > correlated motions are blue-shifted when using undamped or weakly
1073 > damped {\sc sf}.  When using moderate damping ($\alpha = 0.2$
1074 > \AA$^{-1}$) both the {\sc sf} and {\sc sp} methods give nearly identical
1075 > correlated motion to the Ewald method (which has a convergence
1076 > parameter of 0.3119 \AA$^{-1}$).  This weakening of the electrostatic
1077 > interaction with increased damping explains why the long-ranged
1078 > correlated motions are at lower frequencies for the moderately damped
1079 > methods than for undamped or weakly damped methods.
1080 >
1081 > To isolate the role of the damping constant, we have computed the
1082 > spectra for a single method ({\sc sf}) with a range of damping
1083 > constants and compared this with the {\sc spme} spectrum.
1084 > Fig. \ref{fig:dampInc} shows more clearly that increasing the
1085 > electrostatic damping red-shifts the lowest frequency phonon modes.
1086 > However, even without any electrostatic damping, the {\sc sf} method
1087 > has at most a 10 cm$^{-1}$ error in the lowest frequency phonon mode.
1088 > Without the {\sc sf} modifications, an undamped (pure cutoff) method
1089 > would predict the lowest frequency peak near 325 cm$^{-1}$.  {\it
1090 > Most} of the collective behavior in the crystal is accurately captured
1091 > using the {\sc sf} method.  Quantitative agreement with Ewald can be
1092 > obtained using moderate damping in addition to the shifting at the
1093 > cutoff distance.
1094 >
1095   \begin{figure}
1096   \centering
1097   \includegraphics[width = \linewidth]{./comboSquare.pdf}
1098 < \caption{Regions of spectra showing the low-frequency correlated
1099 < motions for NaCl crystals at 1000 K using various electrostatic
1100 < summation methods.  The upper plot is a zoomed inset from figure
1101 < \ref{fig:methodPS}.  As the damping value for the {\sc sf} potential
1102 < increases, the low-frequency peaks red-shift.  The lower plot is of
1103 < spectra when using SPME and a simple damped Coulombic sum with damping
1104 < coefficients ($\alpha$) ranging from 0.15 to 0.3 \AA$^{-1}$.  As
1110 < $\alpha$ increases, the peaks are red-shifted toward and eventually
1111 < beyond the values given by SPME.  The larger $\alpha$ values weaken
1112 < the real-space electrostatics, explaining this shift towards less
1113 < strongly correlated motions in the crystal.}
1098 > \caption{Effect of damping on the two lowest-frequency phonon modes in
1099 > the NaCl crystal at 1000K.  The undamped shifted force ({\sc sf})
1100 > method is off by less than 10 cm$^{-1}$, and increasing the
1101 > electrostatic damping to 0.25 \AA$^{-1}$ gives quantitative agreement
1102 > with the power spectrum obtained using the Ewald sum.  Overdamping can
1103 > result in underestimates of frequencies of the long-wavelength
1104 > motions.}
1105   \label{fig:dampInc}
1106   \end{figure}
1107  
1108   \section{Conclusions}
1109  
1110   This investigation of pairwise electrostatic summation techniques
1111 < shows that there are viable and more computationally efficient
1112 < electrostatic summation techniques than the Ewald summation, chiefly
1113 < methods derived from the damped Coulombic sum originally proposed by
1114 < Wolf \textit{et al.}\cite{Wolf99,Zahn02} In particular, the
1115 < {\sc sf} method, reformulated above as eq. (\ref{eq:DSFPot}),
1116 < shows a remarkable ability to reproduce the energetic and dynamic
1117 < characteristics exhibited by simulations employing lattice summation
1118 < techniques.  The cumulative energy difference results showed the
1119 < undamped {\sc sf} and moderately damped {\sc sp} methods
1120 < produced results nearly identical to SPME.  Similarly for the dynamic
1121 < features, the undamped or moderately damped {\sc sf} and
1122 < moderately damped {\sc sp} methods produce force and torque
1123 < vector magnitude and directions very similar to the expected values.
1124 < These results translate into long-time dynamic behavior equivalent to
1125 < that produced in simulations using SPME.
1111 > shows that there are viable and computationally efficient alternatives
1112 > to the Ewald summation.  These methods are derived from the damped and
1113 > cutoff-neutralized Coulombic sum originally proposed by Wolf
1114 > \textit{et al.}\cite{Wolf99} In particular, the {\sc sf}
1115 > method, reformulated above as eqs. (\ref{eq:DSFPot}) and
1116 > (\ref{eq:DSFForces}), shows a remarkable ability to reproduce the
1117 > energetic and dynamic characteristics exhibited by simulations
1118 > employing lattice summation techniques.  The cumulative energy
1119 > difference results showed the undamped {\sc sf} and moderately damped
1120 > {\sc sp} methods produced results nearly identical to {\sc spme}.  Similarly
1121 > for the dynamic features, the undamped or moderately damped {\sc sf}
1122 > and moderately damped {\sc sp} methods produce force and torque vector
1123 > magnitude and directions very similar to the expected values.  These
1124 > results translate into long-time dynamic behavior equivalent to that
1125 > produced in simulations using {\sc spme}.
1126  
1127 + As in all purely-pairwise cutoff methods, these methods are expected
1128 + to scale approximately {\it linearly} with system size, and they are
1129 + easily parallelizable.  This should result in substantial reductions
1130 + in the computational cost of performing large simulations.
1131 +
1132   Aside from the computational cost benefit, these techniques have
1133   applicability in situations where the use of the Ewald sum can prove
1134 < problematic.  Primary among them is their use in interfacial systems,
1135 < where the unmodified lattice sum techniques artificially accentuate
1136 < the periodicity of the system in an undesirable manner.  There have
1137 < been alterations to the standard Ewald techniques, via corrections and
1138 < reformulations, to compensate for these systems; but the pairwise
1139 < techniques discussed here require no modifications, making them
1140 < natural tools to tackle these problems.  Additionally, this
1141 < transferability gives them benefits over other pairwise methods, like
1142 < reaction field, because estimations of physical properties (e.g. the
1143 < dielectric constant) are unnecessary.
1134 > problematic.  Of greatest interest is their potential use in
1135 > interfacial systems, where the unmodified lattice sum techniques
1136 > artificially accentuate the periodicity of the system in an
1137 > undesirable manner.  There have been alterations to the standard Ewald
1138 > techniques, via corrections and reformulations, to compensate for
1139 > these systems; but the pairwise techniques discussed here require no
1140 > modifications, making them natural tools to tackle these problems.
1141 > Additionally, this transferability gives them benefits over other
1142 > pairwise methods, like reaction field, because estimations of physical
1143 > properties (e.g. the dielectric constant) are unnecessary.
1144  
1145 < We are not suggesting any flaw with the Ewald sum; in fact, it is the
1146 < standard by which these simple pairwise sums are judged.  However,
1147 < these results do suggest that in the typical simulations performed
1148 < today, the Ewald summation may no longer be required to obtain the
1149 < level of accuracy most researchers have come to expect
1145 > If a researcher is using Monte Carlo simulations of large chemical
1146 > systems containing point charges, most structural features will be
1147 > accurately captured using the undamped {\sc sf} method or the {\sc sp}
1148 > method with an electrostatic damping of 0.2 \AA$^{-1}$.  These methods
1149 > would also be appropriate for molecular dynamics simulations where the
1150 > data of interest is either structural or short-time dynamical
1151 > quantities.  For long-time dynamics and collective motions, the safest
1152 > pairwise method we have evaluated is the {\sc sf} method with an
1153 > electrostatic damping between 0.2 and 0.25
1154 > \AA$^{-1}$.
1155  
1156 + We are not suggesting that there is any flaw with the Ewald sum; in
1157 + fact, it is the standard by which these simple pairwise sums have been
1158 + judged.  However, these results do suggest that in the typical
1159 + simulations performed today, the Ewald summation may no longer be
1160 + required to obtain the level of accuracy most researchers have come to
1161 + expect.
1162 +
1163   \section{Acknowledgments}
1164 + Support for this project was provided by the National Science
1165 + Foundation under grant CHE-0134881.  The authors would like to thank
1166 + Steve Corcelli and Ed Maginn for helpful discussions and comments.
1167 +
1168   \newpage
1169  
1170   \bibliographystyle{jcp2}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines