ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
(Generate patch)

Comparing trunk/multipole/multipole_2/multipole2.tex (file contents):
Revision 4174 by gezelter, Thu Jun 5 19:55:14 2014 UTC vs.
Revision 4175 by gezelter, Fri Jun 6 16:01:36 2014 UTC

# Line 36 | Line 36 | preprint,
36   \usepackage{amsmath}
37   \usepackage{times}
38   \usepackage{mathptm}
39 + \usepackage{tabularx}
40   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
41   \usepackage{url}
42   \usepackage[english]{babel}
43  
44 + \newcolumntype{Y}{>{\centering\arraybackslash}X}
45  
46   \begin{document}
47  
48 < \preprint{AIP/123-QED}
48 > %\preprint{AIP/123-QED}
49  
50 < \title[Efficient electrostatics for condensed-phase multipoles]{Real space alternatives to the Ewald
51 < Sum. II. Comparison of Simulation Methodologies} % Force line breaks with \\
50 > \title{Real space alternatives to the Ewald
51 > Sum. II. Comparison of Methods} % Force line breaks with \\
52  
53   \author{Madan Lamichhane}
54   \affiliation{Department of Physics, University
# Line 65 | Line 67 | We have tested our recently developed shifted potentia
67               %  but any date may be explicitly specified
68  
69   \begin{abstract}
70 < We have tested our recently developed shifted potential, gradient-shifted force, and Taylor-shifted force methods for the higher-order multipoles against Ewald’s method in different types of liquid and crystalline system. In this paper, we have also investigated the conservation of total energy in the molecular dynamic simulation using all of these methods. The shifted potential method shows better agreement with the Ewald in the energy differences between different configurations as compared to the direct truncation. Both the gradient shifted force and Taylor-shifted force methods reproduce very good energy conservation. But the absolute energy, force and torque evaluated from the gradient shifted force method shows better result as compared to taylor-shifted force method. Hence the gradient-shifted force method suitably mimics the electrostatic interaction in the molecular dynamic simulation.
70 >  We have tested the real-space shifted potential (SP),
71 >  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
72 >  for multipoles that were developed in the first paper in this series
73 >  against a reference method. The tests were carried out in a variety
74 >  of condensed-phase environments which were designed to test all
75 >  levels of the multipole-multipole interactions.  Comparisons of the
76 >  energy differences between configurations, molecular forces, and
77 >  torques were used to analyze how well the real-space models perform
78 >  relative to the more computationally expensive Ewald sum.  We have
79 >  also investigated the energy conservation properties of the new
80 >  methods in molecular dynamics simulations using all of these
81 >  methods. The SP method shows excellent agreement with
82 >  configurational energy differences, forces, and torques, and would
83 >  be suitable for use in Monte Carlo calculations.  Of the two new
84 >  shifted-force methods, the GSF approach shows the best agreement
85 >  with Ewald-derived energies, forces, and torques and exhibits energy
86 >  conservation properties that make it an excellent choice for
87 >  efficiently computing electrostatic interactions in molecular
88 >  dynamics simulations.
89   \end{abstract}
90  
91 < \pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
91 > %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92                               % Classification Scheme.
93   \keywords{Electrostatics, Multipoles, Real-space}
94  
# Line 315 | Line 335 | f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^
335   number of terms in the truncated Taylor expansion, e.g.,
336   %
337   \begin{equation}
338 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
338 > f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
339   \end{equation}
340   %
341   The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
342   Thus, for $f(r)=1/r$, we find
343   %
344   \begin{equation}
345 < f_1(r)=\frac{1}{r}- \frac{1}{R_c} + (r - R_c) \frac{1}{R_c^2} - \frac{(r-R_c)^2}{R_c^3} .
345 > f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
346   \end{equation}
347   This function is an approximate electrostatic potential that has
348   vanishing second derivatives at the cutoff radius, making it suitable
# Line 349 | Line 369 | $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives
369  
370   Note that increasing the value of $n$ will add additional terms to the
371   electrostatic potential, e.g., $f_2(r)$ includes orders up to
372 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
372 > $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
373   functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
374   f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
375   for computing multipole energies, forces, and torques, and smooth
# Line 377 | Line 397 | presented in the first paper in this series (Reference
397   Eq.~\ref{generic}, accounting for the appropriate number of
398   derivatives. Complete energy, force, and torque expressions are
399   presented in the first paper in this series (Reference
400 < \citep{PaperI}).
400 > \onlinecite{PaperI}).
401  
402   \subsection{Gradient-shifted force (GSF)}
403  
# Line 388 | Line 408 | U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{
408   which have been projected onto the surface of the cutoff sphere
409   without changing their relative orientation,
410   \begin{displaymath}
411 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
412 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
413 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
411 > U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(r_c)
412 >   - (r_{ij}-r_c) \hat{r}_{ij} \cdot
413 >  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{r_c}
414   \end{displaymath}
415   Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
416   and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
# Line 419 | Line 439 | approach zero as $r \rightarrow R_c$.  Both the GSF an
439   function of radius, hence the contribution of the third term is very
440   small for large cutoff radii.  The force and torque derived from
441   equation \ref{generic2} are consistent with the energy expression and
442 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
442 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
443   can be considered generalizations of the original DSF method for
444   higher order multipole interactions. GSF and TSF are also identical up
445   to the charge-dipole interaction but generate different expressions in
446   the energy, force and torque for higher order multipole-multipole
447   interactions. Complete energy, force, and torque expressions for the
448   GSF potential are presented in the first paper in this series
449 < (Reference \citep{PaperI})
449 > (Reference~\onlinecite{PaperI})
450  
451  
452   \subsection{Shifted potential (SP) }
# Line 446 | Line 466 | is not a simple shifting of the total potential at $R_
466   each orientational contribution to the energy (e.g. the direct dipole
467   product contribution is shifted {\it separately} from the
468   dipole-distance terms in dipole-dipole interactions).  Note that this
469 < is not a simple shifting of the total potential at $R_c$. Each radial
469 > is not a simple shifting of the total potential at $r_c$. Each radial
470   contribution is shifted separately.  One consequence of this is that
471   multipoles that reorient after leaving the cutoff sphere can re-enter
472   the cutoff sphere without perturbing the total energy.
473  
474   The potential energy between a central multipole and other multipolar
475 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
475 > sites then goes smoothly to zero as $r \rightarrow r_c$. However, the
476   force and torque obtained from the shifted potential (SP) are
477 < discontinuous at $R_c$. Therefore, MD simulations will still
477 > discontinuous at $r_c$. Therefore, MD simulations will still
478   experience energy drift while operating under the SP potential, but it
479   may be suitable for Monte Carlo approaches where the configurational
480   energy differences are the primary quantity of interest.
# Line 465 | Line 485 | this series (Reference \citep{PaperI})  
485   the cutoff sphere.  This self interaction is nearly identical with the
486   self-terms that arise in the Ewald sum for multipoles.  Complete
487   expressions for the self terms are presented in the first paper in
488 < this series (Reference \citep{PaperI})  
488 > this series (Reference \onlinecite{PaperI}).
489  
490  
491   \section{\label{sec:methodology}Methodology}
# Line 480 | Line 500 | disordered and ordered condensed-phase systems.  These
500   arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
501   used the multipolar Ewald sum as a reference method for comparing
502   energies, forces, and torques for molecular models that mimic
503 < disordered and ordered condensed-phase systems.  These test-cases
504 < include:
485 < \begin{itemize}
486 < \item Soft Dipolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $|D| = 2.35$)
487 < \item Soft Dipolar solids ($\sigma = 2.837$, $\epsilon =1.0$, $|D| = 2.35$)
488 < \item Soft Quadrupolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
489 < \item Soft Quadrupolar solids  ($\sigma = 2.837$, $\epsilon = 1.0$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
490 < \item A mixed multipole model (SSDQ) for water ($\sigma = 3.051$, $\epsilon = 0.152$, $D_z = 2.35$, $Q_{\alpha\alpha} =\left\{-1.35,0,-0.68\right\}$)
491 < \item A mixed multipole models for water with 48 dissolved ions, 24
492 <  \ce{Na+}: ($\sigma = 2.579$, $\epsilon =0.118$, $q = 1e$) and 24
493 <  \ce{Cl-}: ($\sigma = 4.445$, $\epsilon =0.1$l, $q = -1e$)
494 < \end{itemize}
495 < All Lennard-Jones parameters are in units of \AA\ $(\sigma)$ and kcal
496 < / mole $(\epsilon)$.  Partial charges are reported in electrons, while
497 < dipoles are in Debye units, and quadrupoles are in units of Debye-\AA.
503 > disordered and ordered condensed-phase systems.  The parameters used
504 > in the test-cases are given in table~\ref{tab:pars}.
505  
506 < The last test case exercises all levels of the multipole-multipole
507 < interactions we have derived so far and represents the most complete
508 < test of the new methods.  In the following section, we present results
509 < for the total electrostatic energy, as well as the electrostatic
510 < contributions to the force and torque on each molecule.  These
511 < quantities have been computed using the SP, TSF, and GSF methods, as
512 < well as a hard cutoff, and have been compared with the values obtaine
513 < from the multipolar Ewald sum.  In Mote Carlo (MC) simulations, the
514 < energy differences between two configurations is the primary quantity
515 < that governs how the simulation proceeds. These differences are the
516 < most imporant indicators of the reliability of a method even if the
517 < absolute energies are not exact.  For each of the multipolar systems
518 < listed above, we have compared the change in electrostatic potential
519 < energy ($\Delta E$) between 250 statistically-independent
520 < configurations.  In molecular dynamics (MD) simulations, the forces
521 < and torques govern the behavior of the simulation, so we also compute
522 < the electrostatic contributions to the forces and torques.
506 > \begin{table}
507 > \label{tab:pars}
508 > \caption{The parameters used in the systems used to evaluate the new
509 >  real-space methods.  The most comprehensive test was a liquid
510 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
511 >  ions).  This test excercises all orders of the multipolar
512 >  interactions developed in the first paper.}
513 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
514 >             & \multicolumn{2}{c|}{LJ parameters} &
515 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
516 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
517 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
518 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
519 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
520 > \AA\textsuperscript{2})} \\ \hline
521 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
522 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & 10,000  & 17.6 &17.6 & 0 \\
523 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
524 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & 10,000  & 17.6&17.6&0 \\
525 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
526 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
527 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
528 > \end{tabularx}
529 > \end{table}
530 > The systems consist of pure multipolar solids (both dipole and
531 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
532 > fluid composed of sites containing both dipoles and quadrupoles
533 > simultaneously, and a final test case that includes ions with point
534 > charges in addition to the multipolar fluid.  The solid-phase
535 > parameters were chosen so that the systems can explore some
536 > orientational freedom for the multipolar sites, while maintaining
537 > relatively strict translational order.  The SSDQ model used here is
538 > not a particularly accurate water model, but it does test
539 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
540 > interactions at roughly the same magnitudes. The last test case, SSDQ
541 > water with dissolved ions, exercises \textit{all} levels of the
542 > multipole-multipole interactions we have derived so far and represents
543 > the most complete test of the new methods.
544  
545 + In the following section, we present results for the total
546 + electrostatic energy, as well as the electrostatic contributions to
547 + the force and torque on each molecule.  These quantities have been
548 + computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
549 + and have been compared with the values obtaine from the multipolar
550 + Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
551 + between two configurations is the primary quantity that governs how
552 + the simulation proceeds. These differences are the most imporant
553 + indicators of the reliability of a method even if the absolute
554 + energies are not exact.  For each of the multipolar systems listed
555 + above, we have compared the change in electrostatic potential energy
556 + ($\Delta E$) between 250 statistically-independent configurations.  In
557 + molecular dynamics (MD) simulations, the forces and torques govern the
558 + behavior of the simulation, so we also compute the electrostatic
559 + contributions to the forces and torques.
560 +
561 + \subsection{Implementation}
562 + The real-space methods developed in the first paper in this series
563 + have been implemented in our group's open source molecular simulation
564 + program, OpenMD,\cite{openmd} which was used for all calculations in
565 + this work.  The complementary error function can be a relatively slow
566 + function on some processors, so all of the radial functions are
567 + precomputed on a fine grid and are spline-interpolated to provide
568 + values when required.  
569 +
570 + Using the same simulation code, we compare to a multipolar Ewald sum
571 + with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
572 + the Ewald sum is a re-implementation of the algorithm originally
573 + proposed by Smith that does not use the particle mesh or smoothing
574 + approximations.\cite{Smith82,Smith98} In all cases, the quantities
575 + being compared are the electrostatic contributions to energies, force,
576 + and torques.  All other contributions to these quantities (i.e. from
577 + Lennard-Jones interactions) are removed prior to the comparisons.
578 +
579 + The convergence parameter ($\alpha$) also plays a role in the balance
580 + of the real-space and reciprocal-space portions of the Ewald
581 + calculation.  Typical molecular mechanics packages set this to a value
582 + that depends on the cutoff radius and a tolerance (typically less than
583 + $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
584 + associated with increasing accuracy at the expense of computational
585 + time spent on the reciprocal-space portion of the
586 + summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
587 + 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
588 + Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
589 +
590 + The real-space models have self-interactions that provide
591 + contributions to the energies only.  Although the self interaction is
592 + a rapid calculation, we note that in systems with fluctuating charges
593 + or point polarizabilities, the self-term is not static and must be
594 + recomputed at each time step.
595 +
596   \subsection{Model systems}
597   To sample independent configurations of multipolar crystals, a body
598   centered cubic (bcc) crystal which is a minimum energy structure for
# Line 699 | Line 778 | reproduce the Ewald-derived configurational energy dif
778  
779   The combined correlation coefficient and slope for all six systems is
780   shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
781 < reproduce the Ewald-derived configurational energy differences with
782 < remarkable fidelity.  Undamped hard cutoffs introduce a significant
783 < amount of random scatter in the energy differences which is apparent
784 < in the reduced value of the correlation coefficient for this method.
785 < This can be understood easily as configurations which exhibit only
786 < small traversals of a few dipoles or quadrupoles out of the cutoff
787 < sphere will see large energy jumps when hard cutoffs are used.  The
781 > reproduce the Ewald configurational energy differences with remarkable
782 > fidelity.  Undamped hard cutoffs introduce a significant amount of
783 > random scatter in the energy differences which is apparent in the
784 > reduced value of the correlation coefficient for this method.  This
785 > can be easily understood as configurations which exhibit small
786 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
787 > will see large energy jumps when hard cutoffs are used.  The
788   orientations of the multipoles (particularly in the ordered crystals)
789 < mean that these jumps can go either up or down in energy, producing a
790 < significant amount of random scatter.
789 > mean that these energy jumps can go in either direction, producing a
790 > significant amount of random scatter, but no systematic error.
791  
792   The TSF method produces energy differences that are highly correlated
793   with the Ewald results, but it also introduces a significant
# Line 716 | Line 795 | effect on crystalline systems.
795   smaller cutoff values. The TSF method alters the distance dependence
796   of different orientational contributions to the energy in a
797   non-uniform way, so the size of the cutoff sphere can have a large
798 < effect on crystalline systems.
798 > effect, particularly for the crystalline systems.
799  
800   Both the SP and GSF methods appear to reproduce the Ewald results with
801   excellent fidelity, particularly for moderate damping ($\alpha =
802 < 0.1-0.2$\AA$^{-1}$) and commonly-used cutoff values ($r_c = 12$\AA).
803 < With the exception of the undamped hard cutoff, and the TSF method
804 < with short cutoffs, all of the methods would be appropriate for use in
805 < Monte Carlo simulations.
802 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
803 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
804 > method with short cutoffs, all of the methods would be appropriate for
805 > use in Monte Carlo simulations.
806  
807   \subsection{Magnitude of the force and torque vectors}
808  
809 < The comparison of the magnitude of the combined forces and torques for
810 < the data accumulated from all system types are shown in Figures
809 > The comparisons of the magnitudes of the forces and torques for the
810 > data accumulated from all six systems are shown in Figures
811   ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
812   correlation and slope for the forces agree well with the Ewald sum
813 < even for the hard cutoff method.
813 > even for the hard cutoffs.
814  
815 < For the system of molecules with higher order multipoles, the
816 < interaction is quite short ranged. Moreover, the force decays more
817 < rapidly than the electrostatic energy hence the hard cutoff method can
818 < also produces reasonable agreement.  Although the pure cutoff gives
819 < the good match of the electrostatic force for pairs of molecules
820 < included within the cutoff sphere, the discontinuity in the force at
821 < the cutoff radius can potentially cause problems the total energy
822 < conservation as molecules enter and leave the cutoff sphere.  This is
823 < discussed in detail in section \ref{sec:}.
815 > For systems of molecules with only multipolar interactions, the pair
816 > energy contributions are quite short ranged.  Moreover, the force
817 > decays more rapidly than the electrostatic energy, hence the hard
818 > cutoff method can also produce reasonable agreement for this quantity.
819 > Although the pure cutoff gives reasonably good electrostatic forces
820 > for pairs of molecules included within each other's cutoff spheres,
821 > the discontinuity in the force at the cutoff radius can potentially
822 > cause energy conservation problems as molecules enter and leave the
823 > cutoff spheres.  This is discussed in detail in section
824 > \ref{sec:conservation}.
825  
826   The two shifted-force methods (GSF and TSF) exhibit a small amount of
827   systematic variation and scatter compared with the Ewald forces.  The
828   shifted-force models intentionally perturb the forces between pairs of
829 < molecules inside the cutoff sphere in order to correct the energy
830 < conservation issues, so it is not particularly surprising that this
831 < perturbation is evident in these same molecular forces.  The GSF
829 > molecules inside each other's cutoff spheres in order to correct the
830 > energy conservation issues, and this perturbation is evident in the
831 > statistics accumulated for the molecular forces.  The GSF
832   perturbations are minimal, particularly for moderate damping and and
833   commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
834   reasonable agreement in the correlation coefficient but again the
# Line 791 | Line 871 | The other real-space methods can cause some significan
871  
872   The results shows that the torque from the hard cutoff method
873   reproduces the torques in quite good agreement with the Ewald sum.
874 < The other real-space methods can cause some significant deviations,
875 < but excellent agreement with the Ewald sum torques is recovered at
876 < moderate values of the damping coefficient ($\alpha =
877 < 0.1-0.2$\AA$^{-1}$) and cutoff radius ($r_c \ge 12$\AA).  The TSF
878 < method exhibits the only fair agreement in the slope as compared to
879 < Ewald even for larger cutoff radii.  It appears that the severity of
880 < the perturbations in the TSF method are most apparent in the torques.
874 > The other real-space methods can cause some deviations, but excellent
875 > agreement with the Ewald sum torques is recovered at moderate values
876 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
877 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
878 > in the slope when compared with the Ewald torques even for larger
879 > cutoff radii.  It appears that the severity of the perturbations in
880 > the TSF method are most in evidence for the torques.
881  
882   \subsection{Directionality of the force and torque vectors}  
883  
# Line 808 | Line 888 | from the new real-space method exhibit nearly-ideal Fi
888   directionality is shown in terms of circular variance
889   ($\mathrm{Var}(\theta$) in figure
890   \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
891 < from the new real-space method exhibit nearly-ideal Fisher probability
891 > from the new real-space methods exhibit nearly-ideal Fisher probability
892   distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
893   exhibit the best vectorial agreement with the Ewald sum. The force and
894   torque vectors from GSF method also show good agreement with the Ewald
895   method, which can also be systematically improved by using moderate
896 < damping and a reasonable cutoff radius.  For $\alpha = 0.2$ and $r_c =
896 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
897   12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
898 < to a distribution with 95\% of force vectors within $6.37^\circ$ of the
899 < corresponding Ewald forces. The TSF method produces the poorest
898 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
899 > the corresponding Ewald forces. The TSF method produces the poorest
900   agreement with the Ewald force directions.
901  
902 < Torques are again more perturbed by the new real-space methods, than
903 < forces, but even here the variance is reasonably small.  For the same
902 > Torques are again more perturbed than the forces by the new real-space
903 > methods, but even here the variance is reasonably small.  For the same
904   method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
905   the circular variance was 0.01415, corresponds to a distribution which
906   has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
# Line 840 | Line 920 | systematically improved by varying $\alpha$ and $r_c$.
920    \label{fig:slopeCorr_circularVariance}
921   \end{figure}
922  
923 < \subsection{Energy conservation}
923 > \subsection{Energy conservation\label{sec:conservation}}
924  
925   We have tested the conservation of energy one can expect to see with
926   the new real-space methods using the SSDQ water model with a small
# Line 848 | Line 928 | and 48 dissolved ions at a density of 0.98 g cm${-3}$
928   orders of multipole-multipole interactions derived in the first paper
929   in this series and provides the most comprehensive test of the new
930   methods.  A liquid-phase system was created with 2000 water molecules
931 < and 48 dissolved ions at a density of 0.98 g cm${-3}$ and a
931 > and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
932   temperature of 300K.  After equilibration, this liquid-phase system
933   was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
934 < a cutoff radius of 9\AA.  The value of the damping coefficient was
934 > a cutoff radius of 12\AA.  The value of the damping coefficient was
935   also varied from the undamped case ($\alpha = 0$) to a heavily damped
936 < case ($\alpha = 0.3$ \AA$^{-1}$) for the real space methods.  A sample
937 < was also run using the multipolar Ewald sum.
936 > case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
937 > sample was also run using the multipolar Ewald sum with the same
938 > real-space cutoff.
939  
940   In figure~\ref{fig:energyDrift} we show the both the linear drift in
941   energy over time, $\delta E_1$, and the standard deviation of energy
# Line 864 | Line 945 | somewhat reduced.
945   while the hard cutoff is essentially unusable for molecular dynamics.
946   SP provides some benefit over the hard cutoff because the energetic
947   jumps that happen as particles leave and enter the cutoff sphere are
948 < somewhat reduced.
948 > somewhat reduced, but like the Wolf method for charges, the SP method
949 > would not be as useful for molecular dynamics as either of the
950 > shifted-force methods.
951  
952   We note that for all tested values of the cutoff radius, the new
953   real-space methods can provide better energy conservation behavior
# Line 880 | Line 963 | $k$-space cutoff values.
963    energy over time and $\delta \mathrm{E}_0$ is the standard deviation
964    of energy fluctuations around this drift.  All simulations were of a
965    2000-molecule simulation of SSDQ water with 48 ionic charges at 300
966 <  K starting from the same initial configuration.}
966 >  K starting from the same initial configuration. All runs utilized
967 >  the same real-space cutoff, $r_c = 12$\AA.}
968   \end{figure}
969  
970  
971   \section{CONCLUSION}
972 < We have generalized the charged neutralized potential energy
973 < originally developed by the Wolf et al.\cite{Wolf:1999dn} for the
974 < charge-charge interaction to the charge-multipole and
975 < multipole-multipole interaction in the SP method for higher order
976 < multipoles. Also, we have developed GSF and TSF methods by
977 < implementing the modification purposed by Fennel and
978 < Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the
979 < higher order multipoles to ensure consistency and smooth truncation of
980 < the electrostatic energy, force, and torque for the spherical
897 < truncation. The SP methods for multipoles proved its suitability in MC
898 < simulations. On the other hand, the results from the GSF method
899 < produced good agreement with the Ewald's energy, force, and
900 < torque. Also, it shows very good energy conservation in MD
901 < simulations.  The direct truncation of any molecular system without
902 < multipole neutralization creates the fluctuation in the electrostatic
903 < energy. This fluctuation in the energy is very large for the case of
904 < crystal because of long range of multipole ordering (Refer paper
905 < I).\cite{PaperI} This is also significant in the case of the liquid
906 < because of the local multipole ordering in the molecules. If the net
907 < multipole within cutoff radius neutralized within cutoff sphere by
908 < placing image multiples on the surface of the sphere, this fluctuation
909 < in the energy reduced significantly. Also, the multipole
910 < neutralization in the generalized SP method showed very good agreement
911 < with the Ewald as compared to direct truncation for the evaluation of
912 < the $\triangle E$ between the configurations.  In MD simulations, the
913 < energy conservation is very important. The conservation of the total
914 < energy can be ensured by i) enforcing the smooth truncation of the
915 < energy, force and torque in the cutoff radius and ii) making the
916 < energy, force and torque consistent with each other. The GSF and TSF
917 < methods ensure the consistency and smooth truncation of the energy,
918 < force and torque at the cutoff radius, as a result show very good
919 < total energy conservation. But the TSF method does not show good
920 < agreement in the absolute value of the electrostatic energy, force and
921 < torque with the Ewald.  The GSF method has mimicked Ewald’s force,
922 < energy and torque accurately and also conserved energy. Therefore, the
923 < GSF method is the suitable method for evaluating required force field
924 < in MD simulations. In addition, the energy drift and fluctuation from
925 < the GSF method is much better than Ewald’s method for finite-sized
926 < reciprocal space.
972 > In the first paper in this series, we generalized the
973 > charge-neutralized electrostatic energy originally developed by Wolf
974 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
975 > up to quadrupolar order.  The SP method is essentially a
976 > multipole-capable version of the Wolf model.  The SP method for
977 > multipoles provides excellent agreement with Ewald-derived energies,
978 > forces and torques, and is suitable for Monte Carlo simulations,
979 > although the forces and torques retain discontinuities at the cutoff
980 > distance that prevents its use in molecular dynamics.
981  
982 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
983 < that can be made extremely efficient using spline interpolations of
984 < the radial functions.  They require no Fourier transforms or $k$-space
985 < sums, and guarantee the smooth handling of energies, forces, and
986 < torques as multipoles cross the real-space cutoff boundary.  
982 > We also developed two natural extensions of the damped shifted-force
983 > (DSF) model originally proposed by Fennel and
984 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
985 > smooth truncation of energies, forces, and torques at the real-space
986 > cutoff, and both converge to DSF electrostatics for point-charge
987 > interactions.  The TSF model is based on a high-order truncated Taylor
988 > expansion which can be relatively perturbative inside the cutoff
989 > sphere.  The GSF model takes the gradient from an images of the
990 > interacting multipole that has been projected onto the cutoff sphere
991 > to derive shifted force and torque expressions, and is a significantly
992 > more gentle approach.
993  
994 + Of the two newly-developed shifted force models, the GSF method
995 + produced quantitative agreement with Ewald energy, force, and torques.
996 + It also performs well in conserving energy in MD simulations.  The
997 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
998 + place on a potential energy surface that is significantly perturbed
999 + from Ewald-based electrostatics.  
1000 +
1001 + % The direct truncation of any electrostatic potential energy without
1002 + % multipole neutralization creates large fluctuations in molecular
1003 + % simulations.  This fluctuation in the energy is very large for the case
1004 + % of crystal because of long range of multipole ordering (Refer paper
1005 + % I).\cite{PaperI} This is also significant in the case of the liquid
1006 + % because of the local multipole ordering in the molecules. If the net
1007 + % multipole within cutoff radius neutralized within cutoff sphere by
1008 + % placing image multiples on the surface of the sphere, this fluctuation
1009 + % in the energy reduced significantly. Also, the multipole
1010 + % neutralization in the generalized SP method showed very good agreement
1011 + % with the Ewald as compared to direct truncation for the evaluation of
1012 + % the $\triangle E$ between the configurations.  In MD simulations, the
1013 + % energy conservation is very important. The conservation of the total
1014 + % energy can be ensured by i) enforcing the smooth truncation of the
1015 + % energy, force and torque in the cutoff radius and ii) making the
1016 + % energy, force and torque consistent with each other. The GSF and TSF
1017 + % methods ensure the consistency and smooth truncation of the energy,
1018 + % force and torque at the cutoff radius, as a result show very good
1019 + % total energy conservation. But the TSF method does not show good
1020 + % agreement in the absolute value of the electrostatic energy, force and
1021 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1022 + % energy and torque accurately and also conserved energy.
1023 +
1024 + The only cases we have found where the new GSF and SP real-space
1025 + methods can be problematic are those which retain a bulk dipole moment
1026 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1027 + materials, uniform weighting of the orientational contributions can be
1028 + important for converging the total energy.  In these cases, the
1029 + damping function which causes the non-uniform weighting can be
1030 + replaced by the bare electrostatic kernel, and the energies return to
1031 + the expected converged values.
1032 +
1033 + Based on the results of this work, the GSF method is a suitable and
1034 + efficient replacement for the Ewald sum for evaluating electrostatic
1035 + interactions in MD simulations.  Both methods retain excellent
1036 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1037 + energy drift and fluctuations from the GSF electrostatics are better
1038 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1039 + Because they use real-space cutoffs with moderate cutoff radii, the
1040 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1041 + increases.  Additionally, they can be made extremely efficient using
1042 + spline interpolations of the radial functions.  They require no
1043 + Fourier transforms or $k$-space sums, and guarantee the smooth
1044 + handling of energies, forces, and torques as multipoles cross the
1045 + real-space cutoff boundary.
1046 +
1047   %\bibliographystyle{aip}
1048   \newpage
1049   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines