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 4188 by gezelter, Sun Jun 15 16:27:42 2014 UTC vs.
Revision 4210 by gezelter, Fri Aug 15 18:21:49 2014 UTC

# Line 47 | Line 47 | preprint,
47  
48   %\preprint{AIP/123-QED}
49  
50 < \title{Real space alternatives to the Ewald Sum. II. Comparison of Methods}
50 > \title{Real space electrostatics for multipoles. II. Comparisons with
51 >  the Ewald Sum}
52  
53   \author{Madan Lamichhane}
54   \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
# Line 72 | Line 73 | preprint,
73    energy differences between configurations, molecular forces, and
74    torques were used to analyze how well the real-space models perform
75    relative to the more computationally expensive Ewald treatment.  We
76 <  have also investigated the energy conservation properties of the new
77 <  methods in molecular dynamics simulations. The SP method shows
78 <  excellent agreement with configurational energy differences, forces,
79 <  and torques, and would be suitable for use in Monte Carlo
80 <  calculations.  Of the two new shifted-force methods, the GSF
81 <  approach shows the best agreement with Ewald-derived energies,
82 <  forces, and torques and also exhibits energy conservation properties
83 <  that make it an excellent choice for efficient computation of
84 <  electrostatic interactions in molecular dynamics simulations.
76 >  have also investigated the energy conservation, structural, and
77 >  dynamical properties of the new methods in molecular dynamics
78 >  simulations. The SP method shows excellent agreement with
79 >  configurational energy differences, forces, and torques, and would
80 >  be suitable for use in Monte Carlo calculations.  Of the two new
81 >  shifted-force methods, the GSF approach shows the best agreement
82 >  with Ewald-derived energies, forces, and torques and also exhibits
83 >  energy conservation properties that make it an excellent choice for
84 >  efficient computation of electrostatic interactions in molecular
85 >  dynamics simulations.  Both SP and GSF are able to reproduce
86 >  structural and dyanamical properties in the liquid models with
87 >  excellent fidelity.
88   \end{abstract}
89  
90   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
# Line 117 | Line 121 | interactions,\cite{Parry:1975if,Parry:1976fq,Clarke77,
121   the three dimensional Ewald sum is appropriate for slab
122   geometries.\cite{Parry:1975if} Modified Ewald methods that were
123   developed to handle two-dimensional (2-D) electrostatic
124 < interactions,\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
125 < but these methods were originally quite computationally
124 > interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
125 > These methods were originally quite computationally
126   expensive.\cite{Spohr97,Yeh99} There have been several successful
127 < efforts that reduced the computational cost of 2-D lattice
124 < summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
127 > efforts that reduced the computational cost of 2-D lattice summations,
128   bringing them more in line with the scaling for the full 3-D
129 < treatments.  The inherent periodicity in the Ewald method can also
130 < be problematic for interfacial molecular
131 < systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
129 > treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
130 > inherent periodicity required by the Ewald method can also be
131 > problematic in a number of protein/solvent and ionic solution
132 > environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
133  
134   \subsection{Real-space methods}
135   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
# Line 173 | Line 177 | point charges.\cite{Fukuda:2013sf}
177  
178   \begin{figure}
179    \centering
180 <  \includegraphics[width=\linewidth]{schematic.pdf}
180 >  \includegraphics[width=\linewidth]{schematic.eps}
181    \caption{Top: Ionic systems exhibit local clustering of dissimilar
182      charges (in the smaller grey circle), so interactions are
183      effectively charge-multipole at longer distances.  With hard
# Line 184 | Line 188 | point charges.\cite{Fukuda:2013sf}
188      orientational averaging helps to reduce the effective range of the
189      interactions in the fluid.  Placement of reversed image multipoles
190      on the surface of the cutoff sphere recovers the effective
191 <    higher-order multipole behavior.}
188 <  \label{fig:schematic}
191 >    higher-order multipole behavior. \label{fig:schematic}}
192   \end{figure}
193  
194   One can make a similar effective range argument for crystals of point
# Line 219 | Line 222 | negaitve ions and vice versa.  The reversed-charge ima
222   limited to perfect crystals, but can also apply in disordered fluids.
223   Even at elevated temperatures, there is local charge balance in an
224   ionic liquid, where each positive ion has surroundings dominated by
225 < negaitve ions and vice versa.  The reversed-charge images on the
225 > negative ions and vice versa.  The reversed-charge images on the
226   cutoff sphere that are integral to the Wolf and DSF approaches retain
227   the effective multipolar interactions as the charges traverse the
228   cutoff boundary.
# Line 231 | Line 234 | interacting molecules traverse each other's cutoff bou
234   and reduce the effective range of the multipolar interactions as
235   interacting molecules traverse each other's cutoff boundaries.
236  
234 % Because of this reason, although the nature of electrostatic
235 % interaction short ranged, the hard cutoff sphere creates very large
236 % fluctuation in the electrostatic energy for the perfect crystal. In
237 % addition, the charge neutralized potential proposed by Wolf et
238 % al. converged to correct Madelung constant but still holds oscillation
239 % in the energy about correct Madelung energy.\cite{Wolf:1999dn}.  This
240 % oscillation in the energy around its fully converged value can be due
241 % to the non-neutralized value of the higher order moments within the
242 % cutoff sphere.
243
237   Forces and torques acting on atomic sites are fundamental in driving
238   dynamics in molecular simulations, and the damped shifted force (DSF)
239   energy kernel provides consistent energies and forces on charged atoms
# Line 303 | Line 296 | reference method, a full multipolar Ewald treatment.\c
296   reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
297  
298  
306 %\subsection{Conservation of total energy }
307 %To conserve the total energy in MD simulations, the energy, force, and torque on a central molecule due to another molecule should smoothly tends to zero as second molecule approaches to cutoff radius. In addition, the force should be derivable from the energy and vice versa. If only the first condition holds but not the second, the total energy does not conserve.\cite{Fennell:2006lq}. The hard cutoff method does not ensure the smooth transition of the energy, force, and torque at the cutoff radius.\cite{Wolf:1999dn} By placing image charge on the surface of the cutoff sphere, the smooth transition of the energy can be ensured but the force and torque remains discontinuous. Therefore the purposed methods should have smooth transition of the energy, force and torque to ensure the total energy conservation and the expression should be close to idea of placing image multipole on the surface of the cutoff sphere.
308
299   \section{\label{sec:method}Review of Methods}
300   Any real-space electrostatic method that is suitable for MD
301   simulations should have the electrostatic energy, forces and torques
302   between two sites go smoothly to zero as the distance between the
303 < sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$.  Requiring
303 > sites, $r_{ab}$ approaches the cutoff radius, $r_c$.  Requiring
304   this continuity at the cutoff is essential for energy conservation in
305   MD simulations.  The mathematical details of the shifted potential
306   (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
# Line 324 | Line 314 | U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1
314   expressed as the product of two multipole operators and a Coulombic
315   kernel,
316   \begin{equation}
317 < U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
317 > U_{ab}(r)= M_{a} M_{b} \frac{1}{r}  \label{kernel}.
318   \end{equation}
319 < where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
320 < expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
321 <    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
332 < $\bf a$, etc.
333 <
334 < % Interactions between multipoles can be expressed as higher derivatives
335 < % of the bare Coulomb potential, so one way of ensuring that the forces
336 < % and torques vanish at the cutoff distance is to include a larger
337 < % number of terms in the truncated Taylor expansion, e.g.,
338 < % %
339 < % \begin{equation}
340 < % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
341 < % \end{equation}
342 < % %
343 < % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344 < % Thus, for $f(r)=1/r$, we find
345 < % %
346 < % \begin{equation}
347 < % 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} .
348 < % \end{equation}
349 < % This function is an approximate electrostatic potential that has
350 < % vanishing second derivatives at the cutoff radius, making it suitable
351 < % for shifting the forces and torques of charge-dipole interactions.
319 > where the multipole operator for site $a$, $M_{a}$, is
320 > expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
321 > $a$, etc.
322  
323   The TSF potential for any multipole-multipole interaction can be
324   written
# Line 364 | Line 334 | performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
334   force, and torques, a Taylor expansion with $n$ terms must be
335   performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
336  
367 % To carry out the same procedure for a damped electrostatic kernel, we
368 % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369 % Many of the derivatives of the damped kernel are well known from
370 % Smith's early work on multipoles for the Ewald
371 % summation.\cite{Smith82,Smith98}
372
373 % Note that increasing the value of $n$ will add additional terms to the
374 % electrostatic potential, e.g., $f_2(r)$ includes orders up to
375 % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
376 % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377 % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
378 % for computing multipole energies, forces, and torques, and smooth
379 % cutoffs of these quantities can be guaranteed as long as the number of
380 % terms in the Taylor series exceeds the derivative order required.
381
337   For multipole-multipole interactions, following this procedure results
338   in separate radial functions for each of the distinct orientational
339   contributions to the potential, and ensures that the forces and
340   torques from each of these contributions will vanish at the cutoff
341   radius.  For example, the direct dipole dot product
342 < ($\mathbf{D}_{\bf a}
343 < \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
342 > ($\mathbf{D}_{a}
343 > \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
344   dot products:
345   \begin{equation}
346 < U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 <  \mathbf{D}_{\bf a} \cdot
348 < \mathbf{D}_{\bf b} \right) v_{21}(r) +
349 < \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
350 < \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
346 > U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 >  \mathbf{D}_{a} \cdot
348 > \mathbf{D}_{b} \right) v_{21}(r) +
349 > \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
350 > \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
351   \end{equation}
352  
353   For the Taylor shifted (TSF) method with the undamped kernel,
# Line 417 | Line 372 | U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r)
372   which have been projected onto the surface of the cutoff sphere
373   without changing their relative orientation,
374   \begin{equation}
375 < U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
376 < U_{D_{\bf a} D_{\bf b}}(r_c)
377 <   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
378 <  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
375 > U_{D_{a}D_{b}}(r)  = U_{D_{a}D_{b}}(r) -
376 > U_{D_{a}D_{b}}(r_c)
377 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
378 >  \nabla U_{D_{a}D_{b}}(r_c).
379   \end{equation}
380 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
426 <  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
380 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
381   (although the signs are reversed for the dipole that has been
382   projected onto the cutoff sphere).  In many ways, this simpler
383   approach is closer in spirit to the original shifted force method, in
# Line 435 | Line 389 | $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \fra
389   approach than they are in the Taylor-shifted method.
390  
391   For the gradient shifted (GSF) method with the undamped kernel,
392 < $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
392 > $v_{21}(r) = -\frac{1}{r^3} - \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
393   $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
394   Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
395   because the Taylor expansion retains only one term, they are
# Line 444 | Line 398 | and any multipolar site inside the cutoff radius is gi
398   In general, the gradient shifted potential between a central multipole
399   and any multipolar site inside the cutoff radius is given by,
400   \begin{equation}
401 <  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
402 <    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
403 <    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
401 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
402 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
403 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
404   \label{generic2}
405   \end{equation}
406   where the sum describes a separate force-shifting that is applied to
407   each orientational contribution to the energy.  In this expression,
408   $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
409 < ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
409 > ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
410   represent the orientations the multipoles.
411  
412   The third term converges more rapidly than the first two terms as a
# Line 479 | Line 433 | U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf
433   interactions with the central multipole and the image. This
434   effectively shifts the total potential to zero at the cutoff radius,
435   \begin{equation}
436 < U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
437 < U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
436 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
437 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
438   \label{eq:SP}
439   \end{equation}          
440   where the sum describes separate potential shifting that is done for
# Line 525 | Line 479 | in the test cases are given in table~\ref{tab:pars}.
479   used the multipolar Ewald sum as a reference method for comparing
480   energies, forces, and torques for molecular models that mimic
481   disordered and ordered condensed-phase systems.  The parameters used
482 < in the test cases are given in table~\ref{tab:pars}.
482 > in the test cases are given in table~\ref{tab:pars}.
483  
484   \begin{table}
531 \label{tab:pars}
485   \caption{The parameters used in the systems used to evaluate the new
486    real-space methods.  The most comprehensive test was a liquid
487    composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
488    ions).  This test excercises all orders of the multipolar
489 <  interactions developed in the first paper.}
489 >  interactions developed in the first paper.\label{tab:pars}}
490   \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
491               & \multicolumn{2}{c|}{LJ parameters} &
492               \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
# Line 573 | Line 526 | the simulation proceeds. These differences are the mos
526   and have been compared with the values obtained from the multipolar
527   Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
528   between two configurations is the primary quantity that governs how
529 < the simulation proceeds. These differences are the most imporant
529 > the simulation proceeds. These differences are the most important
530   indicators of the reliability of a method even if the absolute
531   energies are not exact.  For each of the multipolar systems listed
532   above, we have compared the change in electrostatic potential energy
# Line 595 | Line 548 | approximations.\cite{Smith82,Smith98} In all cases, th
548   with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
549   the Ewald sum is a re-implementation of the algorithm originally
550   proposed by Smith that does not use the particle mesh or smoothing
551 < approximations.\cite{Smith82,Smith98} In all cases, the quantities
552 < being compared are the electrostatic contributions to energies, force,
553 < and torques.  All other contributions to these quantities (i.e. from
554 < Lennard-Jones interactions) are removed prior to the comparisons.
551 > approximations.\cite{Smith82,Smith98} This implementation was tested
552 > extensively against the analytic energy constants for the multipolar
553 > lattices that are discussed in reference \onlinecite{PaperI}.  In all
554 > cases discussed below, the quantities being compared are the
555 > electrostatic contributions to energies, force, and torques.  All
556 > other contributions to these quantities (i.e. from Lennard-Jones
557 > interactions) are removed prior to the comparisons.
558  
559   The convergence parameter ($\alpha$) also plays a role in the balance
560   of the real-space and reciprocal-space portions of the Ewald
# Line 737 | Line 693 | model must allow for long simulation times with minima
693  
694   \section{\label{sec:result}RESULTS}
695   \subsection{Configurational energy differences}
740 %The magnitude of the fluctuation in the total electrostatic energy per molecule for a dipolar crystal is very high as shown in (Fig … paper I).\cite{PaperI}As soon as, the net dipole moment within a cutoff radius is neutralized in the SP method, the magnitude of the fluctuation in the total electrostatic energy per molecule reduced significantly and rapidly converged to the correct energy constant (Refer figure … Paper I).\cite{PaperI} The GSF potential energy also converged to the correct energy constant for the cutoff radius rc = 6a for the undamped case. The potential energy from the TSF method converges towards the correct value for a very large cutoff radius. The speed of convergence for the all the cutoff methods can be increased by using damping function as shown in figure … Paper I\cite{PaperI}. For the quadrupolar crystals, the fluctuation in the total electrostatic energy for the hard cutoff method is small and short ranged as compared to the dipolar crystals (figure … in the paper I).\cite{PaperI}  Similar to the dipolar crystals, the net quadrupolar neutralization of the cutoff sphere in the SP method reduces oscillation rapidly and converge electrostatic energy to the correct energy constant.
741 %The oscillation in the the electrostatic energy for the hard cutoff method is even true for the dipolar liquids as shown in Figure ~\ref{fig:rcutConvergence_dipolarLiquid}. As we placed image on the surface of the cutoff sphere in SP method, the oscillation in the energy is reduced. The fluctuation in the energy in liquid is much smaller as compared to the crystal (This result is similar to the results observed by \textit{Wolf et al.} in the case of ionic crystal and Mgo melt). The large magnitude in the fluctuation of the electrostatic energy in the crystal is because of large range of multipole ordering in the crystal. When the energy is evaluated by the direct truncation, it breaks up large number of multipolar ordering leaving behind net multipole moment. But in the case of liquid, there is only local ordering of the multipoles and their ordering disappears in the long range. Therefore, the direct truncation results a small oscillation in the electrostatic energy (which is smaller than deviation SP energy from the Ewald Figure ~\ref{fig:rcutConvergence_dipolarLiquid}) in the case of liquid. Although, the oscillation in the energy is very small for the case of liquid, this affects the change in potential energy ($\triangle E$), which is observed when $\triangle E$ evaluated from the SP method compared with Ewald as shown in figure 4a and 4b.
742 %\begin{figure}[h!]
743 %        \centering
744 %        \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
745 %        \caption{The energy per molecule plotted against cutoff radius, rc for i) Hard ii) SP iii) GSF, and iv) TSF method. The hard cutoff method shows fluctuation in the electorstatic energy and it disappers in all other methods.  }
746 %        \label{fig:rcutConvergence_dipolarLiquid}
747 %    \end{figure}
748 %In MC simulations, the electrostatic differences between the molecules are important parameter for sampling. We have compared $\triangle E$ from the different methods (Hard, SP, GSF, and TSF) with the Ewald using linear regression analysis. The correlation coefficient ($R^2$) of the regression line measures the deviation of the evaluated quantities from the mean slope. We know that Ewald method evaluates accurate value of the electrostatic energy. Hence, if the proposed methods can quantify the electrostatic energy as good as Ewald then the correlation coefficient is 1.The correlation coefficient is 0 for the completely random result for any physical quantity measured by the proposed method. The slope is a measure of the accuracy of the average of a physical quantity obtained from the proposed methods. If the slope is 1 then we can conclude that the average of the physical quantity measured by the method is as good as Ewald. The deviation of the slope from 1 state that the method used in quantifying physical quantities is statistically biased as compared to Ewald.
749 %\begin{figure}
750 %        \centering
751 %        \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
752 %        \label{fig:barGraph1}
753 %        \end{figure}
754 %        \begin{figure}
755 %        \centering
756 %       \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
757 %        \caption{}
758      
759 %        \label{fig:barGraph2}
760 %      \end{figure}
761 %The correlation coefficient ($R^2$) and slope of the linear
762 %regression plots for the energy differences for all six different
763 %molecular systems is shown in figure 4a and 4b.The plot shows that
764 %the correlation coefficient improves for the SP cutoff method as
765 %compared to the undamped hard cutoff method in the case of SSDQC,
766 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
767 %crystal and liquid, the correlation coefficient is almost unchanged
768 %and close to 1.  The correlation coefficient is smallest (0.696276
769 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
770 %charge-charge and charge-multipole interactions. Since the
771 %charge-charge and charge-multipole interaction is long ranged, there
772 %is huge deviation of correlation coefficient from 1. Similarly, the
773 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
774 %compared to interactions in the other multipolar systems, thus the
775 %correlation coefficient very close to 1 even for hard cutoff
776 %method. The idea of placing image multipole on the surface of the
777 %cutoff sphere improves the correlation coefficient and makes it close
778 %to 1 for all types of multipolar systems. Similarly the slope is
779 %hugely deviated from the correct value for the lower order
780 %multipole-multipole interaction and slightly deviated for higher
781 %order multipole – multipole interaction. The SP method improves both
782 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
783 %dipolar systems.  The Slope is found to be deviated more in dipolar
784 %crystal as compared to liquid which is associated with the large
785 %fluctuation in the electrostatic energy in crystal. The GSF also
786 %produced better values of correlation coefficient and slope with the
787 %proper selection of the damping alpha (Interested reader can consult
788 %accompanying supporting material). The TSF method gives good value of
789 %correlation coefficient for the dipolar crystal, dipolar liquid,
790 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
791 %regression slopes are significantly deviated.
696  
697   \begin{figure}
698    \centering
699 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
699 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
700    \caption{Statistical analysis of the quality of configurational
701      energy differences for the real-space electrostatic methods
702      compared with the reference Ewald sum.  Results with a value equal
# Line 800 | Line 704 | model must allow for long simulation times with minima
704      from those obtained using the multipolar Ewald sum.  Different
705      values of the cutoff radius are indicated with different symbols
706      (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
707 <    triangles).}
804 <  \label{fig:slopeCorr_energy}
707 >    triangles).\label{fig:slopeCorr_energy}}
708   \end{figure}
709  
710   The combined correlation coefficient and slope for all six systems is
# Line 863 | Line 766 | forces is desired.
766   systematic error in the forces is concerning if replication of Ewald
767   forces is desired.
768  
769 + It is important to note that the forces and torques from the SP and
770 + the Hard cutoffs are not identical. The SP method shifts each
771 + orientational contribution separately (e.g. the dipole-dipole dot
772 + product is shifted by a different function than the dipole-distance
773 + products), while the hard cutoff contains no orientation-dependent
774 + shifting.  The forces and torques for these methods therefore diverge
775 + for multipoles even though the forces for point charges are identical.
776 +
777   \begin{figure}
778    \centering
779 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
779 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
780    \caption{Statistical analysis of the quality of the force vector
781      magnitudes for the real-space electrostatic methods compared with
782      the reference Ewald sum. Results with a value equal to 1 (dashed
783      line) indicate force magnitude values indistinguishable from those
784      obtained using the multipolar Ewald sum.  Different values of the
785      cutoff radius are indicated with different symbols (9\AA\ =
786 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
787 <  \label{fig:slopeCorr_force}
786 >    circles, 12\AA\ = squares, and 15\AA\ = inverted
787 >    triangles).\label{fig:slopeCorr_force}}
788   \end{figure}
789  
790  
791   \begin{figure}
792    \centering
793 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
793 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
794    \caption{Statistical analysis of the quality of the torque vector
795      magnitudes for the real-space electrostatic methods compared with
796      the reference Ewald sum. Results with a value equal to 1 (dashed
797      line) indicate force magnitude values indistinguishable from those
798      obtained using the multipolar Ewald sum.  Different values of the
799      cutoff radius are indicated with different symbols (9\AA\ =
800 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
801 <  \label{fig:slopeCorr_torque}
800 >    circles, 12\AA\ = squares, and 15\AA\ = inverted
801 >    triangles).\label{fig:slopeCorr_torque}}
802   \end{figure}
803  
804   The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
# Line 937 | Line 848 | systematically improved by varying $\alpha$ and $r_c$.
848  
849   \begin{figure}
850    \centering
851 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
851 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
852    \caption{The circular variance of the direction of the force and
853      torque vectors obtained from the real-space methods around the
854      reference Ewald vectors. A variance equal to 0 (dashed line)
855      indicates direction of the force or torque vectors are
856      indistinguishable from those obtained from the Ewald sum. Here
857      different symbols represent different values of the cutoff radius
858 <    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
948 <  \label{fig:slopeCorr_circularVariance}
858 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)\label{fig:slopeCorr_circularVariance}}
859   \end{figure}
860  
861   \subsection{Energy conservation\label{sec:conservation}}
# Line 957 | Line 867 | temperature of 300K.  After equilibration, this liquid
867   in this series and provides the most comprehensive test of the new
868   methods.  A liquid-phase system was created with 2000 water molecules
869   and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
870 < temperature of 300K.  After equilibration, this liquid-phase system
871 < was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
872 < a cutoff radius of 12\AA.  The value of the damping coefficient was
873 < also varied from the undamped case ($\alpha = 0$) to a heavily damped
874 < case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
875 < sample was also run using the multipolar Ewald sum with the same
876 < real-space cutoff.
870 > temperature of 300K.  After equilibration in the canonical (NVT)
871 > ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
872 > was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
873 > Hard, SP, GSF, and TSF methods with a cutoff radius of 12\AA.  The
874 > value of the damping coefficient was also varied from the undamped
875 > case ($\alpha = 0$) to a heavily damped case ($\alpha = 0.3$
876 > \AA$^{-1}$) for all of the real space methods.  A sample was also run
877 > using the multipolar Ewald sum with the same real-space cutoff.
878  
879   In figure~\ref{fig:energyDrift} we show the both the linear drift in
880   energy over time, $\delta E_1$, and the standard deviation of energy
# Line 979 | Line 890 | than the multipolar Ewald sum, even when utilizing a r
890  
891   We note that for all tested values of the cutoff radius, the new
892   real-space methods can provide better energy conservation behavior
893 < than the multipolar Ewald sum, even when utilizing a relatively large
894 < $k$-space cutoff values.
893 > than the multipolar Ewald sum, even when relatively large $k$-space
894 > cutoff values are utilized.
895  
896   \begin{figure}
897    \centering
898 <  \includegraphics[width=\textwidth]{newDrift_12.pdf}
899 < \label{fig:energyDrift}        
900 < \caption{Analysis of the energy conservation of the real-space
901 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
902 <  energy over time (in kcal / mol / particle / ns) and $\delta
903 <  \mathrm{E}_0$ is the standard deviation of energy fluctuations
904 <  around this drift (in kcal / mol / particle).  All simulations were
905 <  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
906 <  300 K starting from the same initial configuration. All runs
996 <  utilized the same real-space cutoff, $r_c = 12$\AA.}
898 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
899 >  \caption{Energy conservation of the real-space methods for the SSDQ
900 >    water/ion system. $\delta \mathrm{E}_1$ is the linear drift in
901 >    energy over time (in kcal/mol/particle/ns) and $\delta
902 >    \mathrm{E}_0$ is the standard deviation of energy fluctuations
903 >    around this drift (in kcal/mol/particle).  Points that appear in
904 >    the green region at the bottom exhibit better energy conservation
905 >    than would be obtained using common parameters for Ewald-based
906 >    electrostatics.\label{fig:energyDrift}}
907   \end{figure}
908  
909 + \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
910 + The most important test of the modified interaction potentials is the
911 + fidelity with which they can reproduce structural features and
912 + dynamical properties in a liquid.  One commonly-utilized measure of
913 + structural ordering is the pair distribution function, $g(r)$, which
914 + measures local density deviations in relation to the bulk density.  In
915 + the electrostatic approaches studied here, the short-range repulsion
916 + from the Lennard-Jones potential is identical for the various
917 + electrostatic methods, and since short range repulsion determines much
918 + of the local liquid ordering, one would not expect to see many
919 + differences in $g(r)$.  Indeed, the pair distributions are essentially
920 + identical for all of the electrostatic methods studied (for each of
921 + the different systems under investigation).  An example of this
922 + agreement for the SSDQ water/ion system is shown in
923 + Fig. \ref{fig:gofr}.
924  
925 + \begin{figure}
926 +  \centering
927 +  \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
928 + \caption{The pair distribution functions, $g(r)$, for the SSDQ
929 +  water/ion system obtained using the different real-space methods are
930 +  essentially identical with the result from the Ewald
931 +  treatment.\label{fig:gofr}}
932 + \end{figure}
933 +
934 + There is a minor overstructuring of the first solvation shell when
935 + using TSF or when overdamping with any of the real-space methods.
936 + With moderate damping, GSF and SP produce pair distributions that are
937 + identical (within numerical noise) to their Ewald counterparts.  The
938 + degree of overstructuring can be measured most easily using the
939 + coordination number,
940 + \begin{equation}
941 + n_C = 4\pi\rho \int_{0}^{a}r^2\text{g}(r)dr,
942 + \end{equation}
943 + where $\rho$ is the number density of the site-site pair interactions,
944 + $a$ and is the radial location of the minima following the first peak
945 + in $g(r)$ ($a = 4.2$ \AA for the SSDQ water/ion system).  The
946 + coordination number is shown as a function of the damping coefficient
947 + for all of the real space methods in Fig.  \ref{fig:Props}.
948 +
949 + A more demanding test of modified electrostatics is the average value
950 + of the electrostatic energy $\langle U_\mathrm{elect} \rangle / N$
951 + which is obtained by sampling the liquid-state configurations
952 + experienced by a liquid evolving entirely under the influence of each
953 + of the methods.  In fig \ref{fig:Props} we demonstrate how $\langle
954 + U_\mathrm{elect} \rangle / N$ varies with the damping parameter,
955 + $\alpha$, for each of the methods.
956 +
957 + As in the crystals studied in the first paper, damping is important
958 + for converging the mean electrostatic energy values, particularly for
959 + the two shifted force methods (GSF and TSF).  A value of $\alpha
960 + \approx 0.2$ \AA$^{-1}$ is sufficient to converge the SP and GSF
961 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
962 + dramatic damping ($\alpha \approx 0.28$ \AA$^{-1}$ for $r_c = 9$ \AA).
963 + Overdamping the real-space electrostatic methods occurs with $\alpha >
964 + 0.3$, causing the estimate of the electrostatic energy to drop below
965 + the Ewald results.
966 +
967 + These ``optimal'' values of the damping coefficient are slightly
968 + larger than what were observed for DSF electrostatics for purely
969 + point-charge systems, although the range $\alpha= 0.175 \rightarrow
970 + 0.225$ \AA$^{-1}$ for $r_c = 12$\AA\ appears to be an excellent
971 + compromise for mixed charge/multipolar systems.
972 +
973 + To test the fidelity of the electrostatic methods at reproducing
974 + \textit{dynamics} in a multipolar liquid, it is also useful to look at
975 + transport properties, particularly the diffusion constant,
976 + \begin{equation}
977 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
978 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
979 + \label{eq:diff}
980 + \end{equation}
981 + which measures long-time behavior and is sensitive to the forces on
982 + the multipoles.  The self-diffusion constants (D) were calculated from
983 + linear fits to the long-time portion of the mean square displacement,
984 + $\langle r^{2}(t) \rangle$.\cite{Allen87} In fig. \ref{fig:Props} we
985 + demonstrate how the diffusion constant depends on the choice of
986 + real-space methods and the damping coefficient.  Both the SP and GSF
987 + methods can obtain excellent agreement with Ewald again using moderate
988 + damping.
989 +
990 + In addition to translational diffusion, orientational relaxation times
991 + were calculated for comparisons with the Ewald simulations and with
992 + experiments. These values were determined from the same 1~ns
993 + microcanonical trajectories used for translational diffusion by
994 + calculating the orientational time correlation function,
995 + \begin{equation}
996 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
997 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
998 + \label{eq:OrientCorr}
999 + \end{equation}
1000 + where $P_l$ is the Legendre polynomial of order $l$ and
1001 + $\hat{\mathbf{A}}_\gamma$ is the unit vector for body axis $\gamma$.
1002 + The reference frame used for our sample dipolar systems has the
1003 + $z$-axis running along the dipoles, and for the SSDQ water model, the
1004 + $y$-axis connects the two implied hydrogen atom positions.  From the
1005 + orientation autocorrelation functions, we can obtain time constants
1006 + for rotational relaxation either by fitting an exponential function or
1007 + by integrating the entire correlation function.  In a good water
1008 + model, these decay times would be comparable to water orientational
1009 + relaxation times from nuclear magnetic resonance (NMR). The relaxation
1010 + constant obtained from $C_2^y(t)$ is normally of experimental interest
1011 + because it describes the relaxation of the principle axis connecting
1012 + the hydrogen atoms. Thus, $C_2^y(t)$ can be compared to the
1013 + intermolecular portion of the dipole-dipole relaxation from a proton
1014 + NMR signal and should provide an estimate of the NMR relaxation time
1015 + constant.\cite{Impey82}
1016 +
1017 + Results for the diffusion constants and orientational relaxation times
1018 + are shown in figure \ref{fig:Props}. From this data, it is apparent
1019 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1020 + reproduced with reasonable fidelity by the GSF method.
1021 +
1022 + \begin{figure}
1023 +  \caption{Comparison of the structural and dynamic properties for the
1024 +    combined multipolar liquid (SSDQ water + ions) for all of the
1025 +    real-space methods with $r_c = 12$\AA. Electrostatic energies,
1026 +    $\langle U_\mathrm{elect} \rangle / N$ (in kcal / mol),
1027 +    coordination numbers, $n_C$, diffusion constants (in cm$^2$
1028 +    s$^{-1}$), and rotational correlation times (in fs) all show
1029 +    excellent agreement with Ewald results for damping coefficients in
1030 +    the range $\alpha= 0.175 \rightarrow 0.225$
1031 +    \AA$^{-1}$. \label{fig:Props}}
1032 +  \includegraphics[width=\textwidth]{properties.eps}
1033 + \end{figure}
1034 +
1035 +
1036   \section{CONCLUSION}
1037   In the first paper in this series, we generalized the
1038   charge-neutralized electrostatic energy originally developed by Wolf
# Line 1009 | Line 1045 | We also developed two natural extensions of the damped
1045   distance that prevents its use in molecular dynamics.
1046  
1047   We also developed two natural extensions of the damped shifted-force
1048 < (DSF) model originally proposed by Fennel and
1049 < Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1050 < smooth truncation of energies, forces, and torques at the real-space
1051 < cutoff, and both converge to DSF electrostatics for point-charge
1052 < interactions.  The TSF model is based on a high-order truncated Taylor
1053 < expansion which can be relatively perturbative inside the cutoff
1054 < sphere.  The GSF model takes the gradient from an images of the
1055 < interacting multipole that has been projected onto the cutoff sphere
1056 < to derive shifted force and torque expressions, and is a significantly
1057 < more gentle approach.
1048 > (DSF) model originally proposed by Zahn {\it et al.} and extended by
1049 > Fennel and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
1050 > approaches provide smooth truncation of energies, forces, and torques
1051 > at the real-space cutoff, and both converge to DSF electrostatics for
1052 > point-charge interactions.  The TSF model is based on a high-order
1053 > truncated Taylor expansion which can be relatively perturbative inside
1054 > the cutoff sphere.  The GSF model takes the gradient from an images of
1055 > the interacting multipole that has been projected onto the cutoff
1056 > sphere to derive shifted force and torque expressions, and is a
1057 > significantly more gentle approach.
1058  
1059 < Of the two newly-developed shifted force models, the GSF method
1060 < produced quantitative agreement with Ewald energy, force, and torques.
1061 < It also performs well in conserving energy in MD simulations.  The
1062 < Taylor-shifted (TSF) model provides smooth dynamics, but these take
1063 < place on a potential energy surface that is significantly perturbed
1064 < from Ewald-based electrostatics.  
1059 > The GSF method produced quantitative agreement with Ewald energy,
1060 > force, and torques.  It also performs well in conserving energy in MD
1061 > simulations.  The Taylor-shifted (TSF) model provides smooth dynamics,
1062 > but these take place on a potential energy surface that is
1063 > significantly perturbed from Ewald-based electrostatics.  Because it
1064 > performs relatively poorly compared with GSF, it may seem odd that
1065 > that the TSF model was included in this work.  However, the functional
1066 > forms derived for the SP and GSF methods depend on the separation of
1067 > orientational contributions that were made visible by the Taylor
1068 > series of the electrostatic kernel at the cutoff radius. The TSF
1069 > method also has the unique property that a large number of derivatives
1070 > can be made to vanish at the cutoff radius.  This property has proven
1071 > useful in past treatments of the corrections to the fluctuation
1072 > formula for dielectric constants.\cite{Izvekov:2008wo}
1073  
1074 < % The direct truncation of any electrostatic potential energy without
1075 < % multipole neutralization creates large fluctuations in molecular
1076 < % simulations.  This fluctuation in the energy is very large for the case
1077 < % of crystal because of long range of multipole ordering (Refer paper
1078 < % I).\cite{PaperI} This is also significant in the case of the liquid
1079 < % because of the local multipole ordering in the molecules. If the net
1080 < % multipole within cutoff radius neutralized within cutoff sphere by
1081 < % placing image multiples on the surface of the sphere, this fluctuation
1082 < % in the energy reduced significantly. Also, the multipole
1083 < % neutralization in the generalized SP method showed very good agreement
1084 < % with the Ewald as compared to direct truncation for the evaluation of
1041 < % the $\triangle E$ between the configurations.  In MD simulations, the
1042 < % energy conservation is very important. The conservation of the total
1043 < % energy can be ensured by i) enforcing the smooth truncation of the
1044 < % energy, force and torque in the cutoff radius and ii) making the
1045 < % energy, force and torque consistent with each other. The GSF and TSF
1046 < % methods ensure the consistency and smooth truncation of the energy,
1047 < % force and torque at the cutoff radius, as a result show very good
1048 < % total energy conservation. But the TSF method does not show good
1049 < % agreement in the absolute value of the electrostatic energy, force and
1050 < % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1051 < % energy and torque accurately and also conserved energy.
1074 > Reproduction of both structural and dynamical features in the liquid
1075 > systems is remarkably good for both the SP and GSF models.  Pair
1076 > distribution functions are essentially equivalent to the same
1077 > functions produced using Ewald-based electrostatics, and with moderate
1078 > damping, a structural feature that directly probes the electrostatic
1079 > interaction (e.g. the mean electrostatic potential energy) can also be
1080 > made quantitative.  Dynamical features are sensitive probes of the
1081 > forces and torques produced by these methods, and even though the
1082 > smooth behavior of forces is produced by perturbing the overall
1083 > potential, the diffusion constants and orientational correlation times
1084 > are quite close to the Ewald-based results.
1085  
1086   The only cases we have found where the new GSF and SP real-space
1087   methods can be problematic are those which retain a bulk dipole moment
# Line 1059 | Line 1092 | Based on the results of this work, the GSF method is a
1092   replaced by the bare electrostatic kernel, and the energies return to
1093   the expected converged values.
1094  
1095 < Based on the results of this work, the GSF method is a suitable and
1096 < efficient replacement for the Ewald sum for evaluating electrostatic
1097 < interactions in MD simulations.  Both methods retain excellent
1098 < fidelity to the Ewald energies, forces and torques.  Additionally, the
1099 < energy drift and fluctuations from the GSF electrostatics are better
1100 < than a multipolar Ewald sum for finite-sized reciprocal spaces.
1101 < Because they use real-space cutoffs with moderate cutoff radii, the
1102 < GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1103 < increases.  Additionally, they can be made extremely efficient using
1071 < spline interpolations of the radial functions.  They require no
1072 < Fourier transforms or $k$-space sums, and guarantee the smooth
1073 < handling of energies, forces, and torques as multipoles cross the
1074 < real-space cutoff boundary.
1095 > Based on the results of this work, we can conclude that the GSF method
1096 > is a suitable and efficient replacement for the Ewald sum for
1097 > evaluating electrostatic interactions in modern MD simulations, and
1098 > the SP meethod would be an excellent choice for Monte Carlo
1099 > simulations where smooth forces and energy conservation are not
1100 > important.  Both the SP and GSF methods retain excellent fidelity to
1101 > the Ewald energies, forces and torques.  Additionally, the energy
1102 > drift and fluctuations from the GSF electrostatics are significantly
1103 > better than a multipolar Ewald sum for finite-sized reciprocal spaces.
1104  
1105 + As in all purely pairwise cutoff methods, the SP, GSF and TSF methods
1106 + are expected to scale approximately {\it linearly} with system size,
1107 + and are easily parallelizable.  This should result in substantial
1108 + reductions in the computational cost of performing large simulations.
1109 + With the proper use of pre-computation and spline interpolation of the
1110 + radial functions, the real-space methods are essentially the same cost
1111 + as a simple real-space cutoff.  They require no Fourier transforms or
1112 + $k$-space sums, and guarantee the smooth handling of energies, forces,
1113 + and torques as multipoles cross the real-space cutoff boundary.
1114 +
1115 + We are not suggesting that there is any flaw with the Ewald sum; in
1116 + fact, it is the standard by which the SP, GSF, and TSF methods have
1117 + been judged in this work.  However, these results provide evidence
1118 + that in the typical simulations performed today, the Ewald summation
1119 + may no longer be required to obtain the level of accuracy most
1120 + researchers have come to expect.
1121 +
1122   \begin{acknowledgments}
1123    JDG acknowledges helpful discussions with Christopher
1124    Fennell. Support for this project was provided by the National

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines