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 4166 by mlamichh, Tue Jun 3 18:49:15 2014 UTC vs.
Revision 4185 by gezelter, Sun Jun 15 00:18:18 2014 UTC

# Line 20 | Line 20
20   % Use this file as a source of example code for your aip document.
21   % Use the file aiptemplate.tex as a template for your document.
22   \documentclass[%
23 < aip,
24 < jmp,
23 > aip,jcp,
24   amsmath,amssymb,
25 < %preprint,%
26 < reprint,%
25 > preprint,
26 > %reprint,%
27   %author-year,%
28   %author-numerical,%
29   ]{revtex4-1}
30  
31   \usepackage{graphicx}% Include figure files
32   \usepackage{dcolumn}% Align table columns on decimal point
33 < \usepackage{bm}% bold math
33 > %\usepackage{bm}% bold math
34   %\usepackage[mathlines]{lineno}% Enable numbering of text and display math
35   %\linenumbers\relax % Commence numbering lines
36   \usepackage{amsmath}
37 + \usepackage{times}
38 + \usepackage{mathptmx}
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. performance in condensed phase simulations}% 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 60 | 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 multipole interactions that were developed in the first paper in
73 >  this series, using the multipolar Ewald sum as a reference
74 >  method. The tests were carried out in a variety of condensed-phase
75 >  environments which were designed to test all levels of the
76 >  multipole-multipole interactions.  Comparisons of the energy
77 >  differences between configurations, molecular forces, and torques
78 >  were used to analyze how well the real-space models perform relative
79 >  to the more computationally expensive Ewald treatment.  We have also
80 >  investigated the energy conservation properties of the new methods
81 >  in molecular dynamics simulations. The SP method shows excellent
82 >  agreement with configurational energy differences, forces, and
83 >  torques, and would be suitable for use in Monte Carlo calculations.
84 >  Of the two new shifted-force methods, the GSF approach shows the
85 >  best agreement with Ewald-derived energies, forces, and torques and
86 >  exhibits energy conservation properties that make it an excellent
87 >  choice for efficient computation of electrostatic interactions in
88 >  molecular 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{Suggested keywords}%Use showkeys class option if keyword
94 <                              %display desired
93 > %\keywords{Electrostatics, Multipoles, Real-space}
94 >
95   \maketitle
96  
97  
98   \section{\label{sec:intro}Introduction}
99 < The interaction between charges has always been the most expensive part in molecular simulations.  There have been many efforts to develop practical and efficient method for handling electrostatic interactions. The Ewald’s method has always been accepted as the most precise method for evaluating electrostatic energies, forces and torques. In this method, the conditionally convergent electrostatic energy is converted into the sum of the rapidly converging real and reciprocal space contribution of artificially made periodic system.\cite{Woodcock86, Woodcock75} Because of this artificially created periodicity, Ewald’s sum is not a suitable method to calculate electrostatic interaction in the interfacial molecular systems such as bicrystals, free surfaces, and liquid-vapor interfaces.\cite{Wolf99}To simulate such interfacial systems, the Parry’s extension of the 3D Ewald sum appropriate for the slab geometry is used,\cite{Parry75} which is computationally very expensive.  Also, the reciprocal part of the Ewald’s sum is computationally expensive which makes it inappropriate to use for the larger molecular system. By using Fast Fourier Transform(FFT) in the  particle-mesh Ewald (PME) and particle-particle particle-mesh  Ewald ($P^3ME$) in the reciprocal space term, the computational cost has been decreased from $O(N^2)$ down to $O(Nlog N)$.\cite{Takada93, Gunsteren94, Gunsteren95, Pedersen93, Pedersen95}. Although the computational time has been reduced, the inherent periodicity in the Ewald’s method can be problematic for the interfacial molecular system.\cite{Gezelter06}  Furthermore, the modified Ewald’s methods developed to handle two-dimensional (2D) electrostatic interactions\cite{Parry75, Parry76, Clarke77, Perram79,Rahman89} in the interfacial systems are also computationally expensive.\cite{Spohr97,Berkowitz99}
99 > Computing the interactions between electrostatic sites is one of the
100 > most expensive aspects of molecular simulations. There have been
101 > significant efforts to develop practical, efficient and convergent
102 > methods for handling these interactions. Ewald's method is perhaps the
103 > best known and most accurate method for evaluating energies, forces,
104 > and torques in explicitly-periodic simulation cells. In this approach,
105 > the conditionally convergent electrostatic energy is converted into
106 > two absolutely convergent contributions, one which is carried out in
107 > real space with a cutoff radius, and one in reciprocal
108 > space. BETTER CITATIONS\cite{Clarke:1986eu,Woodcock75}
109  
110 < \subsection{Real-space methods}
111 < Recently, \textit{Wolf et al.}\cite{Wolf99} proposed a real space $O(N)$ method for calculating electrostatic interaction between charges. They showed that the effective Coulomb interaction in the condensed system is actually short ranged.\cite{Wolf92, Wolf95}. Furthermore, the Madelung energy of an ion considering lattice summation over neutral dipolar molecules decreases as $r^{-5}$.\cite{Wolf92, Wolf95}. Thus, the careful application of the real-space method for a calculation of the electrostatic energy should be able to obtain correct Madelung energy for a significant size of the cutoff sphere. But the direct truncation of the cutoff sphere for the evaluation of the electrostatic energy always create truncation defect. This cutoff defect in the electrostatic energy is due to the existence of the net charge within the cutoff sphere.\cite{Wolf99} To neutralize net charge within the cutoff sphere, \textit{Wolf et al.}\cite{Wolf99} proposed a method of placing an image charge, for every charge within a cutoff sphere, on the surface to evaluate the electrostatic energy and force. Both the electrostatic energy and force for the central charge are evaluated separately from the interaction of the configuration of real charges within the cutoff sphere and image charges on the surface of the sphere. But the energy of an individual charge due to another charge within the cutoff sphere and its image on the surface is not an integral of their force, as a result the total energy does not conserve in molecular dynamic (MD) simulations.\cite{Zahn02}
110 > When carried out as originally formulated, the reciprocal-space
111 > portion of the Ewald sum exhibits relatively poor computational
112 > scaling, making it prohibitive for large systems. By utilizing
113 > particle meshes and three dimensional fast Fourier transforms (FFT),
114 > the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
115 > (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
116 > the computational cost from $O(N^2)$ down to $O(N \log
117 > N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
118  
119 < Considering the interaction of an ion with dipolar molecular shell, the effective Columbic potential for a perfect ionic crystal is found to be decreasing as $r^{-5}$.\cite{Wolf99} Furthermore, viewing the NaCl crystal as simple cubic (SC) structure with octupolar $(NaCl)_{4}$ basis, the electrostatic energy per ion converges more rapidly to Madelong than the dipolar approximation.\cite{Wolf92} Also, to find the correct Madelung constant, Lacman.\cite{Lacman65}suggested that the NaCl structure should be constructed in a such way that the finite crystal terminates with only complete $(NaCl)_4$ molecules.  These facts suggest that the Madelung energy is short ranged for a perfect ionic crystal.  
120 < \begin{figure}[h!]
121 <        \centering
122 <        \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
123 <        \caption{NaCl crystal showing (a) breaking of the charge ordering in the direct spherical truncation, and (b) complete $(NaCl)_{4}$ molecule interacting with the central ion. }
124 <        \label{fig:NaCl}
125 <    \end{figure}
126 <
127 < Any charge in a NaCl crystal is surrounded by opposite charges. Similarly for each pair of charges, there is an opposite pair of charge to its adjacent as shown in Figure ~\ref{fig:NaCl}.  Furthermore for each group of four charges, there should be an oppositely aligned group of four charges as shown in Fig 1b.  If we consider any group of charges, suppose $(NaCl)_4$, far from the central charge, they have little electrostatic interaction with  the central charge (acts like point octopole when it is far from the center ). But if the cutoff sphere passes through the $(NaCl)_4$ molecule leaving behind net positive or negative charge, the electrostatic contribution due to these broken charges going to be very large (for point charge  radial function $1/r_c$ and for point octupole $1/r_c$). Because of this reason, although the nature of electrostatic interaction short ranged, the hard cutoff sphere creates very large fluctuation in the electrostatic energy for the perfect crystal. In addition, the charge neutralized potential proposed by Wolf et al. converged to correct Madelung constant but still holds oscillation in the energy about correct Madelung energy.\cite{Wolf99}.  This oscillation in the energy around its fully converged value can be due to the non-neutralized value of the higher order moments within the cutoff sphere.  Recently, \textit{Ikuo Fukuda} used neutralization of the higher order moments for the calculation of the electrostatic interaction of the point charges system.\cite{Fukuda13}
119 > Because of the artificial periodicity required for the Ewald sum,
120 > interfacial molecular systems such as membranes and liquid-vapor
121 > interfaces require modifications to the
122 > method.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
123 > Parry's extension of the three dimensional Ewald sum is appropriate
124 > for slab geometries.\cite{Parry:1975if} Modified Ewald methods that
125 > were developed to handle two-dimensional (2D) electrostatic
126 > interactions in interfacial systems have not seen similar
127 > particle-mesh treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
128 >  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
129 > with system size. The inherent periodicity in the Ewald’s method can
130 > also be problematic for interfacial molecular
131 > systems.\cite{Fennell:2006lq}
132  
133 < The force and torque acting on molecules are the fundamental factors to drive the dynamics of the molecular simulation. \textit{Fennell and Gezelter} proposed the damped shifted force (DSF) potential energy to obtain consistent configurational force on the central charge by the charges within the cutoff sphere and their image charge on the surface. Since the force is consistent with the energy, MD simulations conserve the total energy. Also, the comparison of accuracy of the potential energy and force from DSF method with the Ewald shows surprisingly good results.\cite{Gezelter06}Now a days, the DSF method is being used in several molecular systems with uniform charge density to calculate electrostatic interaction.\cite{Luebke13, Daivis13, Acevedo13, Space12,English08, Lawrence13, Vergne13}
133 > \subsection{Real-space methods}
134 > Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
135 > method for calculating electrostatic interactions between point
136 > charges. They argued that the effective Coulomb interaction in most
137 > condensed phase systems is effectively short
138 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
139 > computing the Madelung constant of an ionic solid), the material can
140 > be considered as a set of ions interacting with neutral dipolar or
141 > quadrupolar ``molecules'' giving an effective distance dependence for
142 > the electrostatic interactions of $r^{-5}$ (see figure
143 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
144 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
145 > electrostatic energy per ion converges more rapidly to the Madelung
146 > energy than the dipolar approximation.\cite{Wolf92} To find the
147 > correct Madelung constant, Lacman suggested that the NaCl structure
148 > could be constructed in a way that the finite crystal terminates with
149 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
150 > what is effectively a set of octupoles at large distances. These facts
151 > suggest that the Madelung constants are relatively short ranged for
152 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
153 > application of Wolf's method are able to obtain accurate estimates of
154 > Madelung constants using relatively short cutoff radii.
155  
156 < \subsection{Damping function}
157 < The damping function used in our research has been discussed in detail in the paper I.\cite{PaperI} The radial function $1/r$ of the interactions between the charges can be replaced by the complementary error function $erfc(\alpha r)/r$  to accelerate the rate of convergence, where $\alpha$ is damping parameter. We can perform necessary mathematical manipulation by varying $\alpha$ in the damping function for the calculation of the electrostatic energy, force and torque\cite{Wolf99}. By using suitable value of damping alpha ($\alpha = 0.2$) for a cutoff radius ($r_{­c}=9 A$), \textit{Fennel and Gezelter}\cite{Gezelter06} produced very good agreement of the interaction energies, forces and torques for charge-charge interactions.\cite{Gezelter06}
156 > Direct truncation of interactions at a cutoff radius creates numerical
157 > errors.  Wolf \textit{et al.}  argued that truncation errors are due
158 > to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
159 > neutralize this charge they proposed placing an image charge on the
160 > surface of the cutoff sphere for every real charge inside the cutoff.
161 > These charges are present for the evaluation of both the pair
162 > interaction energy and the force, although the force expression
163 > maintained a discontinuity at the cutoff sphere.  In the original Wolf
164 > formulation, the total energy for the charge and image were not equal
165 > to the integral of their force expression, and as a result, the total
166 > energy would not be conserved in molecular dynamics (MD)
167 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
168 > Gezelter later proposed shifted force variants of the Wolf method with
169 > commensurate force and energy expressions that do not exhibit this
170 > problem.\cite{Fennell:2006lq} Related real-space methods were also
171 > proposed by Chen \textit{et
172 >  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
173 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has used
174 > neutralization of the higher order moments for the calculation of the
175 > electrostatic interaction of the point charge
176 > systems.\cite{Fukuda:2013sf}
177  
178 < \subsection{Point multipoles for CG modeling}
179 < Since a molecule consists of equal positive and negative charges, instead taking of the most common case of atomic site-site interaction, the interaction between higher order multipoles can also be used to evaluate molecule-molecule interactions. The short-ranged interaction between the molecules is dominated by Lennard-Jones repulsion. Also, electrons in a molecule is not localized at a specific point, thus a molecule can be coarse-grained to approximate as point multipole.\cite{Ren06, Essex10, Essex11}Recently, water has been modeled with point multipoles up to octupolar order.\cite{Ichiye10_1, Ichiye10_2, Ichiye10_3}. The point multipoles method has also been used in the AMOEBA water model.\cite{Gordon10, Gordon07,Smith80}. But using point multipole in the real space cutoff method without account of multipolar neutrality creates problem in the total energy conservation in MD simulations. In this paper we extended the original idea of the charge neutrality by Wolf’s into point dipoles and quadrupoles. Also, we used the previously developed idea of the damped shifted potential (DSF) for the charge-charge interaction\cite{Gezelter06}and generalized it into higher order multipoles to conserve the total energy in the molecular dynamic simulation (The detail mathematical development of the purposed methods have been discussed in paper I).
178 > \begin{figure}
179 >  \centering
180 >  \includegraphics[width=\linewidth]{schematic.pdf}
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
184 >    cutoffs, motion of individual charges in and out of the cutoff
185 >    sphere can break the effective multipolar ordering.  Bottom:
186 >    dipolar crystals and fluids have a similar effective
187 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
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.}
192 >  \label{fig:schematic}
193 > \end{figure}
194  
195 + One can make a similar effective range argument for crystals of point
196 + \textit{multipoles}. The Luttinger and Tisza treatment of energy
197 + constants for dipolar lattices utilizes 24 basis vectors that contain
198 + dipoles at the eight corners of a unit cube.  Only three of these
199 + basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
200 + moments, while the rest have zero net dipole and retain contributions
201 + only from higher order multipoles.  The lowest energy crystalline
202 + structures are built out of basis vectors that have only residual
203 + quadrupolar moments (e.g. the $Z_5$ array). In these low energy
204 + structures, the effective interaction between a dipole at the center
205 + of a crystal and a group of eight dipoles farther away is
206 + significantly shorter ranged than the $r^{-3}$ that one would expect
207 + for raw dipole-dipole interactions.  Only in crystals which retain a
208 + bulk dipole moment (e.g. ferroelectrics) does the analogy with the
209 + ionic crystal break down -- ferroelectric dipolar crystals can exist,
210 + while ionic crystals with net charge in each unit cell would be
211 + unstable.
212  
213 < %\subsection{Conservation of total energy }
214 < %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{Gezelter06}. The hard cutoff method does not ensure the smooth transition of the energy, force, and torque at the cutoff radius.\cite{Wolf99} 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.
213 > In ionic crystals, real-space truncation can break the effective
214 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
215 > significant swings in the electrostatic energy as individual ions move
216 > back and forth across the boundary.  This is why the image charges are
217 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
218 > the real-space truncation of point multipole interactions breaks
219 > higher order multipole arrangements, and image multipoles are required
220 > for real-space treatments of electrostatic energies.
221  
222 < \section{\label{sec:method}REVIEW OF METHODS}
223 < Any force field associated with MD simulation should have the electrostatic energy, force and the torque between central molecule and any other molecule within cutoff radius should smoothly approach to zero as $r$ tends to $r_c$. This issue of continuous nature of the electrostatic interaction at the cutoff radius is associated with the conservation of total energy in the MD simulation. The mathematical detail for the SP, GSF and TSF has already been discussed in detail in previous paper I.\cite{PaperI}
222 > The shorter effective range of electrostatic interactions is not
223 > limited to perfect crystals, but can also apply in disordered fluids.
224 > Even at elevated temperatures, there is, on average, local charge
225 > balance in an ionic liquid, where each positive ion has surroundings
226 > dominated by negaitve ions and vice versa.  The reversed-charge images
227 > on the cutoff sphere that are integral to the Wolf and DSF approaches
228 > retain the effective multipolar interactions as the charges traverse
229 > the cutoff boundary.
230  
231 < \subsection{Taylor-shifted force(TSF)}
232 < The detail mathematical expression for the multipole-multipole interaction by the TSF method has been described in paper I.\cite{PaperI}. The electrostatic potential energy between groups of charges or multipoles is expressed as the product of operator and potential due to point charge as shown in \textit{equation 4 in Paper I}.\cite{PaperI}  In the Taylor Shifted Force (TSF) method, we shifted kernel $1/r$ (the potential due to a point charge) by $1/r_c$ and performed Taylor Series expansion of the shifted part about the cutoff radius before operating with the operators. To ensure smooth convergence of the energy, force, and torque  to zero at the cut off radius, the required number of terms from Taylor Series expansion are performed for different multipole-multipole interactions. Also, the mathematical consistency between the energy, force and the torque has been established. The potential energy for the multipole-multipole interaction is given by,
231 > In multipolar fluids (see Fig. \ref{fig:schematic}) there is
232 > significant orientational averaging that additionally reduces the
233 > effect of long-range multipolar interactions.  The image multipoles
234 > that are introduced in the TSF, GSF, and SP methods mimic this effect
235 > and reduce the effective range of the multipolar interactions as
236 > interacting molecules traverse each other's cutoff boundaries.
237  
238 < \begin{equation}
239 < \begin{split}
240 < U_{TSF}(\vec r)=\sum_{\alpha=1}^3\sum_{\beta=1}^3(C_a - D_{a \alpha }\frac{\partial}{\partial r_{a \alpha}}+Q_{a \alpha \beta }\frac{\partial}{\partial r_{a \alpha}\partial r_{a \beta}})\\
241 < (C_b - D_{b \alpha }\frac{\partial}{\partial r_{b \alpha}}+Q_{b \alpha \beta }\frac{\partial}{\partial r_{b \alpha}\partial r_{b \beta}})\\
242 < [(\frac{1}{r}-[\frac{1}{r_c}-(r-r_c)\frac{1}{r_c^2}+(r-r_c)^2\frac{1}{r_c^3}+...)]
243 < \end{split}
244 < \label{eq:TSF}
245 < \end{equation}
246 <  
116 < where $C_a = \sum_{k\;in\; a}q_k$ , $D_{a\alpha}=\sum_{k \;in\;a}q_k r_k\alpha$, and $Q_{a\alpha,\beta}=\frac{1}{2}\sum_{k\;in\;a}q_k r_{k\alpha}r_{k\beta}$ stand for charge, dipole and quadrupole moment respectively (detail in paperI\cite{PaperI}). The electrostatic force and torque acting on the central molecule due to a molecule within cutoff sphere are derived from the equation ~\ref{eq:TSF} with the account of appropriate number of terms.  This method is developed on the basis of using kernel potential due to the point charge ($1/r$) and their image charge potential ($1/r_c$) with its Taylor series expansion and considering that the expression for multipole-multipole interaction can be obtained operating the modified kernel by their corresponding operators.
238 > % Because of this reason, although the nature of electrostatic
239 > % interaction short ranged, the hard cutoff sphere creates very large
240 > % fluctuation in the electrostatic energy for the perfect crystal. In
241 > % addition, the charge neutralized potential proposed by Wolf et
242 > % al. converged to correct Madelung constant but still holds oscillation
243 > % in the energy about correct Madelung energy.\cite{Wolf:1999dn}.  This
244 > % oscillation in the energy around its fully converged value can be due
245 > % to the non-neutralized value of the higher order moments within the
246 > % cutoff sphere.
247  
248 < \subsection{Shifted potential (SP) }
249 < A discontinuous truncation of the electrostatic potential at the cutoff sphere introduces severe artifact(Oscillation in the electrostatic energy) even for molecules with the higher-order multipoles.\cite{Paper I} This artifact is due to the existence of multipole moments within the cutoff spheres contributed by the breaking of the multipole ordering at the the surface of the cutoff sphere. The multipole moments of the cutoff sphere can be neutralized by placing image multipole for every multipole within the cutoff sphere. The electrostatic potential between multipoles for the SP method is given by,
250 < \begin{equation}
251 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
252 < \label{eq:SP}
253 < \end{equation}          
254 < The SP method compensates the artifact created by truncation of the multipole ordering by placing image on the cutoff surface.  Also, the potential energy between central multipole and other multipole within sphere approaches smoothly to zero as $r$ tends to $r_c$. But the force and torque obtained from the shifted potential are discontinuous at $r_c$. Therefore, the MD simulation will still have the total energy drift for a longer simulation.  If we derive the force and torque from the direct shifting about $r_c$ like in shifted potential then inconsistency between the force, torque, and potential fails the energy conservation in the dynamic simulation.
248 > The forces and torques acting on atomic sites are the fundamental
249 > factors driving dynamics in molecular simulations. Fennell and
250 > Gezelter proposed the damped shifted force (DSF) energy kernel to
251 > obtain consistent energies and forces on the atoms within the cutoff
252 > sphere. Both the energy and the force go smoothly to zero as an atom
253 > aproaches the cutoff radius. The comparisons of the accuracy these
254 > quantities between the DSF kernel and SPME was surprisingly
255 > good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
256 > calculating electrostatic interactions in molecular systems with
257 > relatively uniform charge
258 > densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
259  
260 < \subsection{Gradient-shifted force (GSF)}
261 < As we mentioned earlier, in the MD simulation the electrostatic energy, force and torque should approach to zero as r tends to $r_c$. Also, the energy, force and torque should be consistent with each other for the total energy conservation. The GSF method is developed to address both the issues of consistency and convergence of the energy, force and the torque. Furthermore, the compensating of charge or multipole ordering breakage in the SP method due to direct spherical truncation will remain intact for large $r_c$. The electrostatic potential energy between central molecule and any molecule inside cutoff radius is given by,
262 <        \begin{equation}
263 < U_{SF}(\vec r)=\sum U(\vec r) - U(\vec r_c)-(\vec r-\vec r_c)\cdot\vec \nabla U(\vec r)|_{r=r_c}
264 < \label{eq:GSF}
265 < \end{equation}    
266 < where the third term converges more rapidly as compared to first two terms hence the contribution of the third term is very small for large $r_c$ value. Hence the GSF method similar to SP method for large $r_c$. Moreover, the force and torque derived from equation 3 are consistent with the energy and approaches to zero as $r$ tends to $r_c$.
267 < Both GSF and TSF methods are the generalization of the original DSF method to higher order multipole-multipole interactions. These two methods are same up to charge-dipole interaction level but generate different expressions in the energy, force and torque for the higher order multipole-multipole interactions.
268 < \subsection{Self term}
260 > \subsection{The damping function}
261 > The damping function has been discussed in detail in the first paper
262 > of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
263 > interactions between point charges can be replaced by the
264 > complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
265 > convergence, where $\alpha$ is a damping parameter with units of
266 > inverse distance.  Altering the value of $\alpha$ is equivalent to
267 > changing the width of Gaussian charge distributions that replace each
268 > point charge, as Coulomb integrals with Gaussian charge distributions
269 > produce complementary error functions when truncated at a finite
270 > distance.
271  
272 < \section{\label{sec:test}Test systems}
273 < We have compared the electrostatic force and torque of each molecule from SP, TSF and GSF method with the multipolar-Ewald method. Furthermore, total electrostatic energies of a molecular system from the different methods have also been compared with total energy from the Ewald. In Mote Carlo (MC) simulation, the energy difference between different configurations of the molecular system is important, even though absolute energies are not accurate. We have compared the change in electrostatic potential energy ($\triangle E$) of 250 different configurations of the various multipolar molecular systems (Section IV B) calculated from the Hard, SP, GSF, and TSF methods with the well-known Ewald method. In MD simulations, the force and torque acting on the molecules drives the whole dynamics of the molecules in a system. The magnitudes of the electrostatic force, torque and their direction for each molecule of the all 250 configurations have also been compared against the Ewald’s method.
272 > With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
273 > produced very good agreement with SPME for interaction energies,
274 > forces and torques for charge-charge
275 > interactions.\cite{Fennell:2006lq}
276  
277 < \subsection{Modeled systems}
278 < We studied the comparison of the energy differences, forces and torques for six different systems; i) dipolar liquid, ii) quadrupolar liquid, iii)  dipolar crystal, iv) quadrupolar crystal v) dipolar-quadrupolar liquid(SSDQ), and vi) ions in dipolar-qudrupolar liquid(SSDQC). To simulate different configurations of the crystals, the body centered cubic (BCC) minimum energy crystal with 3,456 molecules was taken and translationally locked in their respective crystal sites. The thermal energy was supplied to the rotational motion so that dipoles or quadrupoles can freely explore all possible orientation. The crystals were simulated for 10,000 fs in NVE ensemble at 50 K and 250 different configurations was taken in equal time interval for the comparative study.  The crystals were not simulated at high temperature and for a long run time to avoid possible translational deformation of the crystal sites.
279 < For dipolar, quadrupolar, and dipolar-quadrupolar liquids simulation, each molecular system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K temperature after equilibration.  We collected 250 different configurations in equal interval of time. For the ions mixed liquid system, we converted 48 different molecules into 24 $Na^+$ and $24 Cl^-$ ions and equilibrated. After equilibration, the system was run at the same environment for 1ns and 250 configurations were collected. While comparing energies, forces, and torques with Ewald method, Lennad Jone’s potentials were turned off and purely electrostatic interaction had been compared.
277 > \subsection{Point multipoles in molecular modeling}
278 > Coarse-graining approaches which treat entire molecular subsystems as
279 > a single rigid body are now widely used. A common feature of many
280 > coarse-graining approaches is simplification of the electrostatic
281 > interactions between bodies so that fewer site-site interactions are
282 > required to compute configurational
283 > energies.\cite{Ren06,Essex10,Essex11}
284  
285 < \subsection{Statistical analysis}
286 < We have used least square regression analyses for six different molecular systems to compare $\triangle E$ from Hard, SP, GSF, and TSF with the reference method. Molecular systems were run longer enough to explore various configurations and 250 independent configurations were recorded for comparison.  The total numbers of 31,125 energy differences from the proposed methods have been compared with the Ewald.  Similarly, the magnitudes of the forces and torques have also been compared by using least square regression analyses. In the forces and torques comparison, the magnitudes of the forces acting in each molecule for each configuration were evaluated. For example, our dipolar liquid simulation contains 2048 molecules and there are 250 different configurations for each system thus there are 512,000 force and torque comparisons.  The correlation coefficient and correlation slope varies from 0 to 1, where 1 is the best agreement between the two methods.
287 <
288 < \subsection{Analysis of vector quantities}
289 < R.A. Fisher has developed a probablity density function to analyse directional data sets is expressed as below,\cite{fisher53}
290 < \begin{equation}
291 < p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta \exp(\kappa \cos\theta)
292 < \label{eq:pdf}
285 > Because electrons in a molecule are not localized at specific points,
286 > the assignment of partial charges to atomic centers is always an
287 > approximation.  Atomic sites can also be assigned point multipoles and
288 > polarizabilities to increase the accuracy of the molecular model.
289 > Recently, water has been modeled with point multipoles up to octupolar
290 > order using the soft sticky dipole-quadrupole-octupole (SSDQO)
291 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
292 > multipoles up to quadrupolar order have also been coupled with point
293 > polarizabilities in the high-quality AMOEBA and iAMOEBA water
294 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
295 > truncating point multipoles without smoothing the forces and torques
296 > will create energy conservation issues in molecular dynamics simulations.
297 >
298 > In this paper we test a set of real-space methods that were developed
299 > for point multipolar interactions.  These methods extend the damped
300 > shifted force (DSF) and Wolf methods originally developed for
301 > charge-charge interactions and generalize them for higher order
302 > multipoles. The detailed mathematical development of these methods has
303 > been presented in the first paper in this series, while this work
304 > covers the testing the energies, forces, torques, and energy
305 > conservation properties of the methods in realistic simulation
306 > environments.  In all cases, the methods are compared with the
307 > reference method, a full multipolar Ewald treatment.
308 >
309 >
310 > %\subsection{Conservation of total energy }
311 > %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.
312 >
313 > \section{\label{sec:method}Review of Methods}
314 > Any real-space electrostatic method that is suitable for MD
315 > simulations should have the electrostatic energy, forces and torques
316 > between two sites go smoothly to zero as the distance between the
317 > sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$.  Requiring
318 > this continuity at the cutoff is essential for energy conservation in
319 > MD simulations.  The mathematical details of the shifted potential
320 > (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
321 > methods have been discussed in detail in the previous paper in this
322 > series.\cite{PaperI} Here we briefly review the new methods and
323 > describe their essential features.
324 >
325 > \subsection{Taylor-shifted force (TSF)}
326 >
327 > The electrostatic potential energy between point multipoles can be
328 > expressed as the product of two multipole operators and a Coulombic
329 > kernel,
330 > \begin{equation}
331 > U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
332   \end{equation}
333 < where $\kappa$ measures directional dispersion of the data about mean direction can be estimated as a reciprocal of the circular variance for large number of directional data sets.\cite{Allen91} In our calculation, the unit vector from the Ewald method was considered as mean direction and the angle between the vectors from Ewald and the purposed method were evaluated.The total displacement of the unit vectors from the purposed method was calculated as,
333 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
334 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
335 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
336 > $\bf a$, etc.
337 >
338 > % Interactions between multipoles can be expressed as higher derivatives
339 > % of the bare Coulomb potential, so one way of ensuring that the forces
340 > % and torques vanish at the cutoff distance is to include a larger
341 > % number of terms in the truncated Taylor expansion, e.g.,
342 > % %
343 > % \begin{equation}
344 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
345 > % \end{equation}
346 > % %
347 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
348 > % Thus, for $f(r)=1/r$, we find
349 > % %
350 > % \begin{equation}
351 > % 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} .
352 > % \end{equation}
353 > % This function is an approximate electrostatic potential that has
354 > % vanishing second derivatives at the cutoff radius, making it suitable
355 > % for shifting the forces and torques of charge-dipole interactions.
356 >
357 > The TSF potential for any multipole-multipole interaction can be
358 > written
359 > \begin{equation}
360 > U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
361 > \label{generic}
362 > \end{equation}
363 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
364 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
365 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
366 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
367 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
368 > force, and torques, a Taylor expansion with $n$ terms must be
369 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
370 >
371 > % To carry out the same procedure for a damped electrostatic kernel, we
372 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
373 > % Many of the derivatives of the damped kernel are well known from
374 > % Smith's early work on multipoles for the Ewald
375 > % summation.\cite{Smith82,Smith98}
376 >
377 > % Note that increasing the value of $n$ will add additional terms to the
378 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
379 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
380 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
381 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
382 > % for computing multipole energies, forces, and torques, and smooth
383 > % cutoffs of these quantities can be guaranteed as long as the number of
384 > % terms in the Taylor series exceeds the derivative order required.
385 >
386 > For multipole-multipole interactions, following this procedure results
387 > in separate radial functions for each of the distinct orientational
388 > contributions to the potential, and ensures that the forces and
389 > torques from each of these contributions will vanish at the cutoff
390 > radius.  For example, the direct dipole dot product
391 > ($\mathbf{D}_{\bf a}
392 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
393 > dot products:
394 > \begin{equation}
395 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
396 >  \mathbf{D}_{\bf a} \cdot
397 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
398 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
399 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
400 > \end{equation}
401 >
402 > For the Taylor shifted (TSF) method with the undamped kernel,
403 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
404 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
405 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
406 > connection to unmodified electrostatics as well as the smooth
407 > transition to zero in both these functions as $r\rightarrow r_c$.  The
408 > electrostatic forces and torques acting on the central multipole due
409 > to another site within the cutoff sphere are derived from
410 > Eq.~\ref{generic}, accounting for the appropriate number of
411 > derivatives. Complete energy, force, and torque expressions are
412 > presented in the first paper in this series (Reference
413 > \onlinecite{PaperI}).
414 >
415 > \subsection{Gradient-shifted force (GSF)}
416 >
417 > A second (and conceptually simpler) method involves shifting the
418 > gradient of the raw Coulomb potential for each particular multipole
419 > order.  For example, the raw dipole-dipole potential energy may be
420 > shifted smoothly by finding the gradient for two interacting dipoles
421 > which have been projected onto the surface of the cutoff sphere
422 > without changing their relative orientation,
423 > \begin{equation}
424 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
425 > U_{D_{\bf a} D_{\bf b}}(r_c)
426 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
427 >  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
428 > \end{equation}
429 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
430 >  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
431 > (although the signs are reversed for the dipole that has been
432 > projected onto the cutoff sphere).  In many ways, this simpler
433 > approach is closer in spirit to the original shifted force method, in
434 > that it projects a neutralizing multipole (and the resulting forces
435 > from this multipole) onto a cutoff sphere. The resulting functional
436 > forms for the potentials, forces, and torques turn out to be quite
437 > similar in form to the Taylor-shifted approach, although the radial
438 > contributions are significantly less perturbed by the gradient-shifted
439 > approach than they are in the Taylor-shifted method.
440 >
441 > For the gradient shifted (GSF) method with the undamped kernel,
442 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
443 > $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
444 > Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
445 > because the Taylor expansion retains only one term, they are
446 > significantly less perturbed than the TSF functions.
447 >
448 > In general, the gradient shifted potential between a central multipole
449 > and any multipolar site inside the cutoff radius is given by,
450 > \begin{equation}
451 >  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
452 >    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
453 >    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
454 > \label{generic2}
455 > \end{equation}
456 > where the sum describes a separate force-shifting that is applied to
457 > each orientational contribution to the energy.  In this expression,
458 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
459 > ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
460 > represent the orientations the multipoles.
461 >
462 > The third term converges more rapidly than the first two terms as a
463 > function of radius, hence the contribution of the third term is very
464 > small for large cutoff radii.  The force and torque derived from
465 > Eq. \ref{generic2} are consistent with the energy expression and
466 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
467 > can be considered generalizations of the original DSF method for
468 > higher order multipole interactions. GSF and TSF are also identical up
469 > to the charge-dipole interaction but generate different expressions in
470 > the energy, force and torque for higher order multipole-multipole
471 > interactions. Complete energy, force, and torque expressions for the
472 > GSF potential are presented in the first paper in this series
473 > (Reference~\onlinecite{PaperI}).
474 >
475 >
476 > \subsection{Shifted potential (SP) }
477 > A discontinuous truncation of the electrostatic potential at the
478 > cutoff sphere introduces a severe artifact (oscillation in the
479 > electrostatic energy) even for molecules with the higher-order
480 > multipoles.\cite{PaperI} We have also formulated an extension of the
481 > Wolf approach for point multipoles by simply projecting the image
482 > multipole onto the surface of the cutoff sphere, and including the
483 > interactions with the central multipole and the image. This
484 > effectively shifts the total potential to zero at the cutoff radius,
485 > \begin{equation}
486 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
487 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
488 > \label{eq:SP}
489 > \end{equation}          
490 > where the sum describes separate potential shifting that is done for
491 > each orientational contribution to the energy (e.g. the direct dipole
492 > product contribution is shifted {\it separately} from the
493 > dipole-distance terms in dipole-dipole interactions).  Note that this
494 > is not a simple shifting of the total potential at $r_c$. Each radial
495 > contribution is shifted separately.  One consequence of this is that
496 > multipoles that reorient after leaving the cutoff sphere can re-enter
497 > the cutoff sphere without perturbing the total energy.
498 >
499 > For the shifted potential (SP) method with the undamped kernel,
500 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
501 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
502 > central multipole and other multipolar sites goes smoothly to zero as
503 > $r \rightarrow r_c$.  However, the force and torque obtained from the
504 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
505 > will still experience energy drift while operating under the SP
506 > potential, but it may be suitable for Monte Carlo approaches where the
507 > configurational energy differences are the primary quantity of
508 > interest.
509 >
510 > \subsection{The Self Term}
511 > In the TSF, GSF, and SP methods, a self-interaction is retained for
512 > the central multipole interacting with its own image on the surface of
513 > the cutoff sphere.  This self interaction is nearly identical with the
514 > self-terms that arise in the Ewald sum for multipoles.  Complete
515 > expressions for the self terms are presented in the first paper in
516 > this series (Reference \onlinecite{PaperI}).
517 >
518 >
519 > \section{\label{sec:methodology}Methodology}
520 >
521 > To understand how the real-space multipole methods behave in computer
522 > simulations, it is vital to test against established methods for
523 > computing electrostatic interactions in periodic systems, and to
524 > evaluate the size and sources of any errors that arise from the
525 > real-space cutoffs.  In the first paper of this series, we compared
526 > the dipolar and quadrupolar energy expressions against analytic
527 > expressions for ordered dipolar and quadrupolar
528 > arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
529 > used the multipolar Ewald sum as a reference method for comparing
530 > energies, forces, and torques for molecular models that mimic
531 > disordered and ordered condensed-phase systems.  The parameters used
532 > in the test cases are given in table~\ref{tab:pars}.
533 >
534 > \begin{table}
535 > \label{tab:pars}
536 > \caption{The parameters used in the systems used to evaluate the new
537 >  real-space methods.  The most comprehensive test was a liquid
538 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
539 >  ions).  This test excercises all orders of the multipolar
540 >  interactions developed in the first paper.}
541 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
542 >             & \multicolumn{2}{c|}{LJ parameters} &
543 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
544 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
545 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
546 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
547 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
548 > \AA\textsuperscript{2})} \\ \hline
549 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
550 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
551 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
552 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
553 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
554 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
555 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
556 > \end{tabularx}
557 > \end{table}
558 > The systems consist of pure multipolar solids (both dipole and
559 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
560 > fluid composed of sites containing both dipoles and quadrupoles
561 > simultaneously, and a final test case that includes ions with point
562 > charges in addition to the multipolar fluid.  The solid-phase
563 > parameters were chosen so that the systems can explore some
564 > orientational freedom for the multipolar sites, while maintaining
565 > relatively strict translational order.  The SSDQ model used here is
566 > not a particularly accurate water model, but it does test
567 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
568 > interactions at roughly the same magnitudes. The last test case, SSDQ
569 > water with dissolved ions, exercises \textit{all} levels of the
570 > multipole-multipole interactions we have derived so far and represents
571 > the most complete test of the new methods.
572 >
573 > In the following section, we present results for the total
574 > electrostatic energy, as well as the electrostatic contributions to
575 > the force and torque on each molecule.  These quantities have been
576 > computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
577 > and have been compared with the values obtained from the multipolar
578 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
579 > between two configurations is the primary quantity that governs how
580 > the simulation proceeds. These differences are the most imporant
581 > indicators of the reliability of a method even if the absolute
582 > energies are not exact.  For each of the multipolar systems listed
583 > above, we have compared the change in electrostatic potential energy
584 > ($\Delta E$) between 250 statistically-independent configurations.  In
585 > molecular dynamics (MD) simulations, the forces and torques govern the
586 > behavior of the simulation, so we also compute the electrostatic
587 > contributions to the forces and torques.
588 >
589 > \subsection{Implementation}
590 > The real-space methods developed in the first paper in this series
591 > have been implemented in our group's open source molecular simulation
592 > program, OpenMD,\cite{openmd} which was used for all calculations in
593 > this work.  The complementary error function can be a relatively slow
594 > function on some processors, so all of the radial functions are
595 > precomputed on a fine grid and are spline-interpolated to provide
596 > values when required.  
597 >
598 > Using the same simulation code, we compare to a multipolar Ewald sum
599 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
600 > the Ewald sum is a re-implementation of the algorithm originally
601 > proposed by Smith that does not use the particle mesh or smoothing
602 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
603 > being compared are the electrostatic contributions to energies, force,
604 > and torques.  All other contributions to these quantities (i.e. from
605 > Lennard-Jones interactions) are removed prior to the comparisons.
606 >
607 > The convergence parameter ($\alpha$) also plays a role in the balance
608 > of the real-space and reciprocal-space portions of the Ewald
609 > calculation.  Typical molecular mechanics packages set this to a value
610 > that depends on the cutoff radius and a tolerance (typically less than
611 > $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
612 > associated with increasing accuracy at the expense of computational
613 > time spent on the reciprocal-space portion of the
614 > summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
615 > 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
616 > Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
617 >
618 > The real-space models have self-interactions that provide
619 > contributions to the energies only.  Although the self interaction is
620 > a rapid calculation, we note that in systems with fluctuating charges
621 > or point polarizabilities, the self-term is not static and must be
622 > recomputed at each time step.
623 >
624 > \subsection{Model systems}
625 > To sample independent configurations of the multipolar crystals, body
626 > centered cubic (bcc) crystals, which exhibit the minimum energy
627 > structures for point dipoles, were generated using 3,456 molecules.
628 > The multipoles were translationally locked in their respective crystal
629 > sites for equilibration at a relatively low temperature (50K) so that
630 > dipoles or quadrupoles could freely explore all accessible
631 > orientations.  The translational constraints were then removed, the
632 > systems were re-equilibrated, and the crystals were simulated for an
633 > additional 10 ps in the microcanonical (NVE) ensemble with an average
634 > temperature of 50 K.  The balance between moments of inertia and
635 > particle mass were chosen to allow orientational sampling without
636 > significant translational motion.  Configurations were sampled at
637 > equal time intervals in order to compare configurational energy
638 > differences.  The crystals were simulated far from the melting point
639 > in order to avoid translational deformation away of the ideal lattice
640 > geometry.
641 >
642 > For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
643 > simulations, each system was created with 2,048 randomly-oriented
644 > molecules.  These were equilibrated at a temperature of 300K for 1 ns.
645 > Each system was then simulated for 1 ns in the microcanonical (NVE)
646 > ensemble.  We collected 250 different configurations at equal time
647 > intervals. For the liquid system that included ionic species, we
648 > converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
649 > \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
650 > run under the same conditions for 1 ns. A total of 250 configurations
651 > were collected. In the following comparisons of energies, forces, and
652 > torques, the Lennard-Jones potentials were turned off and only the
653 > purely electrostatic quantities were compared with the same values
654 > obtained via the Ewald sum.
655 >
656 > \subsection{Accuracy of Energy Differences, Forces and Torques}
657 > The pairwise summation techniques (outlined above) were evaluated for
658 > use in MC simulations by studying the energy differences between
659 > different configurations.  We took the Ewald-computed energy
660 > difference between two conformations to be the correct behavior. An
661 > ideal performance by one of the new methods would reproduce these
662 > energy differences exactly. The configurational energies being used
663 > here contain only contributions from electrostatic interactions.
664 > Lennard-Jones interactions were omitted from the comparison as they
665 > should be identical for all methods.
666 >
667 > Since none of the real-space methods provide exact energy differences,
668 > we used least square regressions analysis for the six different
669 > molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
670 > with the multipolar Ewald reference method.  Unitary results for both
671 > the correlation (slope) and correlation coefficient for these
672 > regressions indicate perfect agreement between the real-space method
673 > and the multipolar Ewald sum.
674 >
675 > Molecular systems were run long enough to explore independent
676 > configurations and 250 configurations were recorded for comparison.
677 > Each system provided 31,125 energy differences for a total of 186,750
678 > data points.  Similarly, the magnitudes of the forces and torques have
679 > also been compared using least squares regression analysis. In the
680 > forces and torques comparison, the magnitudes of the forces acting in
681 > each molecule for each configuration were evaluated. For example, our
682 > dipolar liquid simulation contains 2048 molecules and there are 250
683 > different configurations for each system resulting in 3,072,000 data
684 > points for comparison of forces and torques.
685 >
686 > \subsection{Analysis of vector quantities}
687 > Getting the magnitudes of the force and torque vectors correct is only
688 > part of the issue for carrying out accurate molecular dynamics
689 > simulations.  Because the real space methods reweight the different
690 > orientational contributions to the energies, it is also important to
691 > understand how the methods impact the \textit{directionality} of the
692 > force and torque vectors. Fisher developed a probablity density
693 > function to analyse directional data sets,
694   \begin{equation}
695 < R = \sqrt{(\sum\limits_{i=1}^N \sin\theta_i)^2 + (\sum\limits_{i=1}^N \sin\theta_i)^2}
695 > p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
696 > \label{eq:pdf}
697 > \end{equation}
698 > where $\kappa$ measures directional dispersion of the data around the
699 > mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
700 > estimated as a reciprocal of the circular variance.\cite{Allen91} To
701 > quantify the directional error, forces obtained from the Ewald sum
702 > were taken as the mean (or correct) direction and the angle between
703 > the forces obtained via the Ewald sum and the real-space methods were
704 > evaluated,
705 > \begin{equation}
706 > \cos\theta_i =  \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
707 >  \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
708 > \end{equation}
709 > The total angular displacement of the vectors was calculated as,
710 > \begin{equation}
711 > R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
712   \label{eq:displacement}
713   \end{equation}
714 < where N is number of directional data sets and $theta_i$ are the angles between unit vectors evaluated from the Ewald and the purposed methods. The circular variance is defined as $ Var(\theta) = 1 -R/N$. The value of circular variance varies from 0 to 1. The lower the value of $Var{\theta}$ is higher the value of $\kappa$, which expresses tighter clustering of the direction sets around Ewald direction.
714 > where $N$ is number of force vectors.  The circular variance is
715 > defined as
716 > \begin{equation}
717 > \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
718 > \end{equation}
719 > The circular variance takes on values between from 0 to 1, with 0
720 > indicating a perfect directional match between the Ewald force vectors
721 > and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
722 > correspond to higher values of $\kappa$, which indicates tighter
723 > clustering of the real-space force vectors around the Ewald forces.
724  
725 + A similar analysis was carried out for the electrostatic contribution
726 + to the molecular torques as well as forces.  
727 +
728   \subsection{Energy conservation}
729 < To test conservation of the energy, the mixed molecular system of 2000 dipolar-quadrupolar molecules with 24 $Na^+$,  and 24 $Cl^-$  was run for 1ns in the microcanonical ensemble at 300 K temperature for different cutoff methods (Ewald, Hard, SP, GSF, and TSF). The molecular system was run in 12 parallel computers and started with same initial positions and velocities for all cutoff methods. The slope and Standard Deviation of the energy about the slope (SD) were evaluated in the total energy versus time plot, where the slope evaluates the total energy drift and SD calculates the energy fluctuation in MD simulations. Also, the time duration for the simulation was recorded to compare efficiency of the purposed methods with the Ewald.
729 > To test conservation the energy for the methods, the mixed molecular
730 > system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
731 > ions was run for 1 ns in the microcanonical ensemble at an average
732 > temperature of 300K.  Each of the different electrostatic methods
733 > (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
734 > damping values. The molecular system was started with same initial
735 > positions and velocities for all cutoff methods. The energy drift
736 > ($\delta E_1$) and standard deviation of the energy about the slope
737 > ($\delta E_0$) were evaluated from the total energy of the system as a
738 > function of time.  Although both measures are valuable at
739 > investigating new methods for molecular dynamics, a useful interaction
740 > model must allow for long simulation times with minimal energy drift.
741  
742   \section{\label{sec:result}RESULTS}
743   \subsection{Configurational energy differences}
# Line 180 | Line 760 | To test conservation of the energy, the mixed molecula
760   %       \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
761   %        \caption{}
762        
763 <        \label{fig:barGraph2}
764 <    \end{figure}
765 < %The correlation coefficient ($R^2$) and slope of the linear regression plots for the energy differences for all six different molecular systems is shown in figure 4a and 4b.The plot shows that the correlation coefficient improves for the SP cutoff method as compared to the undamped hard cutoff method in the case of SSDQC, SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar crystal and liquid, the correlation coefficient is almost unchanged and close to 1.  The correlation coefficient is smallest (0.696276 for $r_c$ = 9 $A^o$) for the SSDQC liquid because of the presence of charge-charge and charge-multipole interactions. Since the charge-charge and charge-multipole interaction is long ranged, there is huge deviation of correlation coefficient from 1. Similarly, the quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with compared to interactions in the other multipolar systems, thus the correlation coefficient very close to 1 even for hard cutoff method. The idea of placing image multipole on the surface of the cutoff sphere improves the correlation coefficient and makes it close to 1 for all types of multipolar systems. Similarly the slope is hugely deviated from the correct value for the lower order multipole-multipole interaction and slightly deviated for higher order multipole – multipole interaction. The SP method improves both correlation coefficient ($R^2$) and slope significantly in SSDQC and dipolar systems.  The Slope is found to be deviated more in dipolar crystal as compared to liquid which is associated with the large fluctuation in the electrostatic energy in crystal. The GSF also produced better values of correlation coefficient and slope with the proper selection of the damping alpha (Interested reader can consult accompanying supporting material). The TSF method gives good value of correlation coefficient for the dipolar crystal, dipolar liquid, SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the regression slopes are significantly deviated.
763 > %        \label{fig:barGraph2}
764 > %      \end{figure}
765 > %The correlation coefficient ($R^2$) and slope of the linear
766 > %regression plots for the energy differences for all six different
767 > %molecular systems is shown in figure 4a and 4b.The plot shows that
768 > %the correlation coefficient improves for the SP cutoff method as
769 > %compared to the undamped hard cutoff method in the case of SSDQC,
770 > %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
771 > %crystal and liquid, the correlation coefficient is almost unchanged
772 > %and close to 1.  The correlation coefficient is smallest (0.696276
773 > %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
774 > %charge-charge and charge-multipole interactions. Since the
775 > %charge-charge and charge-multipole interaction is long ranged, there
776 > %is huge deviation of correlation coefficient from 1. Similarly, the
777 > %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
778 > %compared to interactions in the other multipolar systems, thus the
779 > %correlation coefficient very close to 1 even for hard cutoff
780 > %method. The idea of placing image multipole on the surface of the
781 > %cutoff sphere improves the correlation coefficient and makes it close
782 > %to 1 for all types of multipolar systems. Similarly the slope is
783 > %hugely deviated from the correct value for the lower order
784 > %multipole-multipole interaction and slightly deviated for higher
785 > %order multipole – multipole interaction. The SP method improves both
786 > %correlation coefficient ($R^2$) and slope significantly in SSDQC and
787 > %dipolar systems.  The Slope is found to be deviated more in dipolar
788 > %crystal as compared to liquid which is associated with the large
789 > %fluctuation in the electrostatic energy in crystal. The GSF also
790 > %produced better values of correlation coefficient and slope with the
791 > %proper selection of the damping alpha (Interested reader can consult
792 > %accompanying supporting material). The TSF method gives good value of
793 > %correlation coefficient for the dipolar crystal, dipolar liquid,
794 > %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
795 > %regression slopes are significantly deviated.
796 >
797   \begin{figure}
798 <        \centering
799 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
800 <        \caption{The correlation coefficient and regression slope of configurational energy differences for a given method with compared with the reference Ewald method. The value of result equal to 1(dashed line) indicates energy difference is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle)}
801 <        \label{fig:slopeCorr_energy}
802 <    \end{figure}
803 < The combined correlation coefficient and slope for all six systems is shown in Figure ~\ref{fig:slopeCorr_energy}. The correlation coefficient for the undamped hard cutoff method is does not have good agreement with the Ewald because of the fluctuation of the electrostatic energy in the direct truncation method. This deviation in correlation coefficient is improved by using SP, GSF, and TSF method. But the TSF method worsens the regression slope stating that this method produces statistically more biased result as compared to Ewald. Also the GSF method slightly deviate slope but it can be alleviated by using proper value of damping alpha and cutoff radius. The SP method shows good agreement with Ewald method for all values of damping alpha and radii.
798 >  \centering
799 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
800 >  \caption{Statistical analysis of the quality of configurational
801 >    energy differences for the real-space electrostatic methods
802 >    compared with the reference Ewald sum.  Results with a value equal
803 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
804 >    from those obtained using the multipolar Ewald sum.  Different
805 >    values of the cutoff radius are indicated with different symbols
806 >    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
807 >    triangles).}
808 >  \label{fig:slopeCorr_energy}
809 > \end{figure}
810 >
811 > The combined correlation coefficient and slope for all six systems is
812 > shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
813 > reproduce the Ewald configurational energy differences with remarkable
814 > fidelity.  Undamped hard cutoffs introduce a significant amount of
815 > random scatter in the energy differences which is apparent in the
816 > reduced value of the correlation coefficient for this method.  This
817 > can be easily understood as configurations which exhibit small
818 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
819 > will see large energy jumps when hard cutoffs are used.  The
820 > orientations of the multipoles (particularly in the ordered crystals)
821 > mean that these energy jumps can go in either direction, producing a
822 > significant amount of random scatter, but no systematic error.
823 >
824 > The TSF method produces energy differences that are highly correlated
825 > with the Ewald results, but it also introduces a significant
826 > systematic bias in the values of the energies, particularly for
827 > smaller cutoff values. The TSF method alters the distance dependence
828 > of different orientational contributions to the energy in a
829 > non-uniform way, so the size of the cutoff sphere can have a large
830 > effect, particularly for the crystalline systems.
831 >
832 > Both the SP and GSF methods appear to reproduce the Ewald results with
833 > excellent fidelity, particularly for moderate damping ($\alpha =
834 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
835 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
836 > method with short cutoffs, all of the methods would be appropriate for
837 > use in Monte Carlo simulations.
838 >
839   \subsection{Magnitude of the force and torque vectors}
194 The comparison of the magnitude of the combined forces and torques for the data accumulated from all system types are shown in Figure ~\ref{fig:slopeCorr_force}. The correlation and slope for the forces agree with the Ewald even for the hard cutoff method. For the system of molecules with higher order multipoles, the interaction is short ranged. Moreover, the force decays more rapidly than the electrostatic energy hence the hard cutoff method also produces good results. Although the pure cutoff gives the good match of the electrostatic force, the discontinuity in the force at the cutoff radius causes problem in the total energy conservation in MD simulations, which will be discussed in detail in subsection D. The correlation coefficient for GSF method also perfectly matches with Ewald but the slope is slightly deviated (due to extra term obtained from the angular differentiation). This deviation in the slope can be alleviated with proper selection of the damping alpha and radii ($\alpha = 0.2$ and $r_c = 12 A^o$ are good choice). The TSF method shows good agreement in the correlation coefficient but the slope is not good as compared to the Ewald.
195 \begin{figure}
196        \centering
197        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
198        \caption{The correlation coefficient and regression slope of the magnitude of the force for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle). }
199        \label{fig:slopeCorr_force}
200    \end{figure}
201 The torques appears to be very influenced because of extra term generated when the potential energy is modified to get consistent force and torque.  The result shows that the torque from the hard cutoff method has good agreement with Ewald. As the potential is modified to make it consistent with the force and torque, the correlation and slope is deviated as shown in Figure~\ref{fig:slopeCorr_torque} for SP, GSF and TSF cutoff methods.  But the proper value of the damping alpha and radius can improve the agreement of the GSF with the Ewald method. The TSF method shows worst agreement in the slope as compared to Ewald even for larger cutoff radii.
202 \begin{figure}
203        \centering
204        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
205        \caption{The correlation coefficient and regression slope of the magnitude of the torque for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle).}
206        \label{fig:slopeCorr_torque}
207    \end{figure}
208 \subsection{Directionality of the force and torque vectors}  
209 The accurate evaluation of the direction of the force and torques are also important for the dynamic simulation.In our research, the direction data sets were computed from the purposed method and compared with Ewald using Fisher statistics and results are expressed in terms of circular variance ($Var(\theta$).The force and torque vectors from the purposed method followed Fisher probability distribution function expressed in equation~\ref{eq:pdf}. The circular variance for the force and torque vectors of each molecule in the 250 configurations for all system types is shown in Figure~\ref{fig:slopeCorr_circularVariance}. The direction of the force and torque vectors from hard and SP cutoff methods showed best directional agreement with the Ewald. The force and torque vectors from GSF method also showed good agreement with the Ewald method, which can also be improved by varying damping alpha and cutoff radius.For $\alpha = 0.2$ and $r_c = 12 A^o$, $ Var(\theta) $ for direction of the force was found to be 0.002061 and corresponding value of $\kappa $ was 485.20. Integration of equation ~\ref{eq:pdf} for that corresponding value of $\kappa$ showed that 95\% of force vectors are with in $6.37^o$. The TSF method is the poorest in evaluating accurate direction with compared to Hard, SP, and GSF methods. The circular variance for the direction of the torques is larger as compared to force. For same $\alpha = 0.2, r_c = 12 A^o$ and GSF method, the circular variance was 0.01415, which showed 95\% of torque vectors are within $16.75^o$.The direction of the force and torque vectors can be improved by varying $\alpha$ and $r_c$.
840  
841 + The comparisons of the magnitudes of the forces and torques for the
842 + data accumulated from all six systems are shown in Figures
843 + ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
844 + correlation and slope for the forces agree well with the Ewald sum
845 + even for the hard cutoffs.
846 +
847 + For systems of molecules with only multipolar interactions, the pair
848 + energy contributions are quite short ranged.  Moreover, the force
849 + decays more rapidly than the electrostatic energy, hence the hard
850 + cutoff method can also produce reasonable agreement for this quantity.
851 + Although the pure cutoff gives reasonably good electrostatic forces
852 + for pairs of molecules included within each other's cutoff spheres,
853 + the discontinuity in the force at the cutoff radius can potentially
854 + cause energy conservation problems as molecules enter and leave the
855 + cutoff spheres.  This is discussed in detail in section
856 + \ref{sec:conservation}.
857 +
858 + The two shifted-force methods (GSF and TSF) exhibit a small amount of
859 + systematic variation and scatter compared with the Ewald forces.  The
860 + shifted-force models intentionally perturb the forces between pairs of
861 + molecules inside each other's cutoff spheres in order to correct the
862 + energy conservation issues, and this perturbation is evident in the
863 + statistics accumulated for the molecular forces.  The GSF
864 + perturbations are minimal, particularly for moderate damping and
865 + commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
866 + reasonable agreement in the correlation coefficient but again the
867 + systematic error in the forces is concerning if replication of Ewald
868 + forces is desired.
869 +
870   \begin{figure}
871 <        \centering
872 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
873 <        \caption{The circular variance of the data sets of the direction of the  force and torque vectors obtained from a given method about reference Ewald method. The result equal to 0 (dashed line) indicates direction of the vectors are indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle)}
874 <        \label{fig:slopeCorr_circularVariance}
875 <    \end{figure}
876 < \subsection{Total energy conservation}
877 < We have tested the conservation of energy in the SSDQC liquid system by running system for 1ns in the Hard, SP, GSF and TSF method. The Hard cutoff method shows very high energy drifts 433.53 KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol (measured by the SD from the slope) for the undamped case, which makes it completely unusable in MD simulations. The SP method also shows large value of energy drift 1.289 Kcal/Mol/ns/particle and energy fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method is due to the non-vanishing nature of the torque and force at the cutoff radius. We can improve the energy conservation in some extent by the proper selection of the damping alpha but the improvement is not good enough, which can be observed in Figure 9a and 9b .The GSF and TSF shows very low value of energy drift 0.09016, 0.07371 KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol respectively for the undamped case. Since the absolute value of the evaluated electrostatic energy, force and torque from TSF method are deviated from the Ewald, it does not mimic MD simulations appropriately. The electrostatic energy, force and torque from the GSF method have very good agreement with the Ewald. In addition, the energy drift and energy fluctuation from the GSF method is much better than Ewald’s method for reciprocal space vector value ($k_f$) equal to 7 as shown in Figure~\ref{fig:energyDrift} and ~\ref{fig:fluctuation}. We can improve the total energy fluctuation and drift for the Ewald’s method by increasing size of the reciprocal space, which extremely increseses the simulation time. In our current simulation, the simulation time for the Hard, SP, and GSF methods are about 5.5 times faster than the Ewald method.
871 >  \centering
872 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
873 >  \caption{Statistical analysis of the quality of the force vector
874 >    magnitudes for the real-space electrostatic methods compared with
875 >    the reference Ewald sum. Results with a value equal to 1 (dashed
876 >    line) indicate force magnitude values indistinguishable from those
877 >    obtained using the multipolar Ewald sum.  Different values of the
878 >    cutoff radius are indicated with different symbols (9\AA\ =
879 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
880 >  \label{fig:slopeCorr_force}
881 > \end{figure}
882 >
883 >
884   \begin{figure}
885 <        \centering
886 <        \includegraphics[width=0.5 \textwidth]{log(energyDrift)-crop.pdf}
887 < \label{fig:energyDrift}        
888 <        \end{figure}
885 >  \centering
886 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
887 >  \caption{Statistical analysis of the quality of the torque vector
888 >    magnitudes for the real-space electrostatic methods compared with
889 >    the reference Ewald sum. Results with a value equal to 1 (dashed
890 >    line) indicate force magnitude values indistinguishable from those
891 >    obtained using the multipolar Ewald sum.  Different values of the
892 >    cutoff radius are indicated with different symbols (9\AA\ =
893 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
894 >  \label{fig:slopeCorr_torque}
895 > \end{figure}
896 >
897 > The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
898 > significantly influenced by the choice of real-space method.  The
899 > torque expressions have the same distance dependence as the energies,
900 > which are naturally longer-ranged expressions than the inter-site
901 > forces.  Torques are also quite sensitive to orientations of
902 > neighboring molecules, even those that are near the cutoff distance.
903 >
904 > The results shows that the torque from the hard cutoff method
905 > reproduces the torques in quite good agreement with the Ewald sum.
906 > The other real-space methods can cause some deviations, but excellent
907 > agreement with the Ewald sum torques is recovered at moderate values
908 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
909 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
910 > in the slope when compared with the Ewald torques even for larger
911 > cutoff radii.  It appears that the severity of the perturbations in
912 > the TSF method are most in evidence for the torques.
913 >
914 > \subsection{Directionality of the force and torque vectors}  
915 >
916 > The accurate evaluation of force and torque directions is just as
917 > important for molecular dynamics simulations as the magnitudes of
918 > these quantities. Force and torque vectors for all six systems were
919 > analyzed using Fisher statistics, and the quality of the vector
920 > directionality is shown in terms of circular variance
921 > ($\mathrm{Var}(\theta)$) in figure
922 > \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
923 > from the new real-space methods exhibit nearly-ideal Fisher probability
924 > distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
925 > exhibit the best vectorial agreement with the Ewald sum. The force and
926 > torque vectors from GSF method also show good agreement with the Ewald
927 > method, which can also be systematically improved by using moderate
928 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
929 > 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
930 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
931 > the corresponding Ewald forces. The TSF method produces the poorest
932 > agreement with the Ewald force directions.
933 >
934 > Torques are again more perturbed than the forces by the new real-space
935 > methods, but even here the variance is reasonably small.  For the same
936 > method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
937 > the circular variance was 0.01415, corresponds to a distribution which
938 > has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
939 > results. Again, the direction of the force and torque vectors can be
940 > systematically improved by varying $\alpha$ and $r_c$.
941 >
942   \begin{figure}
943 <        \centering
944 <        \includegraphics[width=0.5 \textwidth]{logSD-crop.pdf}      
945 <        \caption{The plot showing (a) standard deviation, and (b) total energy drift in the total energy conservation plot for different values of the damping alpha for different cut off methods. }
946 <        \label{fig:fluctuation}
947 <    \end{figure}
943 >  \centering
944 >  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
945 >  \caption{The circular variance of the direction of the force and
946 >    torque vectors obtained from the real-space methods around the
947 >    reference Ewald vectors. A variance equal to 0 (dashed line)
948 >    indicates direction of the force or torque vectors are
949 >    indistinguishable from those obtained from the Ewald sum. Here
950 >    different symbols represent different values of the cutoff radius
951 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
952 >  \label{fig:slopeCorr_circularVariance}
953 > \end{figure}
954 >
955 > \subsection{Energy conservation\label{sec:conservation}}
956 >
957 > We have tested the conservation of energy one can expect to see with
958 > the new real-space methods using the SSDQ water model with a small
959 > fraction of solvated ions. This is a test system which exercises all
960 > orders of multipole-multipole interactions derived in the first paper
961 > in this series and provides the most comprehensive test of the new
962 > methods.  A liquid-phase system was created with 2000 water molecules
963 > and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
964 > temperature of 300K.  After equilibration, this liquid-phase system
965 > was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
966 > a cutoff radius of 12\AA.  The value of the damping coefficient was
967 > also varied from the undamped case ($\alpha = 0$) to a heavily damped
968 > case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
969 > sample was also run using the multipolar Ewald sum with the same
970 > real-space cutoff.
971 >
972 > In figure~\ref{fig:energyDrift} we show the both the linear drift in
973 > energy over time, $\delta E_1$, and the standard deviation of energy
974 > fluctuations around this drift $\delta E_0$.  Both of the
975 > shifted-force methods (GSF and TSF) provide excellent energy
976 > conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
977 > while the hard cutoff is essentially unusable for molecular dynamics.
978 > SP provides some benefit over the hard cutoff because the energetic
979 > jumps that happen as particles leave and enter the cutoff sphere are
980 > somewhat reduced, but like the Wolf method for charges, the SP method
981 > would not be as useful for molecular dynamics as either of the
982 > shifted-force methods.
983 >
984 > We note that for all tested values of the cutoff radius, the new
985 > real-space methods can provide better energy conservation behavior
986 > than the multipolar Ewald sum, even when utilizing a relatively large
987 > $k$-space cutoff values.
988 >
989 > \begin{figure}
990 >  \centering
991 >  \includegraphics[width=\textwidth]{newDrift_12.pdf}
992 > \label{fig:energyDrift}        
993 > \caption{Analysis of the energy conservation of the real-space
994 >  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
995 >  energy over time (in kcal / mol / particle / ns) and $\delta
996 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
997 >  around this drift (in kcal / mol / particle).  All simulations were
998 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
999 >  300 K starting from the same initial configuration. All runs
1000 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
1001 > \end{figure}
1002 >
1003 >
1004   \section{CONCLUSION}
1005 < We have generalized the charged neutralized potential energy originally developed by the Wolf et al.\cite{Wolf99} for the charge-charge interaction to the charge-multipole and multipole-multipole interaction in the SP method for higher order multipoles. Also, we have developed GSF and TSF methods by implementing the modification purposed by Fennel and Gezelter\cite{Gezelter06} for the charge-charge interaction to the higher order multipoles to ensure consistency and smooth truncation of the electrostatic energy, force, and torque for the spherical truncation. The SP methods for multipoles proved its suitability in MC simulations. On the other hand, the results from the GSF method produced good agreement with the Ewald's energy, force, and torque. Also, it shows very good energy conservation in MD simulations.
1006 < The direct truncation of any molecular system without multipole neutralization creates the fluctuation in the electrostatic energy. This fluctuation in the energy is very large for the case of crystal because of long range of multipole ordering (Refer paper I).\cite{PaperI} This is also significant in the case of the liquid because of the local multipole ordering in the molecules. If the net multipole within cutoff radius neutralized within cutoff sphere by placing image multiples on the surface of the sphere, this fluctuation in the energy reduced significantly. Also, the multipole neutralization in the generalized SP method showed very good agreement with the Ewald as compared to direct truncation for the evaluation of the $\triangle E$ between the configurations.
1007 < In MD simulations, the energy conservation is very important. The conservation of the total energy can be ensured by  i) enforcing the smooth truncation of the energy, force and torque in the cutoff radius and ii) making the energy, force and torque consistent with each other. The GSF and TSF methods ensure the consistency and smooth truncation of the energy, force and torque at the cutoff radius, as a result show very good total energy conservation. But the TSF method does not show good agreement in the absolute value of the electrostatic energy, force and torque with the Ewald.  The GSF method has mimicked Ewald’s force, energy and torque accurately and also conserved energy. Therefore, the GSF method is the suitable method for evaluating required force field in MD simulations. In addition, the energy drift and fluctuation from the GSF method is much better than Ewald’s method for finite-sized reciprocal space.
1008 < \bibliographystyle{rev4-1}
1005 > In the first paper in this series, we generalized the
1006 > charge-neutralized electrostatic energy originally developed by Wolf
1007 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1008 > up to quadrupolar order.  The SP method is essentially a
1009 > multipole-capable version of the Wolf model.  The SP method for
1010 > multipoles provides excellent agreement with Ewald-derived energies,
1011 > forces and torques, and is suitable for Monte Carlo simulations,
1012 > although the forces and torques retain discontinuities at the cutoff
1013 > distance that prevents its use in molecular dynamics.
1014 >
1015 > We also developed two natural extensions of the damped shifted-force
1016 > (DSF) model originally proposed by Fennel and
1017 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1018 > smooth truncation of energies, forces, and torques at the real-space
1019 > cutoff, and both converge to DSF electrostatics for point-charge
1020 > interactions.  The TSF model is based on a high-order truncated Taylor
1021 > expansion which can be relatively perturbative inside the cutoff
1022 > sphere.  The GSF model takes the gradient from an images of the
1023 > interacting multipole that has been projected onto the cutoff sphere
1024 > to derive shifted force and torque expressions, and is a significantly
1025 > more gentle approach.
1026 >
1027 > Of the two newly-developed shifted force models, the GSF method
1028 > produced quantitative agreement with Ewald energy, force, and torques.
1029 > It also performs well in conserving energy in MD simulations.  The
1030 > Taylor-shifted (TSF) model provides smooth dynamics, but these take
1031 > place on a potential energy surface that is significantly perturbed
1032 > from Ewald-based electrostatics.  
1033 >
1034 > % The direct truncation of any electrostatic potential energy without
1035 > % multipole neutralization creates large fluctuations in molecular
1036 > % simulations.  This fluctuation in the energy is very large for the case
1037 > % of crystal because of long range of multipole ordering (Refer paper
1038 > % I).\cite{PaperI} This is also significant in the case of the liquid
1039 > % because of the local multipole ordering in the molecules. If the net
1040 > % multipole within cutoff radius neutralized within cutoff sphere by
1041 > % placing image multiples on the surface of the sphere, this fluctuation
1042 > % in the energy reduced significantly. Also, the multipole
1043 > % neutralization in the generalized SP method showed very good agreement
1044 > % with the Ewald as compared to direct truncation for the evaluation of
1045 > % the $\triangle E$ between the configurations.  In MD simulations, the
1046 > % energy conservation is very important. The conservation of the total
1047 > % energy can be ensured by i) enforcing the smooth truncation of the
1048 > % energy, force and torque in the cutoff radius and ii) making the
1049 > % energy, force and torque consistent with each other. The GSF and TSF
1050 > % methods ensure the consistency and smooth truncation of the energy,
1051 > % force and torque at the cutoff radius, as a result show very good
1052 > % total energy conservation. But the TSF method does not show good
1053 > % agreement in the absolute value of the electrostatic energy, force and
1054 > % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1055 > % energy and torque accurately and also conserved energy.
1056 >
1057 > The only cases we have found where the new GSF and SP real-space
1058 > methods can be problematic are those which retain a bulk dipole moment
1059 > at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1060 > materials, uniform weighting of the orientational contributions can be
1061 > important for converging the total energy.  In these cases, the
1062 > damping function which causes the non-uniform weighting can be
1063 > replaced by the bare electrostatic kernel, and the energies return to
1064 > the expected converged values.
1065 >
1066 > Based on the results of this work, the GSF method is a suitable and
1067 > efficient replacement for the Ewald sum for evaluating electrostatic
1068 > interactions in MD simulations.  Both methods retain excellent
1069 > fidelity to the Ewald energies, forces and torques.  Additionally, the
1070 > energy drift and fluctuations from the GSF electrostatics are better
1071 > than a multipolar Ewald sum for finite-sized reciprocal spaces.
1072 > Because they use real-space cutoffs with moderate cutoff radii, the
1073 > GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1074 > increases.  Additionally, they can be made extremely efficient using
1075 > spline interpolations of the radial functions.  They require no
1076 > Fourier transforms or $k$-space sums, and guarantee the smooth
1077 > handling of energies, forces, and torques as multipoles cross the
1078 > real-space cutoff boundary.
1079 >
1080 > \begin{acknowledgments}
1081 >  JDG acknowledges helpful discussions with Christopher
1082 >  Fennell. Support for this project was provided by the National
1083 >  Science Foundation under grant CHE-1362211. Computational time was
1084 >  provided by the Center for Research Computing (CRC) at the
1085 >  University of Notre Dame.
1086 > \end{acknowledgments}
1087 >
1088 > %\bibliographystyle{aip}
1089 > \newpage
1090   \bibliography{references}
1091   \end{document}
1092  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines