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 4171 by gezelter, Wed Jun 4 19:31:06 2014 UTC vs.
Revision 4214 by gezelter, Wed Sep 10 21:06:56 2014 UTC

# Line 35 | Line 35 | preprint,
35   %\linenumbers\relax % Commence numbering lines
36   \usepackage{amsmath}
37   \usepackage{times}
38 < \usepackage{mathptm}
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. Comparison of Simulation Methodologies} % Force line breaks with \\
50 > \title{Real space electrostatics for multipoles. II. Comparisons with
51 >  the Ewald Sum}
52  
53   \author{Madan Lamichhane}
54 < \affiliation{Department of Physics, University
53 < of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
54 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
55  
56   \author{Kathie E. Newman}
57 < \affiliation{Department of Physics, University
57 < of Notre Dame, Notre Dame, IN 46556}
57 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
58  
59   \author{J. Daniel Gezelter}%
60   \email{gezelter@nd.edu.}
61 < \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
62 < }%
61 > \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
62 > }
63  
64 < \date{\today}% It is always \today, today,
65 <             %  but any date may be explicitly specified
64 > \date{\today}
65  
66   \begin{abstract}
67 < 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.
67 >  We report on tests of the shifted potential (SP), gradient shifted
68 >  force (GSF), and Taylor shifted force (TSF) real-space methods for
69 >  multipole interactions developed in the first paper in this series,
70 >  using the multipolar Ewald sum as a reference method. The tests were
71 >  carried out in a variety of condensed-phase environments designed to
72 >  test up to quadrupole-quadrupole interactions.  Comparisons of the
73 >  energy differences between configurations, molecular forces, and
74 >  torques were used to analyze how well the real-space models perform
75 >  relative to the more computationally expensive Ewald treatment.  We
76 >  have also investigated the energy conservation, structural, and
77 >  dynamical properties of the new methods in molecular dynamics
78 >  simulations. The SP method shows excellent agreement with
79 >  configurational energy differences, forces, and torques, and would
80 >  be suitable for use in Monte Carlo calculations.  Of the two new
81 >  shifted-force methods, the GSF approach shows the best agreement
82 >  with Ewald-derived energies, forces, and torques and also exhibits
83 >  energy conservation properties that make it an excellent choice for
84 >  efficient computation of electrostatic interactions in molecular
85 >  dynamics simulations.  Both SP and GSF are able to reproduce
86 >  structural and dynamical properties in the liquid models with
87 >  excellent fidelity.
88   \end{abstract}
89  
90 < \pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
90 > %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
91                               % Classification Scheme.
92 < \keywords{Electrostatics, Multipoles, Real-space}
92 > %\keywords{Electrostatics, Multipoles, Real-space}
93  
94   \maketitle
95  
77
96   \section{\label{sec:intro}Introduction}
97   Computing the interactions between electrostatic sites is one of the
98 < most expensive aspects of molecular simulations, which is why there
99 < have been significant efforts to develop practical, efficient and
100 < convergent methods for handling these interactions. Ewald's method is
101 < perhaps the best known and most accurate method for evaluating
102 < energies, forces, and torques in explicitly-periodic simulation
103 < cells. In this approach, the conditionally convergent electrostatic
104 < energy is converted into two absolutely convergent contributions, one
105 < which is carried out in real space with a cutoff radius, and one in
106 < reciprocal space.\cite{Clarke:1986eu,Woodcock75}
98 > most expensive aspects of molecular simulations. There have been
99 > significant efforts to develop practical, efficient and convergent
100 > methods for handling these interactions. Ewald's method is perhaps the
101 > best known and most accurate method for evaluating energies, forces,
102 > and torques in explicitly-periodic simulation cells. In this approach,
103 > the conditionally convergent electrostatic energy is converted into
104 > two absolutely convergent contributions, one which is carried out in
105 > real space with a cutoff radius, and one in reciprocal
106 > space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
107  
108   When carried out as originally formulated, the reciprocal-space
109   portion of the Ewald sum exhibits relatively poor computational
110 < scaling, making it prohibitive for large systems. By utilizing
111 < particle meshes and three dimensional fast Fourier transforms (FFT),
112 < the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
113 < (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
114 < the computational cost from $O(N^2)$ down to $O(N \log
115 < N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
110 > scaling, making it prohibitive for large systems. By utilizing a
111 > particle mesh and three dimensional fast Fourier transforms (FFT), the
112 > particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
113 > (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
114 > methods can decrease the computational cost from $O(N^2)$ down to $O(N
115 > \log
116 > N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
117  
118 < Because of the artificial periodicity required for the Ewald sum, the
100 < method may require modification to compute interactions for
118 > Because of the artificial periodicity required for the Ewald sum,
119   interfacial molecular systems such as membranes and liquid-vapor
120 < interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
121 < To simulate interfacial systems, Parry’s extension of the 3D Ewald sum
122 < is appropriate for slab geometries.\cite{Parry:1975if} The inherent
123 < periodicity in the Ewald’s method can also be problematic for
124 < interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
125 < methods that were developed to handle two-dimensional (2D)
126 < electrostatic interactions in interfacial systems have not had similar
127 < particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
128 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
120 > interfaces require modifications to the method.  Parry's extension of
121 > the three dimensional Ewald sum is appropriate for slab
122 > geometries.\cite{Parry:1975if} Modified Ewald methods that were
123 > developed to handle two-dimensional (2-D) electrostatic
124 > interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
125 > These methods were originally quite computationally
126 > expensive.\cite{Spohr97,Yeh99} There have been several successful
127 > efforts that reduced the computational cost of 2-D lattice summations,
128 > bringing them more in line with the scaling for the full 3-D
129 > treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
130 > inherent periodicity required by the Ewald method can also be
131 > problematic in a number of protein/solvent and ionic solution
132 > environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
133  
134   \subsection{Real-space methods}
135   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
136   method for calculating electrostatic interactions between point
137 < charges. They argued that the effective Coulomb interaction in
138 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
139 < an ordered lattice (e.g. when computing the Madelung constant of an
140 < ionic solid), the material can be considered as a set of ions
141 < interacting with neutral dipolar or quadrupolar ``molecules'' giving
142 < an effective distance dependence for the electrostatic interactions of
143 < $r^{-5}$ (see figure \ref{fig:NaCl}.  For this reason, careful
144 < applications of Wolf's method are able to obtain accurate estimates of
145 < Madelung constants using relatively short cutoff radii.  Recently,
146 < Fukuda used neutralization of the higher order moments for the
147 < calculation of the electrostatic interaction of the point charges
148 < system.\cite{Fukuda:2013sf}
137 > charges. They argued that the effective Coulomb interaction in most
138 > condensed phase systems is effectively short
139 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
140 > computing the Madelung constant of an ionic solid), the material can
141 > be considered as a set of ions interacting with neutral dipolar or
142 > quadrupolar ``molecules'' giving an effective distance dependence for
143 > the electrostatic interactions of $r^{-5}$ (see figure
144 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
145 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
146 > electrostatic energy per ion converges more rapidly to the Madelung
147 > energy than the dipolar approximation.\cite{Wolf92} To find the
148 > correct Madelung constant, Lacman suggested that the NaCl structure
149 > could be constructed in a way that the finite crystal terminates with
150 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
151 > what is effectively a set of octupoles at large distances. These facts
152 > suggest that the Madelung constants are relatively short ranged for
153 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
154 > application of Wolf's method can provide accurate estimates of
155 > Madelung constants using relatively short cutoff radii.
156  
157 < \begin{figure}[h!]
157 > Direct truncation of interactions at a cutoff radius creates numerical
158 > errors.  Wolf \textit{et al.} suggest that truncation errors are due
159 > to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
160 > neutralize this charge they proposed placing an image charge on the
161 > surface of the cutoff sphere for every real charge inside the cutoff.
162 > These charges are present for the evaluation of both the pair
163 > interaction energy and the force, although the force expression
164 > maintains a discontinuity at the cutoff sphere.  In the original Wolf
165 > formulation, the total energy for the charge and image were not equal
166 > to the integral of the force expression, and as a result, the total
167 > energy would not be conserved in molecular dynamics (MD)
168 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennell and
169 > Gezelter later proposed shifted force variants of the Wolf method with
170 > commensurate force and energy expressions that do not exhibit this
171 > problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
172 > were also proposed by Chen \textit{et
173 >  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
174 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfully
175 > used additional neutralization of higher order moments for systems of
176 > point charges.\cite{Fukuda:2013sf}
177 >
178 > \begin{figure}
179    \centering
180 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
181 <  \caption{Top: NaCl crystal showing how spherical truncation can
182 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
183 <    molecules interact with the central ion.  Bottom: A dipolar
184 <    crystal exhibiting similar behavior and illustrating how the
185 <    effective dipole-octupole interactions can be disrupted by
186 <    spherical truncation.}
187 <  \label{fig:NaCl}
180 >  \includegraphics[width=\linewidth]{schematic.eps}
181 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
182 >    charges (in the smaller grey circle), so interactions are
183 >    effectively charge-multipole at longer distances.  With hard
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. \label{fig:schematic}}
192   \end{figure}
193  
194 < The direct truncation of interactions at a cutoff radius creates
195 < truncation defects. Wolf \textit{et al.} further argued that
196 < truncation errors are due to net charge remaining inside the cutoff
197 < sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
198 < placing an image charge on the surface of the cutoff sphere for every
199 < real charge inside the cutoff.  These charges are present for the
200 < evaluation of both the pair interaction energy and the force, although
201 < the force expression maintained a discontinuity at the cutoff sphere.
202 < In the original Wolf formulation, the total energy for the charge and
203 < image were not equal to the integral of their force expression, and as
150 < a result, the total energy would not be conserved in molecular
151 < dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
152 < Fennel and Gezelter later proposed shifted force variants of the Wolf
153 < method with commensurate force and energy expressions that do not
154 < exhibit this problem.\cite{Fennell:2006lq}   Related real-space
155 < methods were also proposed by Chen \textit{et
156 <  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
157 < and by Wu and Brooks.\cite{Wu:044107}
158 <
159 < Considering the interaction of one central ion in an ionic crystal
160 < with a portion of the crystal at some distance, the effective Columbic
161 < potential is found to be decreasing as $r^{-5}$. If one views the
162 < \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
163 < \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
164 < rapidly to the Madelung energy than the dipolar
165 < approximation.\cite{Wolf92} To find the correct Madelung constant,
166 < Lacman suggested that the NaCl structure could be constructed in a way
167 < that the finite crystal terminates with complete \ce{(NaCl)4}
168 < molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
169 < by opposite charges. Similarly for each pair of charges, there is an
170 < opposite pair of charge adjacent to it.  The central ion sees what is
171 < effectively a set of octupoles at large distances. These facts suggest
172 < that the Madelung constants are relatively short ranged for perfect
173 < ionic crystals.\cite{Wolf:1999dn}
174 <
175 < One can make a similar argument for crystals of point multipoles. The
176 < Luttinger and Tisza treatment of energy constants for dipolar lattices
177 < utilizes 24 basis vectors that contain dipoles at the eight corners of
178 < a unit cube.  Only three of these basis vectors, $X_1, Y_1,
179 < \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
180 < zero net dipole and retain contributions only from higher order
181 < multipoles.  The effective interaction between a dipole at the center
194 > One can make a similar effective range argument for crystals of point
195 > \textit{multipoles}. The Luttinger and Tisza treatment of energy
196 > constants for dipolar lattices utilizes 24 basis vectors that contain
197 > dipoles at the eight corners of a unit cube.\cite{LT} Only three of
198 > these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
199 > moments, while the rest have zero net dipole and retain contributions
200 > only from higher order multipoles.  The lowest-energy crystalline
201 > structures are built out of basis vectors that have only residual
202 > quadrupolar moments (e.g. the $Z_5$ array). In these low energy
203 > structures, the effective interaction between a dipole at the center
204   of a crystal and a group of eight dipoles farther away is
205   significantly shorter ranged than the $r^{-3}$ that one would expect
206   for raw dipole-dipole interactions.  Only in crystals which retain a
# Line 188 | Line 210 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
210   unstable.
211  
212   In ionic crystals, real-space truncation can break the effective
213 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
214 < swings in the electrostatic energy as the cutoff radius is increased
215 < (or as individual ions move back and forth across the boundary).  This
216 < is why the image charges were necessary for the Wolf sum to exhibit
217 < rapid convergence.  Similarly, the real-space truncation of point
218 < multipole interactions breaks higher order multipole arrangements, and
219 < image multipoles are required for real-space treatments of
198 < electrostatic energies.
213 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
214 > significant swings in the electrostatic energy as individual ions move
215 > back and forth across the boundary.  This is why the image charges are
216 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
217 > the real-space truncation of point multipole interactions breaks
218 > higher order multipole arrangements, and image multipoles are required
219 > for real-space treatments of electrostatic energies.
220  
221 < % Because of this reason, although the nature of electrostatic
222 < % interaction short ranged, the hard cutoff sphere creates very large
223 < % fluctuation in the electrostatic energy for the perfect crystal. In
224 < % addition, the charge neutralized potential proposed by Wolf et
225 < % al. converged to correct Madelung constant but still holds oscillation
226 < % in the energy about correct Madelung energy.\cite{Wolf:1999dn}.  This
227 < % oscillation in the energy around its fully converged value can be due
228 < % to the non-neutralized value of the higher order moments within the
208 < % cutoff sphere.
221 > The shorter effective range of electrostatic interactions is not
222 > limited to perfect crystals, but can also apply in disordered fluids.
223 > Even at elevated temperatures, there is local charge balance in an
224 > ionic liquid, where each positive ion has surroundings dominated by
225 > negative ions and vice versa.  The reversed-charge images on the
226 > cutoff sphere that are integral to the Wolf and damped shifted force
227 > (DSF) approaches retain the effective multipolar interactions as the
228 > charges traverse the cutoff boundary.
229  
230 < The forces and torques acting on atomic sites are the fundamental
231 < factors driving dynamics in molecular simulations. Fennell and
232 < Gezelter proposed the damped shifted force (DSF) energy kernel to
233 < obtain consistent energies and forces on the atoms within the cutoff
230 > In multipolar fluids (see Fig. \ref{fig:schematic}) there is
231 > significant orientational averaging that additionally reduces the
232 > effect of long-range multipolar interactions.  The image multipoles
233 > that are introduced in the Taylor shifted force (TSF), gradient
234 > shifted force (GSF), and shifted potential (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 > Forces and torques acting on atomic sites are fundamental in driving
239 > dynamics in molecular simulations, and the DSF energy kernel provides
240 > consistent energies and forces on charged atoms within the cutoff
241   sphere. Both the energy and the force go smoothly to zero as an atom
242 < aproaches the cutoff radius. The comparisons of the accuracy these
242 > approaches the cutoff radius. The comparisons of the accuracy these
243   quantities between the DSF kernel and SPME was surprisingly
244 < good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
245 < calculating electrostatic interactions in molecular systems with
246 < relatively uniform charge
220 < densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
244 > good.\cite{Fennell:2006lq} As a result, the DSF method has seen
245 > increasing use in molecular systems with relatively uniform charge
246 > densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
247  
248   \subsection{The damping function}
249 < The damping function used in our research has been discussed in detail
250 < in the first paper of this series.\cite{PaperI} The radial kernel
251 < $1/r$ for the interactions between point charges can be replaced by
252 < the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
253 < accelerate the rate of convergence, where $\alpha$ is a damping
254 < parameter with units of inverse distance.  Altering the value of
255 < $\alpha$ is equivalent to changing the width of Gaussian charge
256 < distributions that replace each point charge -- Gaussian overlap
257 < integrals yield complementary error functions when truncated at a
258 < finite distance.
249 > The damping function has been discussed in detail in the first paper
250 > of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
251 > interactions between point charges can be replaced by the
252 > complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
253 > convergence, where $\alpha$ is a damping parameter with units of
254 > inverse distance.  Altering the value of $\alpha$ is equivalent to
255 > changing the width of Gaussian charge distributions that replace each
256 > point charge, as Coulomb integrals with Gaussian charge distributions
257 > produce complementary error functions when truncated at a finite
258 > distance.
259  
260 < By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
261 < cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
262 < agreement with SPME for the interaction energies, forces and torques
263 < for charge-charge interactions.\cite{Fennell:2006lq}
260 > With moderate damping coefficients, $\alpha \sim 0.2$ \AA$^{-1}$, the DSF method
261 > produced very good agreement with SPME for interaction energies,
262 > forces and torques for charge-charge
263 > interactions.\cite{Fennell:2006lq}
264  
265   \subsection{Point multipoles in molecular modeling}
266   Coarse-graining approaches which treat entire molecular subsystems as
267   a single rigid body are now widely used. A common feature of many
268   coarse-graining approaches is simplification of the electrostatic
269   interactions between bodies so that fewer site-site interactions are
270 < required to compute configurational energies.  Many coarse-grained
271 < molecular structures would normally consist of equal positive and
246 < negative charges, and rather than use multiple site-site interactions,
247 < the interaction between higher order multipoles can also be used to
248 < evaluate a single molecule-molecule
249 < interaction.\cite{Ren06,Essex10,Essex11}
270 > required to compute configurational
271 > energies.\cite{Ren06,Essex10,Essex11}
272  
273 < Because electrons in a molecule are not localized at specific points,
274 < the assignment of partial charges to atomic centers is a relatively
275 < rough approximation.  Atomic sites can also be assigned point
276 < multipoles and polarizabilities to increase the accuracy of the
277 < molecular model.  Recently, water has been modeled with point
278 < multipoles up to octupolar
279 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
273 > Additionally, because electrons in a molecule are not localized at
274 > specific points, the assignment of partial charges to atomic centers
275 > is always an approximation.  For increased accuracy, atomic sites can
276 > also be assigned point multipoles and polarizabilities.  Recently,
277 > water has been modeled with point multipoles up to octupolar order
278 > using the soft sticky dipole-quadrupole-octupole (SSDQO)
279 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
280   multipoles up to quadrupolar order have also been coupled with point
281   polarizabilities in the high-quality AMOEBA and iAMOEBA water
282 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
283 < using point multipole with the real space truncation without
284 < accounting for multipolar neutrality will create energy conservation
285 < issues in molecular dynamics (MD) simulations.
282 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
283 > truncating point multipoles without smoothing the forces and torques
284 > can create energy conservation issues in molecular dynamics
285 > simulations.
286  
287   In this paper we test a set of real-space methods that were developed
288   for point multipolar interactions.  These methods extend the damped
289   shifted force (DSF) and Wolf methods originally developed for
290   charge-charge interactions and generalize them for higher order
291 < multipoles. The detailed mathematical development of these methods has
292 < been presented in the first paper in this series, while this work
293 < covers the testing the energies, forces, torques, and energy
291 > multipoles.  The detailed mathematical development of these methods
292 > has been presented in the first paper in this series, while this work
293 > covers the testing of energies, forces, torques, and energy
294   conservation properties of the methods in realistic simulation
295   environments.  In all cases, the methods are compared with the
296 < reference method, a full multipolar Ewald treatment.
296 > reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
297  
298  
277 %\subsection{Conservation of total energy }
278 %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.
279
299   \section{\label{sec:method}Review of Methods}
300   Any real-space electrostatic method that is suitable for MD
301   simulations should have the electrostatic energy, forces and torques
302   between two sites go smoothly to zero as the distance between the
303 < sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$.  Requiring
303 > sites, $r_{ab}$ approaches the cutoff radius, $r_c$.  Requiring
304   this continuity at the cutoff is essential for energy conservation in
305   MD simulations.  The mathematical details of the shifted potential
306   (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
# Line 295 | Line 314 | U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1
314   expressed as the product of two multipole operators and a Coulombic
315   kernel,
316   \begin{equation}
317 < U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
317 > U_{ab}(r)= M_{a} M_{b} \frac{1}{r}  \label{kernel}.
318   \end{equation}
319 < where the multipole operator for site $\bf a$,
320 < \begin{equation}
321 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
303 < +  Q_{{\bf a}\alpha\beta}
304 < \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
305 < \end{equation}
306 < is expressed in terms of the point charge, $C_{\bf a}$, dipole,
307 < $D_{{\bf a}\alpha}$, and quadrupole, $Q_{{\bf a}\alpha\beta}$, for
308 < object $\bf a$.  Note that in this work, we use the primitive
309 < quadrupole tensor, $Q_{a\alpha,\beta}=\frac{1}{2}\sum_{k\;in\;a}q_k
310 < r_{k\alpha}r_{k\beta}$ to represent point quadrupoles on a site.
319 > where the multipole operator for site $a$, $M_{a}$, is
320 > expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
321 > $a$, etc.
322  
323 < Interactions between multipoles can be expressed as higher derivatives
324 < of the bare Coulomb potential, so one way of ensuring that the forces
314 < and torques vanish at the cutoff distance is to include a larger
315 < number of terms in the truncated Taylor expansion, e.g.,
316 < %
323 > The TSF potential for any multipole-multipole interaction can be
324 > written
325   \begin{equation}
318 f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
319 \end{equation}
320 %
321 The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
322 Thus, for $f(r)=1/r$, we find
323 %
324 \begin{equation}
325 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} .
326 \end{equation}
327 This function is an approximate electrostatic potential that has
328 vanishing second derivatives at the cutoff radius, making it suitable
329 for shifting the forces and torques of charge-dipole interactions.
330
331 In general, the TSF potential for any multipole-multipole interaction
332 can be written
333 \begin{equation}
326   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
327   \label{generic}
328   \end{equation}
329 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
330 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
331 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
332 < energy, force, and torques, the required number of terms from Taylor
333 < series expansion in $f_n(r)$ must be performed for different
334 < multipole-multipole interactions.
329 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
330 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
331 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
332 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
333 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
334 > force, and torques, a Taylor expansion with $n$ terms must be
335 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
336  
344 To carry out the same procedure for a damped electrostatic kernel, we
345 replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
346 Many of the derivatives of the damped kernel are well known from
347 Smith's early work on multipoles for the Ewald
348 summation.\cite{Smith82,Smith98}
349
350 Note that increasing the value of $n$ will add additional terms to the
351 electrostatic potential, e.g., $f_2(r)$ includes orders up to
352 $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
353 functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
354 f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
355 for computing multipole energies, forces, and torques, and smooth
356 cutoffs of these quantities can be guaranteed as long as the number of
357 terms in the Taylor series exceeds the derivative order required.
358
337   For multipole-multipole interactions, following this procedure results
338 < in separate radial functions for each distinct orientational
339 < contribution to the potential, and ensures that the forces and torques
340 < from {\it each} of these contributions will vanish at the cutoff
341 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
342 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
338 > in separate radial functions for each of the distinct orientational
339 > contributions to the potential, and ensures that the forces and
340 > torques from each of these contributions will vanish at the cutoff
341 > radius.  For example, the direct dipole dot product
342 > ($\mathbf{D}_{a}
343 > \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
344   dot products:
345   \begin{equation}
346 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
347 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
348 < -\frac{1}{4\pi \epsilon_0}
349 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
350 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
346 > U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 >  \mathbf{D}_{a} \cdot
348 > \mathbf{D}_{b} \right) v_{21}(r) +
349 > \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
350 > \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
351   \end{equation}
352  
353 < The electrostatic forces and torques acting on the central multipole
354 < site due to another site within cutoff sphere are derived from
353 > For the Taylor shifted (TSF) method with the undamped kernel,
354 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
355 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
356 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
357 > connection to unmodified electrostatics as well as the smooth
358 > transition to zero in both these functions as $r\rightarrow r_c$.  The
359 > electrostatic forces and torques acting on the central multipole due
360 > to another site within the cutoff sphere are derived from
361   Eq.~\ref{generic}, accounting for the appropriate number of
362   derivatives. Complete energy, force, and torque expressions are
363   presented in the first paper in this series (Reference
364 < \citep{PaperI}).
364 > \onlinecite{PaperI}).
365  
366   \subsection{Gradient-shifted force (GSF)}
367  
368 < A second (and significantly simpler) method involves shifting the
369 < gradient of the raw coulomb potential for each particular multipole
368 > A second (and conceptually simpler) method involves shifting the
369 > gradient of the raw Coulomb potential for each particular multipole
370   order.  For example, the raw dipole-dipole potential energy may be
371   shifted smoothly by finding the gradient for two interacting dipoles
372   which have been projected onto the surface of the cutoff sphere
373   without changing their relative orientation,
374 < \begin{displaymath}
375 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
376 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
377 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
378 < \end{displaymath}
379 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
380 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
381 < the signs are reversed for the dipole that has been projected onto the
382 < cutoff sphere).  In many ways, this simpler approach is closer in
383 < spirit to the original shifted force method, in that it projects a
384 < neutralizing multipole (and the resulting forces from this multipole)
385 < onto a cutoff sphere. The resulting functional forms for the
386 < potentials, forces, and torques turn out to be quite similar in form
387 < to the Taylor-shifted approach, although the radial contributions are
388 < significantly less perturbed by the Gradient-shifted approach than
389 < they are in the Taylor-shifted method.
374 > \begin{equation}
375 > U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)  = U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r) -
376 > U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c)
377 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
378 >  \nabla U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c).
379 > \end{equation}
380 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
381 > (although the signs are reversed for the dipole that has been
382 > projected onto the cutoff sphere).  In many ways, this simpler
383 > approach is closer in spirit to the original shifted force method, in
384 > that it projects a neutralizing multipole (and the resulting forces
385 > from this multipole) onto a cutoff sphere. The resulting functional
386 > forms for the potentials, forces, and torques turn out to be quite
387 > similar in form to the Taylor-shifted approach, although the radial
388 > contributions are significantly less perturbed by the gradient-shifted
389 > approach than they are in the Taylor-shifted method.
390 >
391 > For the gradient shifted (GSF) method with the undamped kernel,
392 > $v_{21}(r) = -\frac{1}{r^3} - \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
393 > $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
394 > Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
395 > because the Taylor expansion retains only one term, they are
396 > significantly less perturbed than the TSF functions.
397  
398   In general, the gradient shifted potential between a central multipole
399   and any multipolar site inside the cutoff radius is given by,
400   \begin{equation}
401 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
402 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
403 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
401 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
402 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
403 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
404   \label{generic2}
405   \end{equation}
406   where the sum describes a separate force-shifting that is applied to
407 < each orientational contribution to the energy.
407 > each orientational contribution to the energy.  In this expression,
408 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
409 > ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
410 > represent the orientations the multipoles.
411  
412   The third term converges more rapidly than the first two terms as a
413   function of radius, hence the contribution of the third term is very
414   small for large cutoff radii.  The force and torque derived from
415 < equation \ref{generic2} are consistent with the energy expression and
416 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
415 > Eq. \ref{generic2} are consistent with the energy expression and
416 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
417   can be considered generalizations of the original DSF method for
418   higher order multipole interactions. GSF and TSF are also identical up
419   to the charge-dipole interaction but generate different expressions in
420   the energy, force and torque for higher order multipole-multipole
421   interactions. Complete energy, force, and torque expressions for the
422   GSF potential are presented in the first paper in this series
423 < (Reference \citep{PaperI})
423 > (Reference~\onlinecite{PaperI}).
424  
425  
426   \subsection{Shifted potential (SP) }
# Line 439 | Line 433 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
433   interactions with the central multipole and the image. This
434   effectively shifts the total potential to zero at the cutoff radius,
435   \begin{equation}
436 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
436 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
437 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
438   \label{eq:SP}
439   \end{equation}          
440   where the sum describes separate potential shifting that is done for
441   each orientational contribution to the energy (e.g. the direct dipole
442   product contribution is shifted {\it separately} from the
443   dipole-distance terms in dipole-dipole interactions).  Note that this
444 < is not a simple shifting of the total potential at $R_c$. Each radial
444 > is not a simple shifting of the total potential at $r_c$. Each radial
445   contribution is shifted separately.  One consequence of this is that
446   multipoles that reorient after leaving the cutoff sphere can re-enter
447   the cutoff sphere without perturbing the total energy.
448  
449 < The potential energy between a central multipole and other multipolar
450 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
451 < force and torque obtained from the shifted potential (SP) are
452 < discontinuous at $R_c$. Therefore, MD simulations will still
453 < experience energy drift while operating under the SP potential, but it
454 < may be suitable for Monte Carlo approaches where the configurational
455 < energy differences are the primary quantity of interest.
449 > For the shifted potential (SP) method with the undamped kernel,
450 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
451 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
452 > central multipole and other multipolar sites goes smoothly to zero as
453 > $r \rightarrow r_c$.  However, the force and torque obtained from the
454 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
455 > will still experience energy drift while operating under the SP
456 > potential, but it may be suitable for Monte Carlo approaches where the
457 > configurational energy differences are the primary quantity of
458 > interest.
459  
460 < \subsection{The Self term}
460 > \subsection{The Self Term}
461   In the TSF, GSF, and SP methods, a self-interaction is retained for
462   the central multipole interacting with its own image on the surface of
463   the cutoff sphere.  This self interaction is nearly identical with the
464   self-terms that arise in the Ewald sum for multipoles.  Complete
465   expressions for the self terms are presented in the first paper in
466 < this series (Reference \citep{PaperI})  
466 > this series (Reference \onlinecite{PaperI}).
467  
468  
469   \section{\label{sec:methodology}Methodology}
# Line 477 | Line 475 | arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} Thi
475   real-space cutoffs.  In the first paper of this series, we compared
476   the dipolar and quadrupolar energy expressions against analytic
477   expressions for ordered dipolar and quadrupolar
478 < arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
479 < multipolar Ewald sum as a reference method for comparing energies,
480 < forces, and torques for molecular models that mimic disordered and
481 < ordered condensed-phase systems.  These test-cases include:
478 > arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
479 > used the multipolar Ewald sum as a reference method for comparing
480 > energies, forces, and torques for molecular models that mimic
481 > disordered and ordered condensed-phase systems.  The parameters used
482 > in the test cases are given in table~\ref{tab:pars}.
483  
484 < \begin{itemize}
485 < \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
486 < \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
487 < \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
488 < \item Soft Quadrupolar solids  ($\sigma = , \epsilon = , Q_{xx} = ...$)
489 < \item A mixed multipole model for water
490 < \item A mixed multipole models for water with dissolved ions
491 < \end{itemize}
492 < This last test case exercises all levels of the multipole-multipole
493 < interactions we have derived so far and represents the most complete
494 < test of the new methods.
484 > \begin{table}
485 >  \caption{The parameters used in the systems used to evaluate the new
486 >    real-space methods.  The most comprehensive test was a liquid
487 >    composed of 2000 soft DQ liquid molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
488 >    ions).  This test exercises all orders of the multipolar
489 >    interactions developed in the first paper.\label{tab:pars}}
490 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
491 >             & \multicolumn{2}{c|}{LJ parameters} &
492 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
493 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
494 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
495 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
496 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
497 > \AA\textsuperscript{2})} \\ \hline
498 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
499 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
500 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
501 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
502 >      Soft DQ liquid  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
503 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
504 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
505 > \end{tabularx}
506 > \end{table}
507 > The systems consist of pure multipolar solids (both dipole and
508 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
509 > fluid composed of sites containing both dipoles and quadrupoles
510 > simultaneously, and a final test case that includes ions with point
511 > charges in addition to the multipolar fluid.  The solid-phase
512 > parameters were chosen so that the systems can explore some
513 > orientational freedom for the multipolar sites, while maintaining
514 > relatively strict translational order.  The soft DQ liquid model used
515 > here based loosely on the SSDQO water
516 > model,\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} but is not itself a
517 > particularly accurate water model.  However, the soft DQ model does
518 > test dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
519 > interactions at roughly the same magnitudes. The last test case, a
520 > soft DQ liquid with dissolved ions, exercises \textit{all} levels of
521 > the multipole-multipole interactions we have derived so far and
522 > represents the most complete test of the new methods.
523  
524   In the following section, we present results for the total
525   electrostatic energy, as well as the electrostatic contributions to
526   the force and torque on each molecule.  These quantities have been
527   computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
528 < and have been compared with the values obtaine from the multipolar
529 < Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
528 > and have been compared with the values obtained from the multipolar
529 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
530   between two configurations is the primary quantity that governs how
531 < the simulation proceeds. These differences are the most imporant
531 > the simulation proceeds. These differences are the most important
532   indicators of the reliability of a method even if the absolute
533   energies are not exact.  For each of the multipolar systems listed
534   above, we have compared the change in electrostatic potential energy
# Line 510 | Line 537 | contributions to the forces and torques.
537   behavior of the simulation, so we also compute the electrostatic
538   contributions to the forces and torques.
539  
540 < \subsection{Model systems}
541 < To sample independent configurations of multipolar crystals, a body
542 < centered cubic (BCC) crystal which is a minimum energy structure for
543 < point dipoles was generated using 3,456 molecules.  The multipoles
544 < were translationally locked in their respective crystal sites for
545 < equilibration at a relatively low temperature (50K), so that dipoles
546 < or quadrupoles could freely explore all accessible orientations.  The
547 < translational constraints were removed, and the crystals were
521 < simulated for 10 ps in the microcanonical (NVE) ensemble with an
522 < average temperature of 50 K.  Configurations were sampled at equal
523 < time intervals for the comparison of the configurational energy
524 < differences.  The crystals were not simulated close to the melting
525 < points in order to avoid translational deformation away of the ideal
526 < lattice geometry.
540 > \subsection{Implementation}
541 > The real-space methods developed in the first paper in this series
542 > have been implemented in our group's open source molecular simulation
543 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
544 > this work.  The complementary error function can be a relatively slow
545 > function on some processors, so all of the radial functions are
546 > precomputed on a fine grid and are spline-interpolated to provide
547 > values when required.  
548  
549 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
550 < system was created with 2048 molecules oriented randomly.  These were
549 > Using the same simulation code, we compare to a multipolar Ewald sum
550 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
551 > the Ewald sum is a re-implementation of the algorithm originally
552 > proposed by Smith that does not use the particle mesh or smoothing
553 > approximations.\cite{Smith82,Smith98} This implementation was tested
554 > extensively against the analytic energy constants for the multipolar
555 > lattices that are discussed in reference \onlinecite{PaperI}.  In all
556 > cases discussed below, the quantities being compared are the
557 > electrostatic contributions to energies, force, and torques.  All
558 > other contributions to these quantities (i.e. from Lennard-Jones
559 > interactions) are removed prior to the comparisons.
560  
561 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
562 < temperature after equilibration.  We collected 250 different
563 < configurations in equal interval of time. For the ions mixed liquid
564 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
565 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
566 < at the same environment for 1ns and 250 configurations were
567 < collected. While comparing energies, forces, and torques with Ewald
568 < method, Lennard-Jones potentials were turned off and purely
569 < electrostatic interaction had been compared.
561 > The convergence parameter ($\alpha$) also plays a role in the balance
562 > of the real-space and reciprocal-space portions of the Ewald
563 > calculation.  Typical molecular mechanics packages set this to a value
564 > that depends on the cutoff radius and a tolerance (typically less than
565 > $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
566 > associated with increasing accuracy at the expense of computational
567 > time spent on the reciprocal-space portion of the
568 > summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
569 > 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
570 > Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
571  
572 + The real-space models have self-interactions that provide
573 + contributions to the energies only.  Although the self interaction is
574 + a rapid calculation, we note that in systems with fluctuating charges
575 + or point polarizabilities, the self-term is not static and must be
576 + recomputed at each time step.
577 +
578 + \subsection{Model systems}
579 + To sample independent configurations of the multipolar crystals, body
580 + centered cubic (bcc) crystals, which exhibit the minimum energy
581 + structures for point dipoles, were generated using 3,456 molecules.
582 + The multipoles were translationally locked in their respective crystal
583 + sites for equilibration at a relatively low temperature (50K) so that
584 + dipoles or quadrupoles could freely explore all accessible
585 + orientations.  The translational constraints were then removed, the
586 + systems were re-equilibrated, and the crystals were simulated for an
587 + additional 10 ps in the microcanonical (NVE) ensemble with an average
588 + temperature of 50 K.  The balance between moments of inertia and
589 + particle mass were chosen to allow orientational sampling without
590 + significant translational motion.  Configurations were sampled at
591 + equal time intervals in order to compare configurational energy
592 + differences.  The crystals were simulated far from the melting point
593 + in order to avoid translational deformation away of the ideal lattice
594 + geometry.
595 +
596 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
597 + simulations, each system was created with 2,048 randomly-oriented
598 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
599 + Each system was then simulated for 1 ns in the microcanonical (NVE)
600 + ensemble with the Dullweber, Leimkuhler, and McLachlan (DLM)
601 + symplectic splitting integrator using 1 fs
602 + timesteps.\cite{Dullweber1997} We collected 250 different
603 + configurations at equal time intervals. For the liquid system that
604 + included ionic species, we converted 48 randomly-distributed molecules
605 + into 24 \ce{Na+} and 24 \ce{Cl-} ions and re-equilibrated. After
606 + equilibration, the system was run under the same conditions for 1
607 + ns. A total of 250 configurations were collected. In the following
608 + comparisons of energies, forces, and torques, the Lennard-Jones
609 + potentials were turned off and only the purely electrostatic
610 + quantities were compared with the same values obtained via the Ewald
611 + sum.
612 +
613   \subsection{Accuracy of Energy Differences, Forces and Torques}
614   The pairwise summation techniques (outlined above) were evaluated for
615   use in MC simulations by studying the energy differences between
# Line 550 | Line 622 | we used least square regressions analysiss for the six
622   should be identical for all methods.
623  
624   Since none of the real-space methods provide exact energy differences,
625 < we used least square regressions analysiss for the six different
625 > we used least square regressions analysis for the six different
626   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
627 < with the multipolar Ewald reference method.  Unitary results for both
628 < the correlation (slope) and correlation coefficient for these
629 < regressions indicate perfect agreement between the real-space method
630 < and the multipolar Ewald sum.
627 > with the multipolar Ewald reference method.  A result of unity for
628 > both the correlation (slope) and coefficient of determination ($R^2$)
629 > for these regressions would indicate perfect agreement between the
630 > real-space method and the multipolar Ewald sum.
631  
632   Molecular systems were run long enough to explore independent
633   configurations and 250 configurations were recorded for comparison.
634   Each system provided 31,125 energy differences for a total of 186,750
635   data points.  Similarly, the magnitudes of the forces and torques have
636 < also been compared by using least squares regression analyses. In the
636 > also been compared using least squares regression analysis. In the
637   forces and torques comparison, the magnitudes of the forces acting in
638   each molecule for each configuration were evaluated. For example, our
639   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 574 | Line 646 | force and torque vectors. Fisher developed a probablit
646   simulations.  Because the real space methods reweight the different
647   orientational contributions to the energies, it is also important to
648   understand how the methods impact the \textit{directionality} of the
649 < force and torque vectors. Fisher developed a probablity density
649 > force and torque vectors. Fisher developed a probability density
650   function to analyse directional data sets,
651   \begin{equation}
652   p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
# Line 588 | Line 660 | evaluated,
660   the forces obtained via the Ewald sum and the real-space methods were
661   evaluated,
662   \begin{equation}
663 < \cos\theta_i =  \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
664 <  \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
663 >  \cos\theta_i =  \frac{\mathbf{f}_i^\mathrm{~Ewald} \cdot
664 >    \mathbf{f}_i^\mathrm{~GSF}}{\left|\mathbf{f}_i^\mathrm{~Ewald}\right| \left|\mathbf{f}_i^\mathrm{~GSF}\right|}
665   \end{equation}
666   The total angular displacement of the vectors was calculated as,
667   \begin{equation}
# Line 612 | Line 684 | system of 2000 SSDQ water molecules with 24 \ce{Na+} a
684  
685   \subsection{Energy conservation}
686   To test conservation the energy for the methods, the mixed molecular
687 < system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
688 < ions was run for 1 ns in the microcanonical ensemble at an average
689 < temperature of 300K.  Each of the different electrostatic methods
690 < (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
691 < damping values. The molecular system was started with same initial
692 < positions and velocities for all cutoff methods. The energy drift
693 < ($\delta E_1$) and standard deviation of the energy about the slope
694 < ($\delta E_0$) were evaluated from the total energy of the system as a
695 < function of time.  Although both measures are valuable at
687 > system of 2000 soft DQ liquid molecules with 24 \ce{Na+} and 24
688 > \ce{Cl-} ions was run for 1 ns in the microcanonical ensemble at an
689 > average temperature of 300K.  Each of the different electrostatic
690 > methods (Ewald, Hard, SP, GSF, and TSF) was tested for a range of
691 > different damping values. The molecular system was started with same
692 > initial positions and velocities for all cutoff methods. The energy
693 > drift ($\delta E_1$) and standard deviation of the energy about the
694 > slope ($\delta E_0$) were evaluated from the total energy of the
695 > system as a function of time.  Although both measures are valuable at
696   investigating new methods for molecular dynamics, a useful interaction
697   model must allow for long simulation times with minimal energy drift.
698  
699   \section{\label{sec:result}RESULTS}
700   \subsection{Configurational energy differences}
701 < %The magnitude of the fluctuation in the total electrostatic energy per molecule for a dipolar crystal is very high as shown in (Fig … paper I).\cite{PaperI}As soon as, the net dipole moment within a cutoff radius is neutralized in the SP method, the magnitude of the fluctuation in the total electrostatic energy per molecule reduced significantly and rapidly converged to the correct energy constant (Refer figure … Paper I).\cite{PaperI} The GSF potential energy also converged to the correct energy constant for the cutoff radius rc = 6a for the undamped case. The potential energy from the TSF method converges towards the correct value for a very large cutoff radius. The speed of convergence for the all the cutoff methods can be increased by using damping function as shown in figure … Paper I\cite{PaperI}. For the quadrupolar crystals, the fluctuation in the total electrostatic energy for the hard cutoff method is small and short ranged as compared to the dipolar crystals (figure … in the paper I).\cite{PaperI}  Similar to the dipolar crystals, the net quadrupolar neutralization of the cutoff sphere in the SP method reduces oscillation rapidly and converge electrostatic energy to the correct energy constant.
630 < %The oscillation in the the electrostatic energy for the hard cutoff method is even true for the dipolar liquids as shown in Figure ~\ref{fig:rcutConvergence_dipolarLiquid}. As we placed image on the surface of the cutoff sphere in SP method, the oscillation in the energy is reduced. The fluctuation in the energy in liquid is much smaller as compared to the crystal (This result is similar to the results observed by \textit{Wolf et al.} in the case of ionic crystal and Mgo melt). The large magnitude in the fluctuation of the electrostatic energy in the crystal is because of large range of multipole ordering in the crystal. When the energy is evaluated by the direct truncation, it breaks up large number of multipolar ordering leaving behind net multipole moment. But in the case of liquid, there is only local ordering of the multipoles and their ordering disappears in the long range. Therefore, the direct truncation results a small oscillation in the electrostatic energy (which is smaller than deviation SP energy from the Ewald Figure ~\ref{fig:rcutConvergence_dipolarLiquid}) in the case of liquid. Although, the oscillation in the energy is very small for the case of liquid, this affects the change in potential energy ($\triangle E$), which is observed when $\triangle E$ evaluated from the SP method compared with Ewald as shown in figure 4a and 4b.
631 < %\begin{figure}[h!]
632 < %        \centering
633 < %        \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
634 < %        \caption{The energy per molecule plotted against cutoff radius, rc for i) Hard ii) SP iii) GSF, and iv) TSF method. The hard cutoff method shows fluctuation in the electorstatic energy and it disappers in all other methods.  }
635 < %        \label{fig:rcutConvergence_dipolarLiquid}
636 < %    \end{figure}
637 < %In MC simulations, the electrostatic differences between the molecules are important parameter for sampling. We have compared $\triangle E$ from the different methods (Hard, SP, GSF, and TSF) with the Ewald using linear regression analysis. The correlation coefficient ($R^2$) of the regression line measures the deviation of the evaluated quantities from the mean slope. We know that Ewald method evaluates accurate value of the electrostatic energy. Hence, if the proposed methods can quantify the electrostatic energy as good as Ewald then the correlation coefficient is 1.The correlation coefficient is 0 for the completely random result for any physical quantity measured by the proposed method. The slope is a measure of the accuracy of the average of a physical quantity obtained from the proposed methods. If the slope is 1 then we can conclude that the average of the physical quantity measured by the method is as good as Ewald. The deviation of the slope from 1 state that the method used in quantifying physical quantities is statistically biased as compared to Ewald.
638 < %\begin{figure}
639 < %        \centering
640 < %        \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
641 < %        \label{fig:barGraph1}
642 < %        \end{figure}
643 < %        \begin{figure}
644 < %        \centering
645 < %       \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
646 < %        \caption{}
647 <      
648 < %        \label{fig:barGraph2}
649 < %      \end{figure}
650 < %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.
701 >
702   \begin{figure}
703 <        \centering
704 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
705 <        \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 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle)}
706 <        \label{fig:slopeCorr_energy}
707 <    \end{figure}
708 < 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.
703 >  \centering
704 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined.eps}
705 >  \caption{Statistical analysis of the quality of configurational
706 >    energy differences for the real-space electrostatic methods
707 >    compared with the reference Ewald sum.  Results with a value equal
708 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
709 >    from those obtained using the multipolar Ewald sum.  Different
710 >    values of the cutoff radius are indicated with different symbols
711 >    (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = inverted
712 >    triangles).\label{fig:slopeCorr_energy}}
713 > \end{figure}
714 >
715 > The combined coefficient of determination and slope for all six
716 > systems is shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the
717 > methods reproduce the Ewald configurational energy differences with
718 > remarkable fidelity.  Undamped hard cutoffs introduce a significant
719 > amount of random scatter in the energy differences which is apparent
720 > in the reduced value of $R^2$ for this method.  This can be easily
721 > understood as configurations which exhibit small traversals of a few
722 > dipoles or quadrupoles out of the cutoff sphere will see large energy
723 > jumps when hard cutoffs are used.  The orientations of the multipoles
724 > (particularly in the ordered crystals) mean that these energy jumps
725 > can go in either direction, producing a significant amount of random
726 > scatter, but no systematic error.
727 >
728 > The TSF method produces energy differences that are highly correlated
729 > with the Ewald results, but it also introduces a significant
730 > systematic bias in the values of the energies, particularly for
731 > smaller cutoff values. The TSF method alters the distance dependence
732 > of different orientational contributions to the energy in a
733 > non-uniform way, so the size of the cutoff sphere can have a large
734 > effect, particularly for the crystalline systems.
735 >
736 > Both the SP and GSF methods appear to reproduce the Ewald results with
737 > excellent fidelity, particularly for moderate damping ($\alpha \approx
738 > 0.2$~\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
739 > 12$~\AA).  With the exception of the undamped hard cutoff, and the TSF
740 > method with short cutoffs, all of the methods would be appropriate for
741 > use in Monte Carlo simulations.
742 >
743   \subsection{Magnitude of the force and torque vectors}
744 < 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.
744 >
745 > The comparisons of the magnitudes of the forces and torques for the
746 > data accumulated from all six systems are shown in Figures
747 > ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
748 > correlation and slope for the forces agree well with the Ewald sum
749 > even for the hard cutoffs.
750 >
751 > For systems of molecules with only multipolar interactions, the pair
752 > energy contributions are quite short ranged.  Moreover, the force
753 > decays more rapidly than the electrostatic energy, hence the hard
754 > cutoff method can also produce reasonable agreement for this quantity.
755 > Although the pure cutoff gives reasonably good electrostatic forces
756 > for pairs of molecules included within each other's cutoff spheres,
757 > the discontinuity in the force at the cutoff radius can potentially
758 > cause energy conservation problems as molecules enter and leave the
759 > cutoff spheres.  This is discussed in detail in section
760 > \ref{sec:conservation}.
761 >
762 > The two shifted-force methods (GSF and TSF) exhibit a small amount of
763 > systematic variation and scatter compared with the Ewald forces.  The
764 > shifted-force models intentionally perturb the forces between pairs of
765 > molecules inside each other's cutoff spheres in order to correct the
766 > energy conservation issues, and this perturbation is evident in the
767 > statistics accumulated for the molecular forces.  The GSF
768 > perturbations are minimal, particularly for moderate damping and
769 > commonly-used cutoff values ($r_c = 12$~\AA).  The TSF method shows
770 > reasonable agreement in $R^2$, but again the systematic error in the
771 > forces is concerning if replication of Ewald forces is desired.
772 >
773 > It is important to note that the forces and torques from the SP and
774 > the Hard cutoffs are not identical. The SP method shifts each
775 > orientational contribution separately (e.g. the dipole-dipole dot
776 > product is shifted by a different function than the dipole-distance
777 > products), while the hard cutoff contains no orientation-dependent
778 > shifting.  The forces and torques for these methods therefore diverge
779 > for multipoles even though the forces for point charges are identical.
780 >
781   \begin{figure}
782 <        \centering
783 <        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
784 <        \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 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle). }
785 <        \label{fig:slopeCorr_force}
786 <    \end{figure}
787 < 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.
782 >  \centering
783 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
784 >  \caption{Statistical analysis of the quality of the force vector
785 >    magnitudes for the real-space electrostatic methods compared with
786 >    the reference Ewald sum. Results with a value equal to 1 (dashed
787 >    line) indicate force magnitude values indistinguishable from those
788 >    obtained using the multipolar Ewald sum.  Different values of the
789 >    cutoff radius are indicated with different symbols (9~\AA\ =
790 >    circles, 12~\AA\ = squares, and 15~\AA\ = inverted
791 >    triangles).\label{fig:slopeCorr_force}}
792 > \end{figure}
793 >
794 >
795   \begin{figure}
796 <        \centering
797 <        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
798 <        \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).}
799 <        \label{fig:slopeCorr_torque}
800 <    \end{figure}
796 >  \centering
797 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
798 >  \caption{Statistical analysis of the quality of the torque vector
799 >    magnitudes for the real-space electrostatic methods compared with
800 >    the reference Ewald sum. Results with a value equal to 1 (dashed
801 >    line) indicate force magnitude values indistinguishable from those
802 >    obtained using the multipolar Ewald sum.  Different values of the
803 >    cutoff radius are indicated with different symbols (9~\AA\ =
804 >    circles, 12~\AA\ = squares, and 15~\AA\ = inverted
805 >    triangles).\label{fig:slopeCorr_torque}}
806 > \end{figure}
807 >
808 > The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
809 > significantly influenced by the choice of real-space method.  The
810 > torque expressions have the same distance dependence as the energies,
811 > which are naturally longer-ranged expressions than the inter-site
812 > forces.  Torques are also quite sensitive to orientations of
813 > neighboring molecules, even those that are near the cutoff distance.
814 >
815 > The results shows that the torque from the hard cutoff method
816 > reproduces the torques in quite good agreement with the Ewald sum.
817 > The other real-space methods can cause some deviations, but excellent
818 > agreement with the Ewald sum torques is recovered at moderate values
819 > of the damping coefficient ($\alpha \approx 0.2$~\AA$^{-1}$) and cutoff
820 > radius ($r_c \ge 12$~\AA).  The TSF method exhibits only fair agreement
821 > in the slope when compared with the Ewald torques even for larger
822 > cutoff radii.  It appears that the severity of the perturbations in
823 > the TSF method are most in evidence for the torques.
824 >
825   \subsection{Directionality of the force and torque vectors}  
674 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 $r_c$.
826  
827 + The accurate evaluation of force and torque directions is just as
828 + important for molecular dynamics simulations as the magnitudes of
829 + these quantities. Force and torque vectors for all six systems were
830 + analyzed using Fisher statistics, and the quality of the vector
831 + directionality is shown in terms of circular variance
832 + ($\mathrm{Var}(\theta)$) in
833 + Fig. \ref{fig:slopeCorr_circularVariance}. The force and torque
834 + vectors from the new real-space methods exhibit nearly-ideal Fisher
835 + probability distributions (Eq.~\ref{eq:pdf}). Both the hard and SP
836 + cutoff methods exhibit the best vectorial agreement with the Ewald
837 + sum. The force and torque vectors from GSF method also show good
838 + agreement with the Ewald method, which can also be systematically
839 + improved by using moderate damping and a reasonable cutoff radius. For
840 + $\alpha = 0.2$~\AA$^{-1}$ and $r_c = 12$~\AA, we observe
841 + $\mathrm{Var}(\theta) = 0.00206$, which corresponds to a distribution
842 + with 95\% of force vectors within $6.37^\circ$ of the corresponding
843 + Ewald forces. The TSF method produces the poorest agreement with the
844 + Ewald force directions.
845 +
846 + Torques are again more perturbed than the forces by the new real-space
847 + methods, but even here the variance is reasonably small.  For the same
848 + method (GSF) with the same parameters ($\alpha = 0.2$~\AA$^{-1}$, $r_c
849 + = 12$~\AA), the circular variance was 0.01415, corresponds to a
850 + distribution which has 95\% of torque vectors are within $16.75^\circ$
851 + of the Ewald results. Again, the direction of the force and torque
852 + vectors can be systematically improved by varying $\alpha$ and $r_c$.
853 +
854   \begin{figure}
855 <        \centering
856 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
857 <        \caption{The circular variance of the data sets of the
858 <          direction of the  force and torque vectors obtained from a
859 <          given method about reference Ewald method. The result equal
860 <          to 0 (dashed line) indicates direction of the vectors are
861 <          indistinguishable from the Ewald method. Here different
862 <          symbols represent different value of the cutoff radius (9
863 <          \AA\ = circle, 12 \AA\ = square, 15 \AA\  = inverted triangle)}
864 <        \label{fig:slopeCorr_circularVariance}
687 <    \end{figure}
688 < \subsection{Total energy conservation}
689 < We have tested the conservation of energy in the SSDQC liquid system
690 < by running system for 1ns in the Hard, SP, GSF and TSF method. The
691 < Hard cutoff method shows very high energy drifts 433.53
692 < KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol
693 < (measured by the SD from the slope) for the undamped case, which makes
694 < it completely unusable in MD simulations. The SP method also shows
695 < large value of energy drift 1.289 Kcal/Mol/ns/particle and energy
696 < fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method
697 < is due to the non-vanishing nature of the torque and force at the
698 < cutoff radius. We can improve the energy conservation in some extent
699 < by the proper selection of the damping alpha but the improvement is
700 < not good enough, which can be observed in Figure 9a and 9b .The GSF
701 < and TSF shows very low value of energy drift 0.09016, 0.07371
702 < KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol
703 < respectively for the undamped case. Since the absolute value of the
704 < evaluated electrostatic energy, force and torque from TSF method are
705 < deviated from the Ewald, it does not mimic MD simulations
706 < appropriately. The electrostatic energy, force and torque from the GSF
707 < method have very good agreement with the Ewald. In addition, the
708 < energy drift and energy fluctuation from the GSF method is much better
709 < than Ewald’s method for reciprocal space vector value ($k_f$) equal to
710 < 7 as shown in Figure~\ref{fig:energyDrift} and
711 < ~\ref{fig:fluctuation}. We can improve the total energy fluctuation
712 < and drift for the Ewald’s method by increasing size of the reciprocal
713 < space, which extremely increseses the simulation time. In our current
714 < simulation, the simulation time for the Hard, SP, and GSF methods are
715 < about 5.5 times faster than the Ewald method.
855 >  \centering
856 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
857 >  \caption{The circular variance of the direction of the force and
858 >    torque vectors obtained from the real-space methods around the
859 >    reference Ewald vectors. A variance equal to 0 (dashed line)
860 >    indicates direction of the force or torque vectors are
861 >    indistinguishable from those obtained from the Ewald sum. Here
862 >    different symbols represent different values of the cutoff radius
863 >    (9~\AA\ = circle, 12~\AA\ = square, 15~\AA\ = inverted triangle)\label{fig:slopeCorr_circularVariance}}
864 > \end{figure}
865  
866 < In Fig.~\ref{fig:energyDrift}, $\delta \mbox{E}_1$ is a measure of the
718 < linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
719 < over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
720 < standard deviation of the energy fluctuations in units of $\mbox{kcal
721 <  mol}^{-1}$ per particle. In the bottom plot, it is apparent that the
722 < energy drift is reduced by a significant amount (2 to 3 orders of
723 < magnitude improvement at all values of the damping coefficient) by
724 < chosing either of the shifted-force methods over the hard or SP
725 < methods.  We note that the two shifted-force method can give
726 < significantly better energy conservation than the multipolar Ewald sum
727 < with the same choice of real-space cutoffs.
866 > \subsection{Energy conservation\label{sec:conservation}}
867  
868 + We have tested the conservation of energy one can expect to see with
869 + the new real-space methods using the soft DQ liquid model with a small
870 + fraction of solvated ions. This is a test system which exercises all
871 + orders of multipole-multipole interactions derived in the first paper
872 + in this series and provides the most comprehensive test of the new
873 + methods.  A liquid-phase system was created with 2000 liquid-phase
874 + molecules and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
875 + temperature of 300K.  After equilibration in the canonical (NVT)
876 + ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
877 + was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
878 + Hard, SP, GSF, and TSF methods with a cutoff radius of 12~\AA.  The
879 + value of the damping coefficient was also varied from the undamped
880 + case ($\alpha = 0$) to a heavily damped case ($\alpha =
881 + 0.3$~\AA$^{-1}$) for all of the real space methods.  A sample was also
882 + run using the multipolar Ewald sum with the same real-space cutoff.
883 +
884 + In figure~\ref{fig:energyDrift} we show the both the linear drift in
885 + energy over time, $\delta E_1$, and the standard deviation of energy
886 + fluctuations around this drift $\delta E_0$.  Both of the
887 + shifted-force methods (GSF and TSF) provide excellent energy
888 + conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
889 + while the hard cutoff is essentially unusable for molecular dynamics.
890 + SP provides some benefit over the hard cutoff because the energetic
891 + jumps that happen as particles leave and enter the cutoff sphere are
892 + somewhat reduced, but like the Wolf method for charges, the SP method
893 + would not be as useful for molecular dynamics as either of the
894 + shifted-force methods.
895 +
896 + We note that for all tested values of the cutoff radius, the new
897 + real-space methods can provide better energy conservation behavior
898 + than the multipolar Ewald sum, even when relatively large $k$-space
899 + cutoff values are utilized.
900 +
901   \begin{figure}
902    \centering
903 <  \includegraphics[width=\textwidth]{newDrift.pdf}
904 < \label{fig:energyDrift}        
905 < \caption{Analysis of the energy conservation of the real space
906 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
907 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
908 <  of energy fluctuations around this drift.  All simulations were of a
909 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
910 <  K starting from the same initial configuration.}
903 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
904 >  \caption{Energy conservation of the real-space methods for the soft
905 >    DQ liauid / ion system. $\delta \mathrm{E}_1$ is the linear drift
906 >    in energy over time (in kcal/mol/particle/ns) and $\delta
907 >    \mathrm{E}_0$ is the standard deviation of energy fluctuations
908 >    around this drift (in kcal/mol/particle).  Points that appear in
909 >    the green region at the bottom exhibit better energy conservation
910 >    than would be obtained using common parameters for Ewald-based
911 >    electrostatics.\label{fig:energyDrift}}
912   \end{figure}
913  
914 + \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
915 + The most important test of the modified interaction potentials is the
916 + fidelity with which they can reproduce structural features and
917 + dynamical properties in a liquid.  One commonly-utilized measure of
918 + structural ordering is the pair distribution function, $g(r)$, which
919 + measures local density deviations in relation to the bulk density.  In
920 + the electrostatic approaches studied here, the short-range repulsion
921 + from the Lennard-Jones potential is identical for the various
922 + electrostatic methods, and since short range repulsion determines much
923 + of the local liquid ordering, one would not expect to see many
924 + differences in $g(r)$.  Indeed, the pair distributions are essentially
925 + identical for all of the electrostatic methods studied (for each of
926 + the different systems under investigation).
927 +
928 + % An example of this agreement for the soft DQ liquid/ion system is
929 + % shown in Fig. \ref{fig:gofr}.
930 +
931 + % \begin{figure}
932 + %   \centering
933 + %   \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
934 + % \caption{The pair distribution functions, $g(r)$, for the SSDQ
935 + %   water/ion system obtained using the different real-space methods are
936 + %   essentially identical with the result from the Ewald
937 + %   treatment.\label{fig:gofr}}
938 + % \end{figure}
939 +
940 + There is a minor over-structuring of the first solvation shell when
941 + using TSF or when overdamping with any of the real-space methods.
942 + With moderate damping, GSF and SP produce pair distributions that are
943 + identical (within numerical noise) to their Ewald counterparts.  The
944 + degree of over-structuring can be measured most easily using the
945 + coordination number,
946 + \begin{equation}
947 + n_C = 4\pi\rho \int_{0}^{a}r^2\text{g}(r)dr,
948 + \end{equation}
949 + where $\rho$ is the number density of the site-site pair interactions,
950 + and $a$ is the radial location of the minima following the first peak
951 + in $g(r)$ ($a = 4.2$~\AA\  for the soft DQ liquid / ion system).  The
952 + coordination number is shown as a function of the damping coefficient
953 + for all of the real space methods in Fig. \ref{fig:Props}.
954 +
955 + A more demanding test of modified electrostatics is the average value
956 + of the electrostatic energy $\langle U_\mathrm{elect} \rangle / N$
957 + which is obtained by sampling the liquid-state configurations
958 + experienced by a liquid evolving entirely under the influence of each
959 + of the methods.  In Fig. \ref{fig:Props} we demonstrate how $\langle
960 + U_\mathrm{elect} \rangle / N$ varies with the damping parameter,
961 + $\alpha$, for each of the methods.
962 +
963 + As in the crystals studied in the first paper, damping is important
964 + for converging the mean electrostatic energy values, particularly for
965 + the two shifted force methods (GSF and TSF).  A value of $\alpha
966 + \approx 0.2$~\AA$^{-1}$ is sufficient to converge the SP and GSF
967 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
968 + dramatic damping ($\alpha \approx 0.28$~\AA$^{-1}$ for $r_c = 9$~\AA).
969 + Overdamping the real-space electrostatic methods occurs with $\alpha >
970 + 0.3$~\AA$^{-1}$, causing the estimate of the electrostatic energy to
971 + drop below the Ewald results.
972 +
973 + These ``optimal'' values of the damping coefficient for structural
974 + features are similar to those observed for DSF electrostatics for
975 + purely point-charge systems, and the range $\alpha= 0.175 \rightarrow
976 + 0.225$~\AA$^{-1}$ for $r_c = 12$~\AA\ appears to be an excellent
977 + compromise for mixed charge/multipolar systems.
978 +
979 + To test the fidelity of the electrostatic methods at reproducing
980 + \textit{dynamics} in a multipolar liquid, it is also useful to look at
981 + transport properties, particularly the diffusion constant,
982 + \begin{equation}
983 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
984 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
985 + \label{eq:diff}
986 + \end{equation}
987 + which measures long-time behavior and is sensitive to the forces on
988 + the multipoles. The self-diffusion constants (D) were calculated from
989 + linear fits to the long-time portion of the mean square displacement,
990 + $\langle r^{2}(t) \rangle$.\cite{Allen87} In Fig. \ref{fig:Props} we
991 + demonstrate how the diffusion constant depends on the choice of
992 + real-space methods and the damping coefficient.  Both the SP and GSF
993 + methods can obtain excellent agreement with Ewald again using moderate
994 + damping.
995 +
996 + In addition to translational diffusion, orientational relaxation times
997 + were calculated for comparisons with the Ewald simulations and with
998 + experiments. These values were determined by calculating the
999 + orientational time correlation function,
1000 + \begin{equation}
1001 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1002 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1003 + \label{eq:OrientCorr}
1004 + \end{equation}
1005 + from the same 350 ps microcanonical trajectories that were used for
1006 + translational diffusion.  Here, $P_l$ is the Legendre polynomial of
1007 + order $l$ and $\hat{\mathbf{A}}_\gamma$ is the unit vector for body
1008 + axis $\gamma$.  The reference frame used for our sample dipolar
1009 + systems has the $z$-axis running along the dipoles, and for the soft
1010 + DQ liquid model, the $y$-axis connects the two implied hydrogen-like
1011 + positions.  From the orientation autocorrelation functions, we can
1012 + obtain time constants for rotational relaxation either by fitting to a
1013 + multi-exponential model for the orientational relaxation, or by
1014 + integrating the correlation functions.
1015 +
1016 + In a good model for water, the orientational decay times would be
1017 + comparable to water orientational relaxation times from nuclear
1018 + magnetic resonance (NMR). The relaxation constant obtained from
1019 + $C_2^y(t)$ is normally of experimental interest because it describes
1020 + the relaxation of the principle axis connecting the hydrogen
1021 + atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1022 + of the dipole-dipole relaxation from a proton NMR signal and can
1023 + provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1024 + In Fig. \ref{fig:Props} we compare the $\tau_2^y$ and $\tau_2^z$
1025 + values for the various real-space methods over a range of different
1026 + damping coefficients.  The rotational relaxation for the $z$ axis
1027 + primarily probes the torques on the dipoles, while the relaxation for
1028 + the $y$ axis is sensitive primarily to the quadrupolar torques.
1029 +
1030 + \begin{figure}
1031 +  \includegraphics[width=\textwidth]{properties.eps}
1032 +  \caption{Comparison of the structural and dynamic properties for the
1033 +    combined multipolar liquid (soft DQ liquid + ions) for all of the
1034 +    real-space methods with $r_c = 12$~\AA. Electrostatic energies,
1035 +    $\langle U_\mathrm{elect} \rangle / N$ (in kcal / mol),
1036 +    coordination numbers, $n_C$, diffusion constants (in $10^{-5}
1037 +    \mathrm{cm}^2\mathrm{s}^{-1}$), and rotational correlation times
1038 +    (in ps) all show excellent agreement with Ewald results for
1039 +    damping coefficients in the range $\alpha= 0.175 \rightarrow
1040 +    0.225$~\AA$^{-1}$. \label{fig:Props}}
1041 + \end{figure}
1042 +
1043 + In Fig. \ref{fig:Props} it appears that values for $D$, $\tau_2^y$,
1044 + and $\tau_2^z$ using the Ewald sum are reproduced with excellent
1045 + fidelity by the GSF and SP methods.  All of the real space methods can
1046 + be \textit{overdamped}, which reduces the effective range of multipole
1047 + interactions, causing structural and dynamical changes from the
1048 + correct behavior.  Because overdamping weakens orientational
1049 + preferences between adjacent molecules, it manifests as too-rapid
1050 + orientational decay coupled with faster diffusion and
1051 + over-coordination of the liquid.  Underdamping is less problematic for
1052 + the SP and GSF methods, as their structural and dynamical properties
1053 + still reproduce the Ewald results even in the completely undamped
1054 + ($\alpha = 0$) case.  An optimal range for the electrostatic damping
1055 + parameter appears to be $\alpha= 0.175 \rightarrow 0.225$~\AA$^{-1}$
1056 + for $r_c = 12$~\AA, which similar to the optimal range found for the
1057 + damped shifted force potential for point charges.\cite{Fennell:2006lq}
1058 +
1059   \section{CONCLUSION}
1060 < We have generalized the charged neutralized potential energy originally developed by the Wolf et al.\cite{Wolf:1999dn} 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{Fennell:2006lq} 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.
1061 < 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.
1062 < In MD simulations, the energy conservation is very important. The
1063 < conservation of the total energy can be ensured by  i) enforcing the
1064 < smooth truncation of the energy, force and torque in the cutoff radius
1065 < and ii) making the energy, force and torque consistent with each
1066 < other. The GSF and TSF methods ensure the consistency and smooth
1067 < truncation of the energy, force and torque at the cutoff radius, as a
1068 < result show very good total energy conservation. But the TSF method
751 < does not show good agreement in the absolute value of the
752 < electrostatic energy, force and torque with the Ewald.  The GSF method
753 < has mimicked Ewald’s force, energy and torque accurately and also
754 < conserved energy. Therefore, the GSF method is the suitable method for
755 < evaluating required force field in MD simulations. In addition, the
756 < energy drift and fluctuation from the GSF method is much better than
757 < Ewald’s method for finite-sized reciprocal space.
1060 > In the first paper in this series, we generalized the
1061 > charge-neutralized electrostatic energy originally developed by Wolf
1062 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1063 > up to quadrupolar order.  The SP method is essentially a
1064 > multipole-capable version of the Wolf model.  The SP method for
1065 > multipoles provides excellent agreement with Ewald-derived energies,
1066 > forces and torques, and is suitable for Monte Carlo simulations,
1067 > although the forces and torques retain discontinuities at the cutoff
1068 > distance that prevents its use in molecular dynamics.
1069  
1070 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1071 < that can be made extremely efficient using spline interpolations of
1072 < the radial functions.  They require no Fourier transforms or $k$-space
1073 < sums, and guarantee the smooth handling of energies, forces, and
1074 < torques as multipoles cross the real-space cutoff boundary.  
1070 > We also developed two natural extensions of the damped shifted-force
1071 > (DSF) model originally proposed by Zahn {\it et al.} and extended by
1072 > Fennell and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
1073 > approaches provide smooth truncation of energies, forces, and torques
1074 > at the real-space cutoff, and both converge to DSF electrostatics for
1075 > point-charge interactions.  The TSF model is based on a high-order
1076 > truncated Taylor expansion which can be relatively perturbative inside
1077 > the cutoff sphere.  The GSF model takes the gradient from an images of
1078 > the interacting multipole that has been projected onto the cutoff
1079 > sphere to derive shifted force and torque expressions, and is a
1080 > significantly more gentle approach.
1081  
1082 + The GSF method produces quantitative agreement with Ewald energies,
1083 + forces, and torques.  It also performs well in conserving energy in MD
1084 + simulations.  The Taylor-shifted (TSF) model provides smooth dynamics,
1085 + but these take place on a potential energy surface that is
1086 + significantly perturbed from Ewald-based electrostatics.  Because it
1087 + performs relatively poorly compared with GSF, it may seem odd that
1088 + that the TSF model was included in this work.  However, the functional
1089 + forms derived for the SP and GSF methods depend on the separation of
1090 + orientational contributions that were made visible by the Taylor
1091 + series of the electrostatic kernel at the cutoff radius. The TSF
1092 + method also has the unique property that a large number of derivatives
1093 + can be made to vanish at the cutoff radius.  This property has proven
1094 + useful in past treatments of the corrections to the Clausius-Mossotti
1095 + fluctuation formula for dielectric constants.\cite{Izvekov:2008wo}
1096 +
1097 + Reproduction of both structural and dynamical features in the liquid
1098 + systems is remarkably good for both the SP and GSF models.  Pair
1099 + distribution functions are essentially equivalent to the same
1100 + functions produced using Ewald-based electrostatics, and with moderate
1101 + damping, a structural feature that directly probes the electrostatic
1102 + interaction (e.g. the mean electrostatic potential energy) can also be
1103 + made quantitative.  Dynamical features are sensitive probes of the
1104 + forces and torques produced by these methods, and even though the
1105 + smooth behavior of forces is produced by perturbing the overall
1106 + potential, the diffusion constants and orientational correlation times
1107 + are quite close to the Ewald-based results.
1108 +
1109 + The only cases we have found where the new GSF and SP real-space
1110 + methods can be problematic are those which retain a bulk dipole moment
1111 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1112 + materials, uniform weighting of the orientational contributions can be
1113 + important for converging the total energy.  In these cases, the
1114 + damping function which causes the non-uniform weighting can be
1115 + replaced by the bare electrostatic kernel, and the energies return to
1116 + the expected converged values.
1117 +
1118 + Based on the results of this work, we can conclude that the GSF method
1119 + is a suitable and efficient replacement for the Ewald sum for
1120 + evaluating electrostatic interactions in modern MD simulations, and
1121 + the SP method would be an excellent choice for Monte Carlo
1122 + simulations where smooth forces and energy conservation are not
1123 + important.  Both the SP and GSF methods retain excellent fidelity to
1124 + the Ewald energies, forces and torques.  Additionally, the energy
1125 + drift and fluctuations from the GSF electrostatics are significantly
1126 + better than a multipolar Ewald sum for finite-sized reciprocal spaces,
1127 + and physical properties are reproduced accurately.
1128 +
1129 + As in all purely pairwise cutoff methods, the SP, GSF and TSF methods
1130 + are expected to scale approximately {\it linearly} with system size,
1131 + and are easily parallelizable.  This should result in substantial
1132 + reductions in the computational cost of performing large simulations.
1133 + With the proper use of pre-computation and spline interpolation of the
1134 + radial functions, the real-space methods are essentially the same cost
1135 + as a simple real-space cutoff.  They require no Fourier transforms or
1136 + $k$-space sums, and guarantee the smooth handling of energies, forces,
1137 + and torques as multipoles cross the real-space cutoff boundary.
1138 +
1139 + We are not suggesting that there is any flaw with the Ewald sum; in
1140 + fact, it is the standard by which the SP, GSF, and TSF methods have
1141 + been judged in this work.  However, these results provide evidence
1142 + that in the typical simulations performed today, the Ewald summation
1143 + may no longer be required to obtain the level of accuracy most
1144 + researchers have come to expect.
1145 +
1146 + \begin{acknowledgments}
1147 +  JDG acknowledges helpful discussions with Christopher
1148 +  Fennell. Support for this project was provided by the National
1149 +  Science Foundation under grant CHE-1362211. Computational time was
1150 +  provided by the Center for Research Computing (CRC) at the
1151 +  University of Notre Dame.
1152 + \end{acknowledgments}
1153 +
1154   %\bibliographystyle{aip}
1155   \newpage
1156   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines