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

Comparing trunk/multipole/multipole_2/multipole2.tex (file contents):
Revision 4174 by gezelter, Thu Jun 5 19:55:14 2014 UTC vs.
Revision 4203 by gezelter, Wed Aug 6 19:10:04 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 properties of the new
77 >  methods in molecular dynamics simulations. The SP method shows
78 >  excellent agreement with configurational energy differences, forces,
79 >  and torques, and would be suitable for use in Monte Carlo
80 >  calculations.  Of the two new shifted-force methods, the GSF
81 >  approach shows the best agreement with Ewald-derived energies,
82 >  forces, and torques and also exhibits energy conservation properties
83 >  that make it an excellent choice for efficient computation of
84 >  electrostatic interactions in molecular dynamics simulations.
85   \end{abstract}
86  
87 < \pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
87 > %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
88                               % Classification Scheme.
89 < \keywords{Electrostatics, Multipoles, Real-space}
89 > %\keywords{Electrostatics, Multipoles, Real-space}
90  
91   \maketitle
92  
77
93   \section{\label{sec:intro}Introduction}
94   Computing the interactions between electrostatic sites is one of the
95 < most expensive aspects of molecular simulations, which is why there
96 < have been significant efforts to develop practical, efficient and
97 < convergent methods for handling these interactions. Ewald's method is
98 < perhaps the best known and most accurate method for evaluating
99 < energies, forces, and torques in explicitly-periodic simulation
100 < cells. In this approach, the conditionally convergent electrostatic
101 < energy is converted into two absolutely convergent contributions, one
102 < which is carried out in real space with a cutoff radius, and one in
103 < reciprocal space.\cite{Clarke:1986eu,Woodcock75}
95 > most expensive aspects of molecular simulations. There have been
96 > significant efforts to develop practical, efficient and convergent
97 > methods for handling these interactions. Ewald's method is perhaps the
98 > best known and most accurate method for evaluating energies, forces,
99 > and torques in explicitly-periodic simulation cells. In this approach,
100 > the conditionally convergent electrostatic energy is converted into
101 > two absolutely convergent contributions, one which is carried out in
102 > real space with a cutoff radius, and one in reciprocal
103 > space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
104  
105   When carried out as originally formulated, the reciprocal-space
106   portion of the Ewald sum exhibits relatively poor computational
107 < scaling, making it prohibitive for large systems. By utilizing
108 < particle meshes and three dimensional fast Fourier transforms (FFT),
109 < the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
110 < (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
111 < the computational cost from $O(N^2)$ down to $O(N \log
112 < N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
107 > scaling, making it prohibitive for large systems. By utilizing a
108 > particle mesh and three dimensional fast Fourier transforms (FFT), the
109 > particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
110 > (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
111 > methods can decrease the computational cost from $O(N^2)$ down to $O(N
112 > \log
113 > N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
114  
115 < Because of the artificial periodicity required for the Ewald sum, the
100 < method may require modification to compute interactions for
115 > Because of the artificial periodicity required for the Ewald sum,
116   interfacial molecular systems such as membranes and liquid-vapor
117 < interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
118 < To simulate interfacial systems, Parry’s extension of the 3D Ewald sum
119 < is appropriate for slab geometries.\cite{Parry:1975if} The inherent
120 < periodicity in the Ewald’s method can also be problematic for
121 < interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
122 < methods that were developed to handle two-dimensional (2D)
123 < electrostatic interactions in interfacial systems have not had similar
124 < particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
125 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
117 > interfaces require modifications to the method.  Parry's extension of
118 > the three dimensional Ewald sum is appropriate for slab
119 > geometries.\cite{Parry:1975if} Modified Ewald methods that were
120 > developed to handle two-dimensional (2-D) electrostatic
121 > interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
122 > These methods were originally quite computationally
123 > expensive.\cite{Spohr97,Yeh99} There have been several successful
124 > efforts that reduced the computational cost of 2-D lattice summations,
125 > bringing them more in line with the scaling for the full 3-D
126 > treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
127 > inherent periodicity required by the Ewald method can also be
128 > problematic in a number of protein/solvent and ionic solution
129 > environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
130  
131   \subsection{Real-space methods}
132   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
133   method for calculating electrostatic interactions between point
134 < charges. They argued that the effective Coulomb interaction in
135 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
136 < an ordered lattice (e.g. when computing the Madelung constant of an
137 < ionic solid), the material can be considered as a set of ions
138 < interacting with neutral dipolar or quadrupolar ``molecules'' giving
139 < an effective distance dependence for the electrostatic interactions of
140 < $r^{-5}$ (see figure \ref{fig:NaCl}.  For this reason, careful
141 < applications of Wolf's method are able to obtain accurate estimates of
142 < Madelung constants using relatively short cutoff radii.  Recently,
143 < Fukuda used neutralization of the higher order moments for the
144 < calculation of the electrostatic interaction of the point charges
145 < system.\cite{Fukuda:2013sf}
134 > charges. They argued that the effective Coulomb interaction in most
135 > condensed phase systems is effectively short
136 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
137 > computing the Madelung constant of an ionic solid), the material can
138 > be considered as a set of ions interacting with neutral dipolar or
139 > quadrupolar ``molecules'' giving an effective distance dependence for
140 > the electrostatic interactions of $r^{-5}$ (see figure
141 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
142 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
143 > electrostatic energy per ion converges more rapidly to the Madelung
144 > energy than the dipolar approximation.\cite{Wolf92} To find the
145 > correct Madelung constant, Lacman suggested that the NaCl structure
146 > could be constructed in a way that the finite crystal terminates with
147 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
148 > what is effectively a set of octupoles at large distances. These facts
149 > suggest that the Madelung constants are relatively short ranged for
150 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
151 > application of Wolf's method can provide accurate estimates of
152 > Madelung constants using relatively short cutoff radii.
153  
154 < \begin{figure}[h!]
154 > Direct truncation of interactions at a cutoff radius creates numerical
155 > errors.  Wolf \textit{et al.} suggest that truncation errors are due
156 > to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
157 > neutralize this charge they proposed placing an image charge on the
158 > surface of the cutoff sphere for every real charge inside the cutoff.
159 > These charges are present for the evaluation of both the pair
160 > interaction energy and the force, although the force expression
161 > maintains a discontinuity at the cutoff sphere.  In the original Wolf
162 > formulation, the total energy for the charge and image were not equal
163 > to the integral of the force expression, and as a result, the total
164 > energy would not be conserved in molecular dynamics (MD)
165 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
166 > Gezelter later proposed shifted force variants of the Wolf method with
167 > commensurate force and energy expressions that do not exhibit this
168 > problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
169 > were also proposed by Chen \textit{et
170 >  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
171 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
172 > used additional neutralization of higher order moments for systems of
173 > point charges.\cite{Fukuda:2013sf}
174 >
175 > \begin{figure}
176    \centering
177 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
178 <  \caption{Top: NaCl crystal showing how spherical truncation can
179 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
180 <    molecules interact with the central ion.  Bottom: A dipolar
181 <    crystal exhibiting similar behavior and illustrating how the
182 <    effective dipole-octupole interactions can be disrupted by
183 <    spherical truncation.}
184 <  \label{fig:NaCl}
177 >  \includegraphics[width=\linewidth]{schematic.eps}
178 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
179 >    charges (in the smaller grey circle), so interactions are
180 >    effectively charge-multipole at longer distances.  With hard
181 >    cutoffs, motion of individual charges in and out of the cutoff
182 >    sphere can break the effective multipolar ordering.  Bottom:
183 >    dipolar crystals and fluids have a similar effective
184 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
185 >    orientational averaging helps to reduce the effective range of the
186 >    interactions in the fluid.  Placement of reversed image multipoles
187 >    on the surface of the cutoff sphere recovers the effective
188 >    higher-order multipole behavior.}
189 >  \label{fig:schematic}
190   \end{figure}
191  
192 < The direct truncation of interactions at a cutoff radius creates
193 < truncation defects. Wolf \textit{et al.} further argued that
194 < truncation errors are due to net charge remaining inside the cutoff
195 < sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
196 < placing an image charge on the surface of the cutoff sphere for every
197 < real charge inside the cutoff.  These charges are present for the
198 < evaluation of both the pair interaction energy and the force, although
199 < the force expression maintained a discontinuity at the cutoff sphere.
200 < In the original Wolf formulation, the total energy for the charge and
201 < 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
192 > One can make a similar effective range argument for crystals of point
193 > \textit{multipoles}. The Luttinger and Tisza treatment of energy
194 > constants for dipolar lattices utilizes 24 basis vectors that contain
195 > dipoles at the eight corners of a unit cube.\cite{LT} Only three of
196 > these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
197 > moments, while the rest have zero net dipole and retain contributions
198 > only from higher order multipoles.  The lowest-energy crystalline
199 > structures are built out of basis vectors that have only residual
200 > quadrupolar moments (e.g. the $Z_5$ array). In these low energy
201 > structures, the effective interaction between a dipole at the center
202   of a crystal and a group of eight dipoles farther away is
203   significantly shorter ranged than the $r^{-3}$ that one would expect
204   for raw dipole-dipole interactions.  Only in crystals which retain a
# Line 188 | Line 208 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
208   unstable.
209  
210   In ionic crystals, real-space truncation can break the effective
211 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
212 < swings in the electrostatic energy as the cutoff radius is increased
213 < (or as individual ions move back and forth across the boundary).  This
214 < is why the image charges were necessary for the Wolf sum to exhibit
215 < rapid convergence.  Similarly, the real-space truncation of point
216 < multipole interactions breaks higher order multipole arrangements, and
217 < image multipoles are required for real-space treatments of
198 < electrostatic energies.
211 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
212 > significant swings in the electrostatic energy as individual ions move
213 > back and forth across the boundary.  This is why the image charges are
214 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
215 > the real-space truncation of point multipole interactions breaks
216 > higher order multipole arrangements, and image multipoles are required
217 > for real-space treatments of electrostatic energies.
218  
219 + The shorter effective range of electrostatic interactions is not
220 + limited to perfect crystals, but can also apply in disordered fluids.
221 + Even at elevated temperatures, there is local charge balance in an
222 + ionic liquid, where each positive ion has surroundings dominated by
223 + negaitve ions and vice versa.  The reversed-charge images on the
224 + cutoff sphere that are integral to the Wolf and DSF approaches retain
225 + the effective multipolar interactions as the charges traverse the
226 + cutoff boundary.
227 +
228 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
229 + significant orientational averaging that additionally reduces the
230 + effect of long-range multipolar interactions.  The image multipoles
231 + that are introduced in the TSF, GSF, and SP methods mimic this effect
232 + and reduce the effective range of the multipolar interactions as
233 + interacting molecules traverse each other's cutoff boundaries.
234 +
235   % Because of this reason, although the nature of electrostatic
236   % interaction short ranged, the hard cutoff sphere creates very large
237   % fluctuation in the electrostatic energy for the perfect crystal. In
# Line 207 | Line 242 | The forces and torques acting on atomic sites are the
242   % to the non-neutralized value of the higher order moments within the
243   % cutoff sphere.
244  
245 < The forces and torques acting on atomic sites are the fundamental
246 < factors driving dynamics in molecular simulations. Fennell and
247 < Gezelter proposed the damped shifted force (DSF) energy kernel to
248 < obtain consistent energies and forces on the atoms within the cutoff
249 < sphere. Both the energy and the force go smoothly to zero as an atom
250 < aproaches the cutoff radius. The comparisons of the accuracy these
251 < quantities between the DSF kernel and SPME was surprisingly
252 < good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
253 < calculating electrostatic interactions in molecular systems with
254 < relatively uniform charge
220 < densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
245 > Forces and torques acting on atomic sites are fundamental in driving
246 > dynamics in molecular simulations, and the damped shifted force (DSF)
247 > energy kernel provides consistent energies and forces on charged atoms
248 > within the cutoff sphere. Both the energy and the force go smoothly to
249 > zero as an atom aproaches the cutoff radius. The comparisons of the
250 > accuracy these quantities between the DSF kernel and SPME was
251 > surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
252 > has seen increasing use in molecular systems with relatively uniform
253 > charge
254 > densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
255  
256   \subsection{The damping function}
257 < The damping function used in our research has been discussed in detail
258 < in the first paper of this series.\cite{PaperI} The radial kernel
259 < $1/r$ for the interactions between point charges can be replaced by
260 < the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
261 < accelerate the rate of convergence, where $\alpha$ is a damping
262 < parameter with units of inverse distance.  Altering the value of
263 < $\alpha$ is equivalent to changing the width of Gaussian charge
264 < distributions that replace each point charge -- Gaussian overlap
265 < integrals yield complementary error functions when truncated at a
266 < finite distance.
257 > The damping function has been discussed in detail in the first paper
258 > of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
259 > interactions between point charges can be replaced by the
260 > complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
261 > convergence, where $\alpha$ is a damping parameter with units of
262 > inverse distance.  Altering the value of $\alpha$ is equivalent to
263 > changing the width of Gaussian charge distributions that replace each
264 > point charge, as Coulomb integrals with Gaussian charge distributions
265 > produce complementary error functions when truncated at a finite
266 > distance.
267  
268 < By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
269 < cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
270 < agreement with SPME for the interaction energies, forces and torques
271 < for charge-charge interactions.\cite{Fennell:2006lq}
268 > With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
269 > produced very good agreement with SPME for interaction energies,
270 > forces and torques for charge-charge
271 > interactions.\cite{Fennell:2006lq}
272  
273   \subsection{Point multipoles in molecular modeling}
274   Coarse-graining approaches which treat entire molecular subsystems as
275   a single rigid body are now widely used. A common feature of many
276   coarse-graining approaches is simplification of the electrostatic
277   interactions between bodies so that fewer site-site interactions are
278 < required to compute configurational energies.  Many coarse-grained
279 < 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}
278 > required to compute configurational
279 > energies.\cite{Ren06,Essex10,Essex11}
280  
281 < Because electrons in a molecule are not localized at specific points,
282 < the assignment of partial charges to atomic centers is a relatively
283 < rough approximation.  Atomic sites can also be assigned point
284 < multipoles and polarizabilities to increase the accuracy of the
285 < molecular model.  Recently, water has been modeled with point
286 < multipoles up to octupolar
287 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
281 > Additionally, because electrons in a molecule are not localized at
282 > specific points, the assignment of partial charges to atomic centers
283 > is always an approximation.  For increased accuracy, atomic sites can
284 > also be assigned point multipoles and polarizabilities.  Recently,
285 > water has been modeled with point multipoles up to octupolar order
286 > using the soft sticky dipole-quadrupole-octupole (SSDQO)
287 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
288   multipoles up to quadrupolar order have also been coupled with point
289   polarizabilities in the high-quality AMOEBA and iAMOEBA water
290 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
291 < using point multipole with the real space truncation without
292 < accounting for multipolar neutrality will create energy conservation
293 < issues in molecular dynamics (MD) simulations.
290 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
291 > truncating point multipoles without smoothing the forces and torques
292 > can create energy conservation issues in molecular dynamics
293 > simulations.
294  
295   In this paper we test a set of real-space methods that were developed
296   for point multipolar interactions.  These methods extend the damped
297   shifted force (DSF) and Wolf methods originally developed for
298   charge-charge interactions and generalize them for higher order
299 < multipoles. The detailed mathematical development of these methods has
300 < been presented in the first paper in this series, while this work
301 < covers the testing the energies, forces, torques, and energy
299 > multipoles.  The detailed mathematical development of these methods
300 > has been presented in the first paper in this series, while this work
301 > covers the testing of energies, forces, torques, and energy
302   conservation properties of the methods in realistic simulation
303   environments.  In all cases, the methods are compared with the
304 < reference method, a full multipolar Ewald treatment.
304 > reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
305  
306  
307   %\subsection{Conservation of total energy }
# Line 297 | Line 327 | where the multipole operator for site $\bf a$,
327   \begin{equation}
328   U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
329   \end{equation}
330 < where the multipole operator for site $\bf a$,
331 < \begin{equation}
332 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
333 < +  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.
330 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
331 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
332 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
333 > $\bf a$, etc.
334  
335 < Interactions between multipoles can be expressed as higher derivatives
336 < of the bare Coulomb potential, so one way of ensuring that the forces
337 < and torques vanish at the cutoff distance is to include a larger
338 < number of terms in the truncated Taylor expansion, e.g.,
339 < %
340 < \begin{equation}
341 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
342 < \end{equation}
343 < %
344 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
345 < Thus, for $f(r)=1/r$, we find
346 < %
347 < \begin{equation}
348 < 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} .
349 < \end{equation}
350 < This function is an approximate electrostatic potential that has
351 < vanishing second derivatives at the cutoff radius, making it suitable
352 < for shifting the forces and torques of charge-dipole interactions.
335 > % Interactions between multipoles can be expressed as higher derivatives
336 > % of the bare Coulomb potential, so one way of ensuring that the forces
337 > % and torques vanish at the cutoff distance is to include a larger
338 > % number of terms in the truncated Taylor expansion, e.g.,
339 > % %
340 > % \begin{equation}
341 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
342 > % \end{equation}
343 > % %
344 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
345 > % Thus, for $f(r)=1/r$, we find
346 > % %
347 > % \begin{equation}
348 > % 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} .
349 > % \end{equation}
350 > % This function is an approximate electrostatic potential that has
351 > % vanishing second derivatives at the cutoff radius, making it suitable
352 > % for shifting the forces and torques of charge-dipole interactions.
353  
354 < In general, the TSF potential for any multipole-multipole interaction
355 < can be written
354 > The TSF potential for any multipole-multipole interaction can be
355 > written
356   \begin{equation}
357   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
358   \label{generic}
359   \end{equation}
360 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
361 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
362 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
363 < energy, force, and torques, the required number of terms from Taylor
364 < series expansion in $f_n(r)$ must be performed for different
365 < multipole-multipole interactions.
360 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
361 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
362 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
363 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
364 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
365 > force, and torques, a Taylor expansion with $n$ terms must be
366 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
367  
368 < To carry out the same procedure for a damped electrostatic kernel, we
369 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
370 < Many of the derivatives of the damped kernel are well known from
371 < Smith's early work on multipoles for the Ewald
372 < summation.\cite{Smith82,Smith98}
368 > % To carry out the same procedure for a damped electrostatic kernel, we
369 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
370 > % Many of the derivatives of the damped kernel are well known from
371 > % Smith's early work on multipoles for the Ewald
372 > % summation.\cite{Smith82,Smith98}
373  
374 < Note that increasing the value of $n$ will add additional terms to the
375 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
376 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
377 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
378 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
379 < for computing multipole energies, forces, and torques, and smooth
380 < cutoffs of these quantities can be guaranteed as long as the number of
381 < terms in the Taylor series exceeds the derivative order required.
374 > % Note that increasing the value of $n$ will add additional terms to the
375 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
376 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
377 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
378 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
379 > % for computing multipole energies, forces, and torques, and smooth
380 > % cutoffs of these quantities can be guaranteed as long as the number of
381 > % terms in the Taylor series exceeds the derivative order required.
382  
383   For multipole-multipole interactions, following this procedure results
384 < in separate radial functions for each distinct orientational
385 < contribution to the potential, and ensures that the forces and torques
386 < from {\it each} of these contributions will vanish at the cutoff
387 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
388 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
384 > in separate radial functions for each of the distinct orientational
385 > contributions to the potential, and ensures that the forces and
386 > torques from each of these contributions will vanish at the cutoff
387 > radius.  For example, the direct dipole dot product
388 > ($\mathbf{D}_{\bf a}
389 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
390   dot products:
391   \begin{equation}
392 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
393 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
394 < -\frac{1}{4\pi \epsilon_0}
395 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
396 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
392 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
393 >  \mathbf{D}_{\bf a} \cdot
394 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
395 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
396 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
397   \end{equation}
398  
399 < The electrostatic forces and torques acting on the central multipole
400 < site due to another site within cutoff sphere are derived from
399 > For the Taylor shifted (TSF) method with the undamped kernel,
400 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
401 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
402 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
403 > connection to unmodified electrostatics as well as the smooth
404 > transition to zero in both these functions as $r\rightarrow r_c$.  The
405 > electrostatic forces and torques acting on the central multipole due
406 > to another site within the cutoff sphere are derived from
407   Eq.~\ref{generic}, accounting for the appropriate number of
408   derivatives. Complete energy, force, and torque expressions are
409   presented in the first paper in this series (Reference
410 < \citep{PaperI}).
410 > \onlinecite{PaperI}).
411  
412   \subsection{Gradient-shifted force (GSF)}
413  
414 < A second (and significantly simpler) method involves shifting the
415 < gradient of the raw coulomb potential for each particular multipole
414 > A second (and conceptually simpler) method involves shifting the
415 > gradient of the raw Coulomb potential for each particular multipole
416   order.  For example, the raw dipole-dipole potential energy may be
417   shifted smoothly by finding the gradient for two interacting dipoles
418   which have been projected onto the surface of the cutoff sphere
419   without changing their relative orientation,
420 < \begin{displaymath}
421 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
422 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
423 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
424 < \end{displaymath}
425 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
426 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
427 < the signs are reversed for the dipole that has been projected onto the
428 < cutoff sphere).  In many ways, this simpler approach is closer in
429 < spirit to the original shifted force method, in that it projects a
430 < neutralizing multipole (and the resulting forces from this multipole)
431 < onto a cutoff sphere. The resulting functional forms for the
432 < potentials, forces, and torques turn out to be quite similar in form
433 < to the Taylor-shifted approach, although the radial contributions are
434 < significantly less perturbed by the Gradient-shifted approach than
435 < they are in the Taylor-shifted method.
420 > \begin{equation}
421 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
422 > U_{D_{\bf a} D_{\bf b}}(r_c)
423 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
424 >  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
425 > \end{equation}
426 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
427 >  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
428 > (although the signs are reversed for the dipole that has been
429 > projected onto the cutoff sphere).  In many ways, this simpler
430 > approach is closer in spirit to the original shifted force method, in
431 > that it projects a neutralizing multipole (and the resulting forces
432 > from this multipole) onto a cutoff sphere. The resulting functional
433 > forms for the potentials, forces, and torques turn out to be quite
434 > similar in form to the Taylor-shifted approach, although the radial
435 > contributions are significantly less perturbed by the gradient-shifted
436 > approach than they are in the Taylor-shifted method.
437  
438 + For the gradient shifted (GSF) method with the undamped kernel,
439 + $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
440 + $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
441 + Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
442 + because the Taylor expansion retains only one term, they are
443 + significantly less perturbed than the TSF functions.
444 +
445   In general, the gradient shifted potential between a central multipole
446   and any multipolar site inside the cutoff radius is given by,
447   \begin{equation}
448 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
449 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
450 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
448 >  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
449 >    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
450 >    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
451   \label{generic2}
452   \end{equation}
453   where the sum describes a separate force-shifting that is applied to
454 < each orientational contribution to the energy.
454 > each orientational contribution to the energy.  In this expression,
455 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
456 > ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
457 > represent the orientations the multipoles.
458  
459   The third term converges more rapidly than the first two terms as a
460   function of radius, hence the contribution of the third term is very
461   small for large cutoff radii.  The force and torque derived from
462 < equation \ref{generic2} are consistent with the energy expression and
463 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
462 > Eq. \ref{generic2} are consistent with the energy expression and
463 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
464   can be considered generalizations of the original DSF method for
465   higher order multipole interactions. GSF and TSF are also identical up
466   to the charge-dipole interaction but generate different expressions in
467   the energy, force and torque for higher order multipole-multipole
468   interactions. Complete energy, force, and torque expressions for the
469   GSF potential are presented in the first paper in this series
470 < (Reference \citep{PaperI})
470 > (Reference~\onlinecite{PaperI}).
471  
472  
473   \subsection{Shifted potential (SP) }
# Line 439 | Line 480 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
480   interactions with the central multipole and the image. This
481   effectively shifts the total potential to zero at the cutoff radius,
482   \begin{equation}
483 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
483 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
484 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
485   \label{eq:SP}
486   \end{equation}          
487   where the sum describes separate potential shifting that is done for
488   each orientational contribution to the energy (e.g. the direct dipole
489   product contribution is shifted {\it separately} from the
490   dipole-distance terms in dipole-dipole interactions).  Note that this
491 < is not a simple shifting of the total potential at $R_c$. Each radial
491 > is not a simple shifting of the total potential at $r_c$. Each radial
492   contribution is shifted separately.  One consequence of this is that
493   multipoles that reorient after leaving the cutoff sphere can re-enter
494   the cutoff sphere without perturbing the total energy.
495  
496 < The potential energy between a central multipole and other multipolar
497 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
498 < force and torque obtained from the shifted potential (SP) are
499 < discontinuous at $R_c$. Therefore, MD simulations will still
500 < experience energy drift while operating under the SP potential, but it
501 < may be suitable for Monte Carlo approaches where the configurational
502 < energy differences are the primary quantity of interest.
496 > For the shifted potential (SP) method with the undamped kernel,
497 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
498 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
499 > central multipole and other multipolar sites goes smoothly to zero as
500 > $r \rightarrow r_c$.  However, the force and torque obtained from the
501 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
502 > will still experience energy drift while operating under the SP
503 > potential, but it may be suitable for Monte Carlo approaches where the
504 > configurational energy differences are the primary quantity of
505 > interest.
506  
507 < \subsection{The Self term}
507 > \subsection{The Self Term}
508   In the TSF, GSF, and SP methods, a self-interaction is retained for
509   the central multipole interacting with its own image on the surface of
510   the cutoff sphere.  This self interaction is nearly identical with the
511   self-terms that arise in the Ewald sum for multipoles.  Complete
512   expressions for the self terms are presented in the first paper in
513 < this series (Reference \citep{PaperI})  
513 > this series (Reference \onlinecite{PaperI}).
514  
515  
516   \section{\label{sec:methodology}Methodology}
# Line 480 | Line 525 | disordered and ordered condensed-phase systems.  These
525   arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
526   used the multipolar Ewald sum as a reference method for comparing
527   energies, forces, and torques for molecular models that mimic
528 < disordered and ordered condensed-phase systems.  These test-cases
529 < include:
485 < \begin{itemize}
486 < \item Soft Dipolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $|D| = 2.35$)
487 < \item Soft Dipolar solids ($\sigma = 2.837$, $\epsilon =1.0$, $|D| = 2.35$)
488 < \item Soft Quadrupolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
489 < \item Soft Quadrupolar solids  ($\sigma = 2.837$, $\epsilon = 1.0$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
490 < \item A mixed multipole model (SSDQ) for water ($\sigma = 3.051$, $\epsilon = 0.152$, $D_z = 2.35$, $Q_{\alpha\alpha} =\left\{-1.35,0,-0.68\right\}$)
491 < \item A mixed multipole models for water with 48 dissolved ions, 24
492 <  \ce{Na+}: ($\sigma = 2.579$, $\epsilon =0.118$, $q = 1e$) and 24
493 <  \ce{Cl-}: ($\sigma = 4.445$, $\epsilon =0.1$l, $q = -1e$)
494 < \end{itemize}
495 < All Lennard-Jones parameters are in units of \AA\ $(\sigma)$ and kcal
496 < / mole $(\epsilon)$.  Partial charges are reported in electrons, while
497 < dipoles are in Debye units, and quadrupoles are in units of Debye-\AA.
528 > disordered and ordered condensed-phase systems.  The parameters used
529 > in the test cases are given in table~\ref{tab:pars}.
530  
531 < The last test case exercises all levels of the multipole-multipole
532 < interactions we have derived so far and represents the most complete
533 < test of the new methods.  In the following section, we present results
534 < for the total electrostatic energy, as well as the electrostatic
535 < contributions to the force and torque on each molecule.  These
536 < quantities have been computed using the SP, TSF, and GSF methods, as
537 < well as a hard cutoff, and have been compared with the values obtaine
538 < from the multipolar Ewald sum.  In Mote Carlo (MC) simulations, the
539 < energy differences between two configurations is the primary quantity
540 < that governs how the simulation proceeds. These differences are the
541 < most imporant indicators of the reliability of a method even if the
542 < absolute energies are not exact.  For each of the multipolar systems
543 < listed above, we have compared the change in electrostatic potential
544 < energy ($\Delta E$) between 250 statistically-independent
545 < configurations.  In molecular dynamics (MD) simulations, the forces
546 < and torques govern the behavior of the simulation, so we also compute
547 < the electrostatic contributions to the forces and torques.
531 > \begin{table}
532 > \label{tab:pars}
533 > \caption{The parameters used in the systems used to evaluate the new
534 >  real-space methods.  The most comprehensive test was a liquid
535 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
536 >  ions).  This test excercises all orders of the multipolar
537 >  interactions developed in the first paper.}
538 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
539 >             & \multicolumn{2}{c|}{LJ parameters} &
540 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
541 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
542 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
543 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
544 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
545 > \AA\textsuperscript{2})} \\ \hline
546 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
547 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
548 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
549 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
550 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
551 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
552 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
553 > \end{tabularx}
554 > \end{table}
555 > The systems consist of pure multipolar solids (both dipole and
556 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
557 > fluid composed of sites containing both dipoles and quadrupoles
558 > simultaneously, and a final test case that includes ions with point
559 > charges in addition to the multipolar fluid.  The solid-phase
560 > parameters were chosen so that the systems can explore some
561 > orientational freedom for the multipolar sites, while maintaining
562 > relatively strict translational order.  The SSDQ model used here is
563 > not a particularly accurate water model, but it does test
564 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
565 > interactions at roughly the same magnitudes. The last test case, SSDQ
566 > water with dissolved ions, exercises \textit{all} levels of the
567 > multipole-multipole interactions we have derived so far and represents
568 > the most complete test of the new methods.
569  
570 < \subsection{Model systems}
571 < To sample independent configurations of multipolar crystals, a body
572 < centered cubic (bcc) crystal which is a minimum energy structure for
573 < point dipoles was generated using 3,456 molecules.  The multipoles
574 < were translationally locked in their respective crystal sites for
575 < equilibration at a relatively low temperature (50K), so that dipoles
576 < or quadrupoles could freely explore all accessible orientations.  The
577 < translational constraints were removed, and the crystals were
578 < simulated for 10 ps in the microcanonical (NVE) ensemble with an
579 < average temperature of 50 K.  Configurations were sampled at equal
580 < time intervals for the comparison of the configurational energy
581 < differences.  The crystals were not simulated close to the melting
582 < points in order to avoid translational deformation away of the ideal
583 < lattice geometry.
570 > In the following section, we present results for the total
571 > electrostatic energy, as well as the electrostatic contributions to
572 > the force and torque on each molecule.  These quantities have been
573 > computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
574 > and have been compared with the values obtained from the multipolar
575 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
576 > between two configurations is the primary quantity that governs how
577 > the simulation proceeds. These differences are the most important
578 > indicators of the reliability of a method even if the absolute
579 > energies are not exact.  For each of the multipolar systems listed
580 > above, we have compared the change in electrostatic potential energy
581 > ($\Delta E$) between 250 statistically-independent configurations.  In
582 > molecular dynamics (MD) simulations, the forces and torques govern the
583 > behavior of the simulation, so we also compute the electrostatic
584 > contributions to the forces and torques.
585  
586 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
587 < system was created with 2048 molecules oriented randomly.  These were
586 > \subsection{Implementation}
587 > The real-space methods developed in the first paper in this series
588 > have been implemented in our group's open source molecular simulation
589 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
590 > this work.  The complementary error function can be a relatively slow
591 > function on some processors, so all of the radial functions are
592 > precomputed on a fine grid and are spline-interpolated to provide
593 > values when required.  
594  
595 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
596 < temperature after equilibration.  We collected 250 different
597 < configurations in equal interval of time. For the ions mixed liquid
598 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
599 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
600 < at the same environment for 1ns and 250 configurations were
601 < collected. While comparing energies, forces, and torques with Ewald
602 < method, Lennard-Jones potentials were turned off and purely
543 < electrostatic interaction had been compared.
595 > Using the same simulation code, we compare to a multipolar Ewald sum
596 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
597 > the Ewald sum is a re-implementation of the algorithm originally
598 > proposed by Smith that does not use the particle mesh or smoothing
599 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
600 > being compared are the electrostatic contributions to energies, force,
601 > and torques.  All other contributions to these quantities (i.e. from
602 > Lennard-Jones interactions) are removed prior to the comparisons.
603  
604 + The convergence parameter ($\alpha$) also plays a role in the balance
605 + of the real-space and reciprocal-space portions of the Ewald
606 + calculation.  Typical molecular mechanics packages set this to a value
607 + that depends on the cutoff radius and a tolerance (typically less than
608 + $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
609 + associated with increasing accuracy at the expense of computational
610 + time spent on the reciprocal-space portion of the
611 + summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
612 + 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
613 + Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
614 +
615 + The real-space models have self-interactions that provide
616 + contributions to the energies only.  Although the self interaction is
617 + a rapid calculation, we note that in systems with fluctuating charges
618 + or point polarizabilities, the self-term is not static and must be
619 + recomputed at each time step.
620 +
621 + \subsection{Model systems}
622 + To sample independent configurations of the multipolar crystals, body
623 + centered cubic (bcc) crystals, which exhibit the minimum energy
624 + structures for point dipoles, were generated using 3,456 molecules.
625 + The multipoles were translationally locked in their respective crystal
626 + sites for equilibration at a relatively low temperature (50K) so that
627 + dipoles or quadrupoles could freely explore all accessible
628 + orientations.  The translational constraints were then removed, the
629 + systems were re-equilibrated, and the crystals were simulated for an
630 + additional 10 ps in the microcanonical (NVE) ensemble with an average
631 + temperature of 50 K.  The balance between moments of inertia and
632 + particle mass were chosen to allow orientational sampling without
633 + significant translational motion.  Configurations were sampled at
634 + equal time intervals in order to compare configurational energy
635 + differences.  The crystals were simulated far from the melting point
636 + in order to avoid translational deformation away of the ideal lattice
637 + geometry.
638 +
639 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
640 + simulations, each system was created with 2,048 randomly-oriented
641 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
642 + Each system was then simulated for 1 ns in the microcanonical (NVE)
643 + ensemble.  We collected 250 different configurations at equal time
644 + intervals. For the liquid system that included ionic species, we
645 + converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
646 + \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
647 + run under the same conditions for 1 ns. A total of 250 configurations
648 + were collected. In the following comparisons of energies, forces, and
649 + torques, the Lennard-Jones potentials were turned off and only the
650 + purely electrostatic quantities were compared with the same values
651 + obtained via the Ewald sum.
652 +
653   \subsection{Accuracy of Energy Differences, Forces and Torques}
654   The pairwise summation techniques (outlined above) were evaluated for
655   use in MC simulations by studying the energy differences between
# Line 554 | Line 662 | we used least square regressions analysiss for the six
662   should be identical for all methods.
663  
664   Since none of the real-space methods provide exact energy differences,
665 < we used least square regressions analysiss for the six different
665 > we used least square regressions analysis for the six different
666   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
667   with the multipolar Ewald reference method.  Unitary results for both
668   the correlation (slope) and correlation coefficient for these
# Line 565 | Line 673 | also been compared by using least squares regression a
673   configurations and 250 configurations were recorded for comparison.
674   Each system provided 31,125 energy differences for a total of 186,750
675   data points.  Similarly, the magnitudes of the forces and torques have
676 < also been compared by using least squares regression analyses. In the
676 > also been compared using least squares regression analysis. In the
677   forces and torques comparison, the magnitudes of the forces acting in
678   each molecule for each configuration were evaluated. For example, our
679   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 685 | Line 793 | model must allow for long simulation times with minima
793  
794   \begin{figure}
795    \centering
796 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
796 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
797    \caption{Statistical analysis of the quality of configurational
798      energy differences for the real-space electrostatic methods
799      compared with the reference Ewald sum.  Results with a value equal
# Line 699 | Line 807 | reproduce the Ewald-derived configurational energy dif
807  
808   The combined correlation coefficient and slope for all six systems is
809   shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
810 < reproduce the Ewald-derived configurational energy differences with
811 < remarkable fidelity.  Undamped hard cutoffs introduce a significant
812 < amount of random scatter in the energy differences which is apparent
813 < in the reduced value of the correlation coefficient for this method.
814 < This can be understood easily as configurations which exhibit only
815 < small traversals of a few dipoles or quadrupoles out of the cutoff
816 < sphere will see large energy jumps when hard cutoffs are used.  The
810 > reproduce the Ewald configurational energy differences with remarkable
811 > fidelity.  Undamped hard cutoffs introduce a significant amount of
812 > random scatter in the energy differences which is apparent in the
813 > reduced value of the correlation coefficient for this method.  This
814 > can be easily understood as configurations which exhibit small
815 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
816 > will see large energy jumps when hard cutoffs are used.  The
817   orientations of the multipoles (particularly in the ordered crystals)
818 < mean that these jumps can go either up or down in energy, producing a
819 < significant amount of random scatter.
818 > mean that these energy jumps can go in either direction, producing a
819 > significant amount of random scatter, but no systematic error.
820  
821   The TSF method produces energy differences that are highly correlated
822   with the Ewald results, but it also introduces a significant
# Line 716 | Line 824 | effect on crystalline systems.
824   smaller cutoff values. The TSF method alters the distance dependence
825   of different orientational contributions to the energy in a
826   non-uniform way, so the size of the cutoff sphere can have a large
827 < effect on crystalline systems.
827 > effect, particularly for the crystalline systems.
828  
829   Both the SP and GSF methods appear to reproduce the Ewald results with
830   excellent fidelity, particularly for moderate damping ($\alpha =
831 < 0.1-0.2$\AA$^{-1}$) and commonly-used cutoff values ($r_c = 12$\AA).
832 < With the exception of the undamped hard cutoff, and the TSF method
833 < with short cutoffs, all of the methods would be appropriate for use in
834 < Monte Carlo simulations.
831 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
832 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
833 > method with short cutoffs, all of the methods would be appropriate for
834 > use in Monte Carlo simulations.
835  
836   \subsection{Magnitude of the force and torque vectors}
837  
838 < The comparison of the magnitude of the combined forces and torques for
839 < the data accumulated from all system types are shown in Figures
838 > The comparisons of the magnitudes of the forces and torques for the
839 > data accumulated from all six systems are shown in Figures
840   ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
841   correlation and slope for the forces agree well with the Ewald sum
842 < even for the hard cutoff method.
842 > even for the hard cutoffs.
843  
844 < For the system of molecules with higher order multipoles, the
845 < interaction is quite short ranged. Moreover, the force decays more
846 < rapidly than the electrostatic energy hence the hard cutoff method can
847 < also produces reasonable agreement.  Although the pure cutoff gives
848 < the good match of the electrostatic force for pairs of molecules
849 < included within the cutoff sphere, the discontinuity in the force at
850 < the cutoff radius can potentially cause problems the total energy
851 < conservation as molecules enter and leave the cutoff sphere.  This is
852 < discussed in detail in section \ref{sec:}.
844 > For systems of molecules with only multipolar interactions, the pair
845 > energy contributions are quite short ranged.  Moreover, the force
846 > decays more rapidly than the electrostatic energy, hence the hard
847 > cutoff method can also produce reasonable agreement for this quantity.
848 > Although the pure cutoff gives reasonably good electrostatic forces
849 > for pairs of molecules included within each other's cutoff spheres,
850 > the discontinuity in the force at the cutoff radius can potentially
851 > cause energy conservation problems as molecules enter and leave the
852 > cutoff spheres.  This is discussed in detail in section
853 > \ref{sec:conservation}.
854  
855   The two shifted-force methods (GSF and TSF) exhibit a small amount of
856   systematic variation and scatter compared with the Ewald forces.  The
857   shifted-force models intentionally perturb the forces between pairs of
858 < molecules inside the cutoff sphere in order to correct the energy
859 < conservation issues, so it is not particularly surprising that this
860 < perturbation is evident in these same molecular forces.  The GSF
861 < perturbations are minimal, particularly for moderate damping and and
858 > molecules inside each other's cutoff spheres in order to correct the
859 > energy conservation issues, and this perturbation is evident in the
860 > statistics accumulated for the molecular forces.  The GSF
861 > perturbations are minimal, particularly for moderate damping and
862   commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
863   reasonable agreement in the correlation coefficient but again the
864   systematic error in the forces is concerning if replication of Ewald
# Line 757 | Line 866 | forces is desired.
866  
867   \begin{figure}
868    \centering
869 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
869 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
870    \caption{Statistical analysis of the quality of the force vector
871      magnitudes for the real-space electrostatic methods compared with
872      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 771 | Line 880 | forces is desired.
880  
881   \begin{figure}
882    \centering
883 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
883 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
884    \caption{Statistical analysis of the quality of the torque vector
885      magnitudes for the real-space electrostatic methods compared with
886      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 791 | Line 900 | The other real-space methods can cause some significan
900  
901   The results shows that the torque from the hard cutoff method
902   reproduces the torques in quite good agreement with the Ewald sum.
903 < The other real-space methods can cause some significant deviations,
904 < but excellent agreement with the Ewald sum torques is recovered at
905 < moderate values of the damping coefficient ($\alpha =
906 < 0.1-0.2$\AA$^{-1}$) and cutoff radius ($r_c \ge 12$\AA).  The TSF
907 < method exhibits the only fair agreement in the slope as compared to
908 < Ewald even for larger cutoff radii.  It appears that the severity of
909 < the perturbations in the TSF method are most apparent in the torques.
903 > The other real-space methods can cause some deviations, but excellent
904 > agreement with the Ewald sum torques is recovered at moderate values
905 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
906 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
907 > in the slope when compared with the Ewald torques even for larger
908 > cutoff radii.  It appears that the severity of the perturbations in
909 > the TSF method are most in evidence for the torques.
910  
911   \subsection{Directionality of the force and torque vectors}  
912  
# Line 806 | Line 915 | directionality is shown in terms of circular variance
915   these quantities. Force and torque vectors for all six systems were
916   analyzed using Fisher statistics, and the quality of the vector
917   directionality is shown in terms of circular variance
918 < ($\mathrm{Var}(\theta$) in figure
918 > ($\mathrm{Var}(\theta)$) in figure
919   \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
920 < from the new real-space method exhibit nearly-ideal Fisher probability
920 > from the new real-space methods exhibit nearly-ideal Fisher probability
921   distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
922   exhibit the best vectorial agreement with the Ewald sum. The force and
923   torque vectors from GSF method also show good agreement with the Ewald
924   method, which can also be systematically improved by using moderate
925 < damping and a reasonable cutoff radius.  For $\alpha = 0.2$ and $r_c =
925 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
926   12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
927 < to a distribution with 95\% of force vectors within $6.37^\circ$ of the
928 < corresponding Ewald forces. The TSF method produces the poorest
927 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
928 > the corresponding Ewald forces. The TSF method produces the poorest
929   agreement with the Ewald force directions.
930  
931 < Torques are again more perturbed by the new real-space methods, than
932 < forces, but even here the variance is reasonably small.  For the same
931 > Torques are again more perturbed than the forces by the new real-space
932 > methods, but even here the variance is reasonably small.  For the same
933   method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
934   the circular variance was 0.01415, corresponds to a distribution which
935   has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
# Line 829 | Line 938 | systematically improved by varying $\alpha$ and $r_c$.
938  
939   \begin{figure}
940    \centering
941 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
941 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
942    \caption{The circular variance of the direction of the force and
943      torque vectors obtained from the real-space methods around the
944      reference Ewald vectors. A variance equal to 0 (dashed line)
# Line 840 | Line 949 | systematically improved by varying $\alpha$ and $r_c$.
949    \label{fig:slopeCorr_circularVariance}
950   \end{figure}
951  
952 < \subsection{Energy conservation}
952 > \subsection{Energy conservation\label{sec:conservation}}
953  
954   We have tested the conservation of energy one can expect to see with
955   the new real-space methods using the SSDQ water model with a small
# Line 848 | Line 957 | and 48 dissolved ions at a density of 0.98 g cm${-3}$
957   orders of multipole-multipole interactions derived in the first paper
958   in this series and provides the most comprehensive test of the new
959   methods.  A liquid-phase system was created with 2000 water molecules
960 < and 48 dissolved ions at a density of 0.98 g cm${-3}$ and a
960 > and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
961   temperature of 300K.  After equilibration, this liquid-phase system
962   was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
963 < a cutoff radius of 9\AA.  The value of the damping coefficient was
963 > a cutoff radius of 12\AA.  The value of the damping coefficient was
964   also varied from the undamped case ($\alpha = 0$) to a heavily damped
965 < case ($\alpha = 0.3$ \AA$^{-1}$) for the real space methods.  A sample
966 < was also run using the multipolar Ewald sum.
965 > case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
966 > sample was also run using the multipolar Ewald sum with the same
967 > real-space cutoff.
968  
969   In figure~\ref{fig:energyDrift} we show the both the linear drift in
970   energy over time, $\delta E_1$, and the standard deviation of energy
971   fluctuations around this drift $\delta E_0$.  Both of the
972   shifted-force methods (GSF and TSF) provide excellent energy
973 < conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
973 > conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
974   while the hard cutoff is essentially unusable for molecular dynamics.
975   SP provides some benefit over the hard cutoff because the energetic
976   jumps that happen as particles leave and enter the cutoff sphere are
977 < somewhat reduced.
977 > somewhat reduced, but like the Wolf method for charges, the SP method
978 > would not be as useful for molecular dynamics as either of the
979 > shifted-force methods.
980  
981   We note that for all tested values of the cutoff radius, the new
982   real-space methods can provide better energy conservation behavior
# Line 873 | Line 985 | $k$-space cutoff values.
985  
986   \begin{figure}
987    \centering
988 <  \includegraphics[width=\textwidth]{newDrift.pdf}
988 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
989   \label{fig:energyDrift}        
990   \caption{Analysis of the energy conservation of the real-space
991    electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
992 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
993 <  of energy fluctuations around this drift.  All simulations were of a
994 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
995 <  K starting from the same initial configuration.}
992 >  energy over time (in kcal / mol / particle / ns) and $\delta
993 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
994 >  around this drift (in kcal / mol / particle).  All simulations were
995 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
996 >  300 K starting from the same initial configuration. All runs
997 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
998   \end{figure}
999  
1000 + \subsection{Reproduction of Structural Features\label{sec:structure}}
1001 + One of the best tests of modified interaction potentials is the
1002 + fidelity with which they can reproduce structural features in a
1003 + liquid.  One commonly-utilized measure of structural ordering is the
1004 + pair distribution function, $g(r)$, which measures local density
1005 + deviations in relation to the bulk density.  In the electrostatic
1006 + approaches studied here, the short-range repulsion from the
1007 + Lennard-Jones potential is identical for the various electrostatic
1008 + methods, and since short range repulsion determines much of the local
1009 + liquid ordering, one would not expect to see any differences in
1010 + $g(r)$.  Indeed, the pair distributions are essentially identical for
1011 + all of the electrostatic methods studied (for each of the different
1012 + systems under investigation).  Interested readers may consult the
1013 + supplementary information for plots of these pair distribution
1014 + functions.
1015  
1016 + A direct measure of the structural features that is a more
1017 + enlightening test of the modified electrostatic methods is the average
1018 + value of the electrostatic energy $\langle U_\mathrm{elect} \rangle$
1019 + which is obtained by sampling the liquid-state configurations
1020 + experienced by a liquid evolving entirely under the influence of the
1021 + methods being investigated.  In figure \ref{fig:Uelect} we show how
1022 + $\langle U_\mathrm{elect} \rangle$ for varies with the damping parameter,
1023 + $\alpha$, for each of the methods.
1024 +
1025 + \begin{figure}
1026 +  \centering
1027 +  \includegraphics[width=\textwidth]{averagePotentialEnergy_r9_12.eps}
1028 + \label{fig:Uelect}        
1029 + \caption{The average electrostatic potential energy,
1030 +  $\langle U_\mathrm{elect} \rangle$ for the SSDQ water with ions as a function
1031 +  of the damping parameter, $\alpha$, for each of the real-space
1032 +  electrostatic methods. Top panel: simulations run with a real-space
1033 +  cutoff, $r_c = 9$\AA.  Bottom panel: the same quantity, but with a
1034 +  larger cutoff, $r_c = 12$\AA.}
1035 + \end{figure}
1036 +
1037 + It is clear that moderate damping is important for converging the mean
1038 + potential energy values, particularly for the two shifted force
1039 + methods (GSF and TSF).  A value of $\alpha \approx 0.18$ \AA$^{-1}$ is
1040 + sufficient to converge the SP and GSF energies with a cutoff of 12
1041 + \AA, while shorter cutoffs require more dramatic damping ($\alpha
1042 + \approx 0.36$ \AA$^{-1}$ for $r_c = 9$ \AA).  It is also clear from
1043 + fig. \ref{fig:Uelect} that it is possible to overdamp the real-space
1044 + electrostatic methods, causing the estimate of the energy to drop
1045 + below the Ewald results.
1046 +
1047 + These ``optimal'' values of the damping coefficient are slightly
1048 + larger than what were observed for DSF electrostatics for purely
1049 + point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1050 + $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1051 + multipole systems.
1052 +
1053 + \subsection{Reproduction of Dynamic Properties\label{sec:structure}}
1054 + To test the fidelity of the electrostatic methods at reproducing
1055 + dynamics in a multipolar liquid, it is also useful to look at
1056 + transport properties, particularly the diffusion constant,
1057 + \begin{equation}
1058 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1059 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1060 + \label{eq:diff}
1061 + \end{equation}
1062 + which measures long-time behavior and is sensitive to the forces on
1063 + the multipoles.  For the soft dipolar fluid, and the SSDQ liquid
1064 + systems, the self-diffusion constants (D) were calculated from linear
1065 + fits to the long-time portion of the mean square displacement
1066 + ($\langle r^{2}(t) \rangle$).\cite{Allen87}
1067 +
1068 + In addition to translational diffusion, orientational relaxation times
1069 + were calculated for comparisons with the Ewald simulations and with
1070 + experiments. These values were determined from the same 1~ns $NVE$
1071 + trajectories used for translational diffusion by calculating the
1072 + orientational time correlation function,
1073 + \begin{equation}
1074 + C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\gamma(t)
1075 +                \cdot\hat{\mathbf{u}}_i^\gamma(0)\right]\right\rangle,
1076 + \label{eq:OrientCorr}
1077 + \end{equation}
1078 + where $P_l$ is the Legendre polynomial of order $l$ and
1079 + $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
1080 + axis $\gamma$.  The body-fixed reference frame used for our
1081 + orientational correlation functions has the $z$-axis running along the
1082 + dipoles, and for the SSDQ water model, the $y$-axis connects the two
1083 + implied hydrogen atoms.
1084 +
1085 + From the orientation autocorrelation functions, we can obtain time
1086 + constants for rotational relaxation either by fitting an exponential
1087 + function or by integrating the entire correlation function.  These
1088 + decay times are directly comparable to water orientational relaxation
1089 + times from nuclear magnetic resonance (NMR). The relaxation constant
1090 + obtained from $C_2^y(t)$ is normally of experimental interest because
1091 + it describes the relaxation of the principle axis connecting the
1092 + hydrogen atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular
1093 + portion of the dipole-dipole relaxation from a proton NMR signal and
1094 + should provide an estimate of the NMR relaxation time
1095 + constant.\cite{Impey82}
1096 +
1097 + Results for the diffusion constants and orientational relaxation times
1098 + are shown in figure \ref{fig:dynamics}. From this data, it is apparent
1099 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1100 + reproduced with high fidelity by the GSF method.
1101 +
1102 + The $\tau_2$ results in \ref{fig:dynamics} show a much greater
1103 + difference between the real-space and the Ewald results.
1104 +
1105 +
1106   \section{CONCLUSION}
1107 < We have generalized the charged neutralized potential energy
1108 < originally developed by the Wolf et al.\cite{Wolf:1999dn} for the
1109 < charge-charge interaction to the charge-multipole and
1110 < multipole-multipole interaction in the SP method for higher order
1111 < multipoles. Also, we have developed GSF and TSF methods by
1112 < implementing the modification purposed by Fennel and
1113 < Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the
1114 < higher order multipoles to ensure consistency and smooth truncation of
1115 < the electrostatic energy, force, and torque for the spherical
897 < truncation. The SP methods for multipoles proved its suitability in MC
898 < simulations. On the other hand, the results from the GSF method
899 < produced good agreement with the Ewald's energy, force, and
900 < torque. Also, it shows very good energy conservation in MD
901 < simulations.  The direct truncation of any molecular system without
902 < multipole neutralization creates the fluctuation in the electrostatic
903 < energy. This fluctuation in the energy is very large for the case of
904 < crystal because of long range of multipole ordering (Refer paper
905 < I).\cite{PaperI} This is also significant in the case of the liquid
906 < because of the local multipole ordering in the molecules. If the net
907 < multipole within cutoff radius neutralized within cutoff sphere by
908 < placing image multiples on the surface of the sphere, this fluctuation
909 < in the energy reduced significantly. Also, the multipole
910 < neutralization in the generalized SP method showed very good agreement
911 < with the Ewald as compared to direct truncation for the evaluation of
912 < the $\triangle E$ between the configurations.  In MD simulations, the
913 < energy conservation is very important. The conservation of the total
914 < energy can be ensured by i) enforcing the smooth truncation of the
915 < energy, force and torque in the cutoff radius and ii) making the
916 < energy, force and torque consistent with each other. The GSF and TSF
917 < methods ensure the consistency and smooth truncation of the energy,
918 < force and torque at the cutoff radius, as a result show very good
919 < total energy conservation. But the TSF method does not show good
920 < agreement in the absolute value of the electrostatic energy, force and
921 < torque with the Ewald.  The GSF method has mimicked Ewald’s force,
922 < energy and torque accurately and also conserved energy. Therefore, the
923 < GSF method is the suitable method for evaluating required force field
924 < in MD simulations. In addition, the energy drift and fluctuation from
925 < the GSF method is much better than Ewald’s method for finite-sized
926 < reciprocal space.
1107 > In the first paper in this series, we generalized the
1108 > charge-neutralized electrostatic energy originally developed by Wolf
1109 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1110 > up to quadrupolar order.  The SP method is essentially a
1111 > multipole-capable version of the Wolf model.  The SP method for
1112 > multipoles provides excellent agreement with Ewald-derived energies,
1113 > forces and torques, and is suitable for Monte Carlo simulations,
1114 > although the forces and torques retain discontinuities at the cutoff
1115 > distance that prevents its use in molecular dynamics.
1116  
1117 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1118 < that can be made extremely efficient using spline interpolations of
1119 < the radial functions.  They require no Fourier transforms or $k$-space
1120 < sums, and guarantee the smooth handling of energies, forces, and
1121 < torques as multipoles cross the real-space cutoff boundary.  
1117 > We also developed two natural extensions of the damped shifted-force
1118 > (DSF) model originally proposed by Fennel and
1119 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1120 > smooth truncation of energies, forces, and torques at the real-space
1121 > cutoff, and both converge to DSF electrostatics for point-charge
1122 > interactions.  The TSF model is based on a high-order truncated Taylor
1123 > expansion which can be relatively perturbative inside the cutoff
1124 > sphere.  The GSF model takes the gradient from an images of the
1125 > interacting multipole that has been projected onto the cutoff sphere
1126 > to derive shifted force and torque expressions, and is a significantly
1127 > more gentle approach.
1128  
1129 + Of the two newly-developed shifted force models, the GSF method
1130 + produced quantitative agreement with Ewald energy, force, and torques.
1131 + It also performs well in conserving energy in MD simulations.  The
1132 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1133 + place on a potential energy surface that is significantly perturbed
1134 + from Ewald-based electrostatics.  
1135 +
1136 + % The direct truncation of any electrostatic potential energy without
1137 + % multipole neutralization creates large fluctuations in molecular
1138 + % simulations.  This fluctuation in the energy is very large for the case
1139 + % of crystal because of long range of multipole ordering (Refer paper
1140 + % I).\cite{PaperI} This is also significant in the case of the liquid
1141 + % because of the local multipole ordering in the molecules. If the net
1142 + % multipole within cutoff radius neutralized within cutoff sphere by
1143 + % placing image multiples on the surface of the sphere, this fluctuation
1144 + % in the energy reduced significantly. Also, the multipole
1145 + % neutralization in the generalized SP method showed very good agreement
1146 + % with the Ewald as compared to direct truncation for the evaluation of
1147 + % the $\triangle E$ between the configurations.  In MD simulations, the
1148 + % energy conservation is very important. The conservation of the total
1149 + % energy can be ensured by i) enforcing the smooth truncation of the
1150 + % energy, force and torque in the cutoff radius and ii) making the
1151 + % energy, force and torque consistent with each other. The GSF and TSF
1152 + % methods ensure the consistency and smooth truncation of the energy,
1153 + % force and torque at the cutoff radius, as a result show very good
1154 + % total energy conservation. But the TSF method does not show good
1155 + % agreement in the absolute value of the electrostatic energy, force and
1156 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1157 + % energy and torque accurately and also conserved energy.
1158 +
1159 + The only cases we have found where the new GSF and SP real-space
1160 + methods can be problematic are those which retain a bulk dipole moment
1161 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1162 + materials, uniform weighting of the orientational contributions can be
1163 + important for converging the total energy.  In these cases, the
1164 + damping function which causes the non-uniform weighting can be
1165 + replaced by the bare electrostatic kernel, and the energies return to
1166 + the expected converged values.
1167 +
1168 + Based on the results of this work, the GSF method is a suitable and
1169 + efficient replacement for the Ewald sum for evaluating electrostatic
1170 + interactions in MD simulations.  Both methods retain excellent
1171 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1172 + energy drift and fluctuations from the GSF electrostatics are better
1173 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1174 + Because they use real-space cutoffs with moderate cutoff radii, the
1175 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1176 + increases.  Additionally, they can be made extremely efficient using
1177 + spline interpolations of the radial functions.  They require no
1178 + Fourier transforms or $k$-space sums, and guarantee the smooth
1179 + handling of energies, forces, and torques as multipoles cross the
1180 + real-space cutoff boundary.
1181 +
1182 + \begin{acknowledgments}
1183 +  JDG acknowledges helpful discussions with Christopher
1184 +  Fennell. Support for this project was provided by the National
1185 +  Science Foundation under grant CHE-1362211. Computational time was
1186 +  provided by the Center for Research Computing (CRC) at the
1187 +  University of Notre Dame.
1188 + \end{acknowledgments}
1189 +
1190   %\bibliographystyle{aip}
1191   \newpage
1192   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines