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 4206 by gezelter, Thu Aug 7 20:53:27 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 295 | Line 325 | U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1
325   expressed as the product of two multipole operators and a Coulombic
326   kernel,
327   \begin{equation}
328 < U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
328 > U_{ab}(r)= M_{a} M_{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}}
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.
330 > where the multipole operator for site $a$, $M_{a}$, is
331 > expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
332 > $a$, etc.
333  
334 < Interactions between multipoles can be expressed as higher derivatives
335 < of the bare Coulomb potential, so one way of ensuring that the forces
336 < and torques vanish at the cutoff distance is to include a larger
337 < number of terms in the truncated Taylor expansion, e.g.,
338 < %
339 < \begin{equation}
340 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
341 < \end{equation}
342 < %
343 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344 < Thus, for $f(r)=1/r$, we find
345 < %
346 < \begin{equation}
347 < f_1(r)=\frac{1}{r}- \frac{1}{R_c} + (r - R_c) \frac{1}{R_c^2} - \frac{(r-R_c)^2}{R_c^3} .
348 < \end{equation}
349 < This function is an approximate electrostatic potential that has
350 < vanishing second derivatives at the cutoff radius, making it suitable
351 < for shifting the forces and torques of charge-dipole interactions.
334 > % Interactions between multipoles can be expressed as higher derivatives
335 > % of the bare Coulomb potential, so one way of ensuring that the forces
336 > % and torques vanish at the cutoff distance is to include a larger
337 > % number of terms in the truncated Taylor expansion, e.g.,
338 > % %
339 > % \begin{equation}
340 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
341 > % \end{equation}
342 > % %
343 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344 > % Thus, for $f(r)=1/r$, we find
345 > % %
346 > % \begin{equation}
347 > % f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
348 > % \end{equation}
349 > % This function is an approximate electrostatic potential that has
350 > % vanishing second derivatives at the cutoff radius, making it suitable
351 > % for shifting the forces and torques of charge-dipole interactions.
352  
353 < In general, the TSF potential for any multipole-multipole interaction
354 < can be written
353 > The TSF potential for any multipole-multipole interaction can be
354 > written
355   \begin{equation}
356   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
357   \label{generic}
358   \end{equation}
359 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
360 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
361 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
362 < energy, force, and torques, the required number of terms from Taylor
363 < series expansion in $f_n(r)$ must be performed for different
364 < multipole-multipole interactions.
359 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
360 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
361 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
362 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
363 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
364 > force, and torques, a Taylor expansion with $n$ terms must be
365 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
366  
367 < To carry out the same procedure for a damped electrostatic kernel, we
368 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369 < Many of the derivatives of the damped kernel are well known from
370 < Smith's early work on multipoles for the Ewald
371 < summation.\cite{Smith82,Smith98}
367 > % To carry out the same procedure for a damped electrostatic kernel, we
368 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369 > % Many of the derivatives of the damped kernel are well known from
370 > % Smith's early work on multipoles for the Ewald
371 > % summation.\cite{Smith82,Smith98}
372  
373 < Note that increasing the value of $n$ will add additional terms to the
374 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
375 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
376 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
378 < for computing multipole energies, forces, and torques, and smooth
379 < cutoffs of these quantities can be guaranteed as long as the number of
380 < terms in the Taylor series exceeds the derivative order required.
373 > % Note that increasing the value of $n$ will add additional terms to the
374 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
375 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
376 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
378 > % for computing multipole energies, forces, and torques, and smooth
379 > % cutoffs of these quantities can be guaranteed as long as the number of
380 > % terms in the Taylor series exceeds the derivative order required.
381  
382   For multipole-multipole interactions, following this procedure results
383 < in separate radial functions for each distinct orientational
384 < contribution to the potential, and ensures that the forces and torques
385 < from {\it each} of these contributions will vanish at the cutoff
386 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
387 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
383 > in separate radial functions for each of the distinct orientational
384 > contributions to the potential, and ensures that the forces and
385 > torques from each of these contributions will vanish at the cutoff
386 > radius.  For example, the direct dipole dot product
387 > ($\mathbf{D}_{a}
388 > \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
389   dot products:
390   \begin{equation}
391 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
392 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
393 < -\frac{1}{4\pi \epsilon_0}
394 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
395 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
391 > U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
392 >  \mathbf{D}_{a} \cdot
393 > \mathbf{D}_{b} \right) v_{21}(r) +
394 > \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
395 > \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
396   \end{equation}
397  
398 < The electrostatic forces and torques acting on the central multipole
399 < site due to another site within cutoff sphere are derived from
398 > For the Taylor shifted (TSF) method with the undamped kernel,
399 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
400 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
401 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
402 > connection to unmodified electrostatics as well as the smooth
403 > transition to zero in both these functions as $r\rightarrow r_c$.  The
404 > electrostatic forces and torques acting on the central multipole due
405 > to another site within the cutoff sphere are derived from
406   Eq.~\ref{generic}, accounting for the appropriate number of
407   derivatives. Complete energy, force, and torque expressions are
408   presented in the first paper in this series (Reference
409 < \citep{PaperI}).
409 > \onlinecite{PaperI}).
410  
411   \subsection{Gradient-shifted force (GSF)}
412  
413 < A second (and significantly simpler) method involves shifting the
414 < gradient of the raw coulomb potential for each particular multipole
413 > A second (and conceptually simpler) method involves shifting the
414 > gradient of the raw Coulomb potential for each particular multipole
415   order.  For example, the raw dipole-dipole potential energy may be
416   shifted smoothly by finding the gradient for two interacting dipoles
417   which have been projected onto the surface of the cutoff sphere
418   without changing their relative orientation,
419 < \begin{displaymath}
420 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
421 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
422 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
423 < \end{displaymath}
424 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
425 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
426 < the signs are reversed for the dipole that has been projected onto the
427 < cutoff sphere).  In many ways, this simpler approach is closer in
428 < spirit to the original shifted force method, in that it projects a
429 < neutralizing multipole (and the resulting forces from this multipole)
430 < onto a cutoff sphere. The resulting functional forms for the
431 < potentials, forces, and torques turn out to be quite similar in form
432 < to the Taylor-shifted approach, although the radial contributions are
433 < significantly less perturbed by the Gradient-shifted approach than
434 < they are in the Taylor-shifted method.
419 > \begin{equation}
420 > U_{D_{a}D_{b}}(r)  = U_{D_{a}D_{b}}(r) -
421 > U_{D_{a}D_{b}}(r_c)
422 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
423 >  \nabla U_{D_{a}D_{b}}(r_c).
424 > \end{equation}
425 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
426 > (although the signs are reversed for the dipole that has been
427 > projected onto the cutoff sphere).  In many ways, this simpler
428 > approach is closer in spirit to the original shifted force method, in
429 > that it projects a neutralizing multipole (and the resulting forces
430 > from this multipole) onto a cutoff sphere. The resulting functional
431 > forms for the potentials, forces, and torques turn out to be quite
432 > similar in form to the Taylor-shifted approach, although the radial
433 > contributions are significantly less perturbed by the gradient-shifted
434 > approach than they are in the Taylor-shifted method.
435  
436 + For the gradient shifted (GSF) method with the undamped kernel,
437 + $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
438 + $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
439 + Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
440 + because the Taylor expansion retains only one term, they are
441 + significantly less perturbed than the TSF functions.
442 +
443   In general, the gradient shifted potential between a central multipole
444   and any multipolar site inside the cutoff radius is given by,
445   \begin{equation}
446 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
447 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
448 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
446 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
447 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
448 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
449   \label{generic2}
450   \end{equation}
451   where the sum describes a separate force-shifting that is applied to
452 < each orientational contribution to the energy.
452 > each orientational contribution to the energy.  In this expression,
453 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
454 > ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
455 > represent the orientations the multipoles.
456  
457   The third term converges more rapidly than the first two terms as a
458   function of radius, hence the contribution of the third term is very
459   small for large cutoff radii.  The force and torque derived from
460 < equation \ref{generic2} are consistent with the energy expression and
461 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
460 > Eq. \ref{generic2} are consistent with the energy expression and
461 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
462   can be considered generalizations of the original DSF method for
463   higher order multipole interactions. GSF and TSF are also identical up
464   to the charge-dipole interaction but generate different expressions in
465   the energy, force and torque for higher order multipole-multipole
466   interactions. Complete energy, force, and torque expressions for the
467   GSF potential are presented in the first paper in this series
468 < (Reference \citep{PaperI})
468 > (Reference~\onlinecite{PaperI}).
469  
470  
471   \subsection{Shifted potential (SP) }
# Line 439 | Line 478 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
478   interactions with the central multipole and the image. This
479   effectively shifts the total potential to zero at the cutoff radius,
480   \begin{equation}
481 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
481 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
482 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
483   \label{eq:SP}
484   \end{equation}          
485   where the sum describes separate potential shifting that is done for
486   each orientational contribution to the energy (e.g. the direct dipole
487   product contribution is shifted {\it separately} from the
488   dipole-distance terms in dipole-dipole interactions).  Note that this
489 < is not a simple shifting of the total potential at $R_c$. Each radial
489 > is not a simple shifting of the total potential at $r_c$. Each radial
490   contribution is shifted separately.  One consequence of this is that
491   multipoles that reorient after leaving the cutoff sphere can re-enter
492   the cutoff sphere without perturbing the total energy.
493  
494 < The potential energy between a central multipole and other multipolar
495 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
496 < force and torque obtained from the shifted potential (SP) are
497 < discontinuous at $R_c$. Therefore, MD simulations will still
498 < experience energy drift while operating under the SP potential, but it
499 < may be suitable for Monte Carlo approaches where the configurational
500 < energy differences are the primary quantity of interest.
494 > For the shifted potential (SP) method with the undamped kernel,
495 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
496 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
497 > central multipole and other multipolar sites goes smoothly to zero as
498 > $r \rightarrow r_c$.  However, the force and torque obtained from the
499 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
500 > will still experience energy drift while operating under the SP
501 > potential, but it may be suitable for Monte Carlo approaches where the
502 > configurational energy differences are the primary quantity of
503 > interest.
504  
505 < \subsection{The Self term}
505 > \subsection{The Self Term}
506   In the TSF, GSF, and SP methods, a self-interaction is retained for
507   the central multipole interacting with its own image on the surface of
508   the cutoff sphere.  This self interaction is nearly identical with the
509   self-terms that arise in the Ewald sum for multipoles.  Complete
510   expressions for the self terms are presented in the first paper in
511 < this series (Reference \citep{PaperI})  
511 > this series (Reference \onlinecite{PaperI}).
512  
513  
514   \section{\label{sec:methodology}Methodology}
# Line 480 | Line 523 | disordered and ordered condensed-phase systems.  These
523   arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
524   used the multipolar Ewald sum as a reference method for comparing
525   energies, forces, and torques for molecular models that mimic
526 < disordered and ordered condensed-phase systems.  These test-cases
527 < 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.
526 > disordered and ordered condensed-phase systems.  The parameters used
527 > in the test cases are given in table~\ref{tab:pars}.
528  
529 < The last test case exercises all levels of the multipole-multipole
530 < interactions we have derived so far and represents the most complete
531 < test of the new methods.  In the following section, we present results
532 < for the total electrostatic energy, as well as the electrostatic
533 < contributions to the force and torque on each molecule.  These
534 < quantities have been computed using the SP, TSF, and GSF methods, as
535 < well as a hard cutoff, and have been compared with the values obtaine
536 < from the multipolar Ewald sum.  In Mote Carlo (MC) simulations, the
537 < energy differences between two configurations is the primary quantity
538 < that governs how the simulation proceeds. These differences are the
539 < most imporant indicators of the reliability of a method even if the
540 < absolute energies are not exact.  For each of the multipolar systems
541 < listed above, we have compared the change in electrostatic potential
542 < energy ($\Delta E$) between 250 statistically-independent
543 < configurations.  In molecular dynamics (MD) simulations, the forces
544 < and torques govern the behavior of the simulation, so we also compute
545 < the electrostatic contributions to the forces and torques.
529 > \begin{table}
530 > \label{tab:pars}
531 > \caption{The parameters used in the systems used to evaluate the new
532 >  real-space methods.  The most comprehensive test was a liquid
533 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
534 >  ions).  This test excercises all orders of the multipolar
535 >  interactions developed in the first paper.}
536 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
537 >             & \multicolumn{2}{c|}{LJ parameters} &
538 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
539 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
540 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
541 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
542 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
543 > \AA\textsuperscript{2})} \\ \hline
544 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
545 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
546 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
547 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
548 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
549 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
550 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
551 > \end{tabularx}
552 > \end{table}
553 > The systems consist of pure multipolar solids (both dipole and
554 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
555 > fluid composed of sites containing both dipoles and quadrupoles
556 > simultaneously, and a final test case that includes ions with point
557 > charges in addition to the multipolar fluid.  The solid-phase
558 > parameters were chosen so that the systems can explore some
559 > orientational freedom for the multipolar sites, while maintaining
560 > relatively strict translational order.  The SSDQ model used here is
561 > not a particularly accurate water model, but it does test
562 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
563 > interactions at roughly the same magnitudes. The last test case, SSDQ
564 > water with dissolved ions, exercises \textit{all} levels of the
565 > multipole-multipole interactions we have derived so far and represents
566 > the most complete test of the new methods.
567  
568 < \subsection{Model systems}
569 < To sample independent configurations of multipolar crystals, a body
570 < centered cubic (bcc) crystal which is a minimum energy structure for
571 < point dipoles was generated using 3,456 molecules.  The multipoles
572 < were translationally locked in their respective crystal sites for
573 < equilibration at a relatively low temperature (50K), so that dipoles
574 < or quadrupoles could freely explore all accessible orientations.  The
575 < translational constraints were removed, and the crystals were
576 < simulated for 10 ps in the microcanonical (NVE) ensemble with an
577 < average temperature of 50 K.  Configurations were sampled at equal
578 < time intervals for the comparison of the configurational energy
579 < differences.  The crystals were not simulated close to the melting
580 < points in order to avoid translational deformation away of the ideal
581 < lattice geometry.
568 > In the following section, we present results for the total
569 > electrostatic energy, as well as the electrostatic contributions to
570 > the force and torque on each molecule.  These quantities have been
571 > computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
572 > and have been compared with the values obtained from the multipolar
573 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
574 > between two configurations is the primary quantity that governs how
575 > the simulation proceeds. These differences are the most important
576 > indicators of the reliability of a method even if the absolute
577 > energies are not exact.  For each of the multipolar systems listed
578 > above, we have compared the change in electrostatic potential energy
579 > ($\Delta E$) between 250 statistically-independent configurations.  In
580 > molecular dynamics (MD) simulations, the forces and torques govern the
581 > behavior of the simulation, so we also compute the electrostatic
582 > contributions to the forces and torques.
583  
584 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
585 < system was created with 2048 molecules oriented randomly.  These were
584 > \subsection{Implementation}
585 > The real-space methods developed in the first paper in this series
586 > have been implemented in our group's open source molecular simulation
587 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
588 > this work.  The complementary error function can be a relatively slow
589 > function on some processors, so all of the radial functions are
590 > precomputed on a fine grid and are spline-interpolated to provide
591 > values when required.  
592  
593 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
594 < temperature after equilibration.  We collected 250 different
595 < configurations in equal interval of time. For the ions mixed liquid
596 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
597 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
598 < at the same environment for 1ns and 250 configurations were
599 < collected. While comparing energies, forces, and torques with Ewald
600 < method, Lennard-Jones potentials were turned off and purely
543 < electrostatic interaction had been compared.
593 > Using the same simulation code, we compare to a multipolar Ewald sum
594 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
595 > the Ewald sum is a re-implementation of the algorithm originally
596 > proposed by Smith that does not use the particle mesh or smoothing
597 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
598 > being compared are the electrostatic contributions to energies, force,
599 > and torques.  All other contributions to these quantities (i.e. from
600 > Lennard-Jones interactions) are removed prior to the comparisons.
601  
602 + The convergence parameter ($\alpha$) also plays a role in the balance
603 + of the real-space and reciprocal-space portions of the Ewald
604 + calculation.  Typical molecular mechanics packages set this to a value
605 + that depends on the cutoff radius and a tolerance (typically less than
606 + $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
607 + associated with increasing accuracy at the expense of computational
608 + time spent on the reciprocal-space portion of the
609 + summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
610 + 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
611 + Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
612 +
613 + The real-space models have self-interactions that provide
614 + contributions to the energies only.  Although the self interaction is
615 + a rapid calculation, we note that in systems with fluctuating charges
616 + or point polarizabilities, the self-term is not static and must be
617 + recomputed at each time step.
618 +
619 + \subsection{Model systems}
620 + To sample independent configurations of the multipolar crystals, body
621 + centered cubic (bcc) crystals, which exhibit the minimum energy
622 + structures for point dipoles, were generated using 3,456 molecules.
623 + The multipoles were translationally locked in their respective crystal
624 + sites for equilibration at a relatively low temperature (50K) so that
625 + dipoles or quadrupoles could freely explore all accessible
626 + orientations.  The translational constraints were then removed, the
627 + systems were re-equilibrated, and the crystals were simulated for an
628 + additional 10 ps in the microcanonical (NVE) ensemble with an average
629 + temperature of 50 K.  The balance between moments of inertia and
630 + particle mass were chosen to allow orientational sampling without
631 + significant translational motion.  Configurations were sampled at
632 + equal time intervals in order to compare configurational energy
633 + differences.  The crystals were simulated far from the melting point
634 + in order to avoid translational deformation away of the ideal lattice
635 + geometry.
636 +
637 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
638 + simulations, each system was created with 2,048 randomly-oriented
639 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
640 + Each system was then simulated for 1 ns in the microcanonical (NVE)
641 + ensemble.  We collected 250 different configurations at equal time
642 + intervals. For the liquid system that included ionic species, we
643 + converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
644 + \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
645 + run under the same conditions for 1 ns. A total of 250 configurations
646 + were collected. In the following comparisons of energies, forces, and
647 + torques, the Lennard-Jones potentials were turned off and only the
648 + purely electrostatic quantities were compared with the same values
649 + obtained via the Ewald sum.
650 +
651   \subsection{Accuracy of Energy Differences, Forces and Torques}
652   The pairwise summation techniques (outlined above) were evaluated for
653   use in MC simulations by studying the energy differences between
# Line 554 | Line 660 | we used least square regressions analysiss for the six
660   should be identical for all methods.
661  
662   Since none of the real-space methods provide exact energy differences,
663 < we used least square regressions analysiss for the six different
663 > we used least square regressions analysis for the six different
664   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
665   with the multipolar Ewald reference method.  Unitary results for both
666   the correlation (slope) and correlation coefficient for these
# Line 565 | Line 671 | also been compared by using least squares regression a
671   configurations and 250 configurations were recorded for comparison.
672   Each system provided 31,125 energy differences for a total of 186,750
673   data points.  Similarly, the magnitudes of the forces and torques have
674 < also been compared by using least squares regression analyses. In the
674 > also been compared using least squares regression analysis. In the
675   forces and torques comparison, the magnitudes of the forces acting in
676   each molecule for each configuration were evaluated. For example, our
677   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 685 | Line 791 | model must allow for long simulation times with minima
791  
792   \begin{figure}
793    \centering
794 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
794 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
795    \caption{Statistical analysis of the quality of configurational
796      energy differences for the real-space electrostatic methods
797      compared with the reference Ewald sum.  Results with a value equal
# Line 699 | Line 805 | reproduce the Ewald-derived configurational energy dif
805  
806   The combined correlation coefficient and slope for all six systems is
807   shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
808 < reproduce the Ewald-derived configurational energy differences with
809 < remarkable fidelity.  Undamped hard cutoffs introduce a significant
810 < amount of random scatter in the energy differences which is apparent
811 < in the reduced value of the correlation coefficient for this method.
812 < This can be understood easily as configurations which exhibit only
813 < small traversals of a few dipoles or quadrupoles out of the cutoff
814 < sphere will see large energy jumps when hard cutoffs are used.  The
808 > reproduce the Ewald configurational energy differences with remarkable
809 > fidelity.  Undamped hard cutoffs introduce a significant amount of
810 > random scatter in the energy differences which is apparent in the
811 > reduced value of the correlation coefficient for this method.  This
812 > can be easily understood as configurations which exhibit small
813 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
814 > will see large energy jumps when hard cutoffs are used.  The
815   orientations of the multipoles (particularly in the ordered crystals)
816 < mean that these jumps can go either up or down in energy, producing a
817 < significant amount of random scatter.
816 > mean that these energy jumps can go in either direction, producing a
817 > significant amount of random scatter, but no systematic error.
818  
819   The TSF method produces energy differences that are highly correlated
820   with the Ewald results, but it also introduces a significant
# Line 716 | Line 822 | effect on crystalline systems.
822   smaller cutoff values. The TSF method alters the distance dependence
823   of different orientational contributions to the energy in a
824   non-uniform way, so the size of the cutoff sphere can have a large
825 < effect on crystalline systems.
825 > effect, particularly for the crystalline systems.
826  
827   Both the SP and GSF methods appear to reproduce the Ewald results with
828   excellent fidelity, particularly for moderate damping ($\alpha =
829 < 0.1-0.2$\AA$^{-1}$) and commonly-used cutoff values ($r_c = 12$\AA).
830 < With the exception of the undamped hard cutoff, and the TSF method
831 < with short cutoffs, all of the methods would be appropriate for use in
832 < Monte Carlo simulations.
829 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
830 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
831 > method with short cutoffs, all of the methods would be appropriate for
832 > use in Monte Carlo simulations.
833  
834   \subsection{Magnitude of the force and torque vectors}
835  
836 < The comparison of the magnitude of the combined forces and torques for
837 < the data accumulated from all system types are shown in Figures
836 > The comparisons of the magnitudes of the forces and torques for the
837 > data accumulated from all six systems are shown in Figures
838   ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
839   correlation and slope for the forces agree well with the Ewald sum
840 < even for the hard cutoff method.
840 > even for the hard cutoffs.
841  
842 < For the system of molecules with higher order multipoles, the
843 < interaction is quite short ranged. Moreover, the force decays more
844 < rapidly than the electrostatic energy hence the hard cutoff method can
845 < also produces reasonable agreement.  Although the pure cutoff gives
846 < the good match of the electrostatic force for pairs of molecules
847 < included within the cutoff sphere, the discontinuity in the force at
848 < the cutoff radius can potentially cause problems the total energy
849 < conservation as molecules enter and leave the cutoff sphere.  This is
850 < discussed in detail in section \ref{sec:}.
842 > For systems of molecules with only multipolar interactions, the pair
843 > energy contributions are quite short ranged.  Moreover, the force
844 > decays more rapidly than the electrostatic energy, hence the hard
845 > cutoff method can also produce reasonable agreement for this quantity.
846 > Although the pure cutoff gives reasonably good electrostatic forces
847 > for pairs of molecules included within each other's cutoff spheres,
848 > the discontinuity in the force at the cutoff radius can potentially
849 > cause energy conservation problems as molecules enter and leave the
850 > cutoff spheres.  This is discussed in detail in section
851 > \ref{sec:conservation}.
852  
853   The two shifted-force methods (GSF and TSF) exhibit a small amount of
854   systematic variation and scatter compared with the Ewald forces.  The
855   shifted-force models intentionally perturb the forces between pairs of
856 < molecules inside the cutoff sphere in order to correct the energy
857 < conservation issues, so it is not particularly surprising that this
858 < perturbation is evident in these same molecular forces.  The GSF
859 < perturbations are minimal, particularly for moderate damping and and
856 > molecules inside each other's cutoff spheres in order to correct the
857 > energy conservation issues, and this perturbation is evident in the
858 > statistics accumulated for the molecular forces.  The GSF
859 > perturbations are minimal, particularly for moderate damping and
860   commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
861   reasonable agreement in the correlation coefficient but again the
862   systematic error in the forces is concerning if replication of Ewald
# Line 757 | Line 864 | forces is desired.
864  
865   \begin{figure}
866    \centering
867 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
867 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
868    \caption{Statistical analysis of the quality of the force vector
869      magnitudes for the real-space electrostatic methods compared with
870      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 771 | Line 878 | forces is desired.
878  
879   \begin{figure}
880    \centering
881 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
881 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
882    \caption{Statistical analysis of the quality of the torque vector
883      magnitudes for the real-space electrostatic methods compared with
884      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 791 | Line 898 | The other real-space methods can cause some significan
898  
899   The results shows that the torque from the hard cutoff method
900   reproduces the torques in quite good agreement with the Ewald sum.
901 < The other real-space methods can cause some significant deviations,
902 < but excellent agreement with the Ewald sum torques is recovered at
903 < moderate values of the damping coefficient ($\alpha =
904 < 0.1-0.2$\AA$^{-1}$) and cutoff radius ($r_c \ge 12$\AA).  The TSF
905 < method exhibits the only fair agreement in the slope as compared to
906 < Ewald even for larger cutoff radii.  It appears that the severity of
907 < the perturbations in the TSF method are most apparent in the torques.
901 > The other real-space methods can cause some deviations, but excellent
902 > agreement with the Ewald sum torques is recovered at moderate values
903 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
904 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
905 > in the slope when compared with the Ewald torques even for larger
906 > cutoff radii.  It appears that the severity of the perturbations in
907 > the TSF method are most in evidence for the torques.
908  
909   \subsection{Directionality of the force and torque vectors}  
910  
# Line 806 | Line 913 | directionality is shown in terms of circular variance
913   these quantities. Force and torque vectors for all six systems were
914   analyzed using Fisher statistics, and the quality of the vector
915   directionality is shown in terms of circular variance
916 < ($\mathrm{Var}(\theta$) in figure
916 > ($\mathrm{Var}(\theta)$) in figure
917   \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
918 < from the new real-space method exhibit nearly-ideal Fisher probability
918 > from the new real-space methods exhibit nearly-ideal Fisher probability
919   distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
920   exhibit the best vectorial agreement with the Ewald sum. The force and
921   torque vectors from GSF method also show good agreement with the Ewald
922   method, which can also be systematically improved by using moderate
923 < damping and a reasonable cutoff radius.  For $\alpha = 0.2$ and $r_c =
923 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
924   12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
925 < to a distribution with 95\% of force vectors within $6.37^\circ$ of the
926 < corresponding Ewald forces. The TSF method produces the poorest
925 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
926 > the corresponding Ewald forces. The TSF method produces the poorest
927   agreement with the Ewald force directions.
928  
929 < Torques are again more perturbed by the new real-space methods, than
930 < forces, but even here the variance is reasonably small.  For the same
929 > Torques are again more perturbed than the forces by the new real-space
930 > methods, but even here the variance is reasonably small.  For the same
931   method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
932   the circular variance was 0.01415, corresponds to a distribution which
933   has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
# Line 829 | Line 936 | systematically improved by varying $\alpha$ and $r_c$.
936  
937   \begin{figure}
938    \centering
939 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
939 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
940    \caption{The circular variance of the direction of the force and
941      torque vectors obtained from the real-space methods around the
942      reference Ewald vectors. A variance equal to 0 (dashed line)
# Line 840 | Line 947 | systematically improved by varying $\alpha$ and $r_c$.
947    \label{fig:slopeCorr_circularVariance}
948   \end{figure}
949  
950 < \subsection{Energy conservation}
950 > \subsection{Energy conservation\label{sec:conservation}}
951  
952   We have tested the conservation of energy one can expect to see with
953   the new real-space methods using the SSDQ water model with a small
# Line 848 | Line 955 | and 48 dissolved ions at a density of 0.98 g cm${-3}$
955   orders of multipole-multipole interactions derived in the first paper
956   in this series and provides the most comprehensive test of the new
957   methods.  A liquid-phase system was created with 2000 water molecules
958 < and 48 dissolved ions at a density of 0.98 g cm${-3}$ and a
959 < temperature of 300K.  After equilibration, this liquid-phase system
960 < was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
961 < a cutoff radius of 9\AA.  The value of the damping coefficient was
962 < also varied from the undamped case ($\alpha = 0$) to a heavily damped
963 < case ($\alpha = 0.3$ \AA$^{-1}$) for the real space methods.  A sample
964 < was also run using the multipolar Ewald sum.
958 > and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
959 > temperature of 300K.  After equilibration in the canonical (NVT)
960 > ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
961 > was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
962 > Hard, SP, GSF, and TSF methods with a cutoff radius of 12\AA.  The
963 > value of the damping coefficient was also varied from the undamped
964 > case ($\alpha = 0$) to a heavily damped case ($\alpha = 0.3$
965 > \AA$^{-1}$) for all of the real space methods.  A sample was also run
966 > using the multipolar Ewald sum with the same real-space cutoff.
967  
968   In figure~\ref{fig:energyDrift} we show the both the linear drift in
969   energy over time, $\delta E_1$, and the standard deviation of energy
970   fluctuations around this drift $\delta E_0$.  Both of the
971   shifted-force methods (GSF and TSF) provide excellent energy
972 < conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
972 > conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
973   while the hard cutoff is essentially unusable for molecular dynamics.
974   SP provides some benefit over the hard cutoff because the energetic
975   jumps that happen as particles leave and enter the cutoff sphere are
976 < somewhat reduced.
976 > somewhat reduced, but like the Wolf method for charges, the SP method
977 > would not be as useful for molecular dynamics as either of the
978 > shifted-force methods.
979  
980   We note that for all tested values of the cutoff radius, the new
981   real-space methods can provide better energy conservation behavior
# Line 873 | Line 984 | $k$-space cutoff values.
984  
985   \begin{figure}
986    \centering
987 <  \includegraphics[width=\textwidth]{newDrift.pdf}
987 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
988   \label{fig:energyDrift}        
989   \caption{Analysis of the energy conservation of the real-space
990 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
991 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
992 <  of energy fluctuations around this drift.  All simulations were of a
993 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
994 <  K starting from the same initial configuration.}
990 >  methods. $\delta \mathrm{E}_1$ is the linear drift in energy over
991 >  time (in kcal / mol / particle / ns) and $\delta \mathrm{E}_0$ is
992 >  the standard deviation of energy fluctuations around this drift (in
993 >  kcal / mol / particle).  Points that appear below the dashed grey
994 >  (Ewald) lines exhibit better energy conservation than commonly-used
995 >  parameters for Ewald-based electrostatics.  All simulations were of
996 >  a 2000-molecule simulation of SSDQ water with 48 ionic charges at
997 >  300 K starting from the same initial configuration. All runs
998 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
999   \end{figure}
1000  
1001 + \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
1002 + The most important test of the modified interaction potentials is the
1003 + fidelity with which they can reproduce structural features and
1004 + dynamical properties in a liquid.  One commonly-utilized measure of
1005 + structural ordering is the pair distribution function, $g(r)$, which
1006 + measures local density deviations in relation to the bulk density.  In
1007 + the electrostatic approaches studied here, the short-range repulsion
1008 + from the Lennard-Jones potential is identical for the various
1009 + electrostatic methods, and since short range repulsion determines much
1010 + of the local liquid ordering, one would not expect to see many
1011 + differences in $g(r)$.  Indeed, the pair distributions are essentially
1012 + identical for all of the electrostatic methods studied (for each of
1013 + the different systems under investigation).  An example of this
1014 + agreement for the SSDQ water/ion system is shown in
1015 + Fig. \ref{fig:gofr}.
1016  
1017 + \begin{figure}
1018 +  \centering
1019 +  \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
1020 + \label{fig:gofr}        
1021 + \caption{The pair distribution functions, $g(r)$, for the SSDQ
1022 +  water/ion system obtained using the different real-space methods are
1023 +  essentially identical with the result from the Ewald
1024 +  treatment.}
1025 + \end{figure}
1026 +
1027 + There is a very slight overstructuring of the first solvation shell
1028 + when using when using TSF at lower values of the damping coefficient
1029 + ($\alpha \le 0.1$) or when using undamped GSF.  With moderate damping,
1030 + GSF and SP produce pair distributions that are identical (within
1031 + numerical noise) to their Ewald counterparts.
1032 +
1033 + A structural property that is a more demanding test of modified
1034 + electrostatics is the mean value of the electrostatic energy $\langle
1035 + U_\mathrm{elect} \rangle / N$ which is obtained by sampling the
1036 + liquid-state configurations experienced by a liquid evolving entirely
1037 + under the influence of each of the methods.  In table \ref{tab:Props}
1038 + we demonstrate how $\langle U_\mathrm{elect} \rangle / N$ varies with
1039 + the damping parameter, $\alpha$, for each of the methods.
1040 +
1041 + As in the crystals studied in the first paper, damping is important
1042 + for converging the mean electrostatic energy values, particularly for
1043 + the two shifted force methods (GSF and TSF).  A value of $\alpha
1044 + \approx 0.2$ \AA$^{-1}$ is sufficient to converge the SP and GSF
1045 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
1046 + dramatic damping ($\alpha \approx 0.3$ \AA$^{-1}$ for $r_c = 9$ \AA).
1047 + Overdamping the real-space electrostatic methods occurs with $\alpha >
1048 + 0.4$, causing the estimate of the energy to drop below the Ewald
1049 + results.
1050 +
1051 + These ``optimal'' values of the damping coefficient are slightly
1052 + larger than what were observed for DSF electrostatics for purely
1053 + point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1054 + $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1055 + multipole systems.
1056 +
1057 + To test the fidelity of the electrostatic methods at reproducing
1058 + dynamics in a multipolar liquid, it is also useful to look at
1059 + transport properties, particularly the diffusion constant,
1060 + \begin{equation}
1061 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1062 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1063 + \label{eq:diff}
1064 + \end{equation}
1065 + which measures long-time behavior and is sensitive to the forces on
1066 + the multipoles.  For the soft dipolar fluid and the SSDQ liquid
1067 + systems, the self-diffusion constants (D) were calculated from linear
1068 + fits to the long-time portion of the mean square displacement,
1069 + $\langle r^{2}(t) \rangle$.\cite{Allen87}
1070 +
1071 + In addition to translational diffusion, orientational relaxation times
1072 + were calculated for comparisons with the Ewald simulations and with
1073 + experiments. These values were determined from the same 1~ns
1074 + microcanonical trajectories used for translational diffusion by
1075 + calculating the orientational time correlation function,
1076 + \begin{equation}
1077 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1078 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1079 + \label{eq:OrientCorr}
1080 + \end{equation}
1081 + where $P_l$ is the Legendre polynomial of order $l$ and
1082 + $\hat{\mathbf{A}}_\gamma$ is the space-frame unit vector for body axis
1083 + $\gamma$ on a molecule..  Th body-fixed reference frame used for our
1084 + models has the $z$-axis running along the dipoles, and for the SSDQ
1085 + water model, the $y$-axis connects the two implied hydrogen atom
1086 + positions.  From the orientation autocorrelation functions, we can
1087 + obtain time constants for rotational relaxation either by fitting an
1088 + exponential function or by integrating the entire correlation
1089 + function.  In a good water model, these decay times would be
1090 + comparable to water orientational relaxation times from nuclear
1091 + magnetic resonance (NMR). The relaxation constant obtained from
1092 + $C_2^y(t)$ is normally of experimental interest because it describes
1093 + the relaxation of the principle axis connecting the hydrogen
1094 + atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1095 + of the dipole-dipole relaxation from a proton NMR signal and should
1096 + provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1097 +
1098 + Results for the diffusion constants and orientational relaxation times
1099 + are shown in figure \ref{tab:Props}. From this data, it is apparent
1100 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1101 + reproduced with reasonable fidelity by the GSF method.
1102 +
1103 + The $\tau_2$ results in \ref{tab:Props} show a much greater difference
1104 + between the real-space and the Ewald results.
1105 +
1106 + \begin{table}
1107 + \label{tab:Props}
1108 + \caption{Comparison of the structural and dynamic properties for the
1109 +  soft dipolar liquid test for all of the real-space methods.}
1110 + \begin{tabular}{l|c|cccc|cccc|cccc}
1111 +         & Ewald & \multicolumn{4}{c|}{SP} & \multicolumn{4}{c|}{GSF} & \multicolumn{4}{c|}{TSF} \\
1112 + $\alpha$ (\AA$^{-1}$) & &      
1113 + 0.0 & 0.1 & 0.2 & 0.3 &
1114 + 0.0 & 0.1 & 0.2 & 0.3 &
1115 + 0.0 & 0.1 & 0.2 & 0.3 \\ \cline{2-6}\cline{6-10}\cline{10-14}
1116 +
1117 + $\langle U_\mathrm{elect} \rangle /N$ &&&&&&&&&&&&&\\
1118 + D ($10^{-4}~\mathrm{cm}^2/\mathrm{s}$)&
1119 + 470.2(6) &
1120 + 416.6(5) &
1121 + 379.6(5) &
1122 + 438.6(5) &
1123 + 476.0(6) &
1124 + 412.8(5) &
1125 + 421.1(5) &
1126 + 400.5(5) &
1127 + 437.5(6) &
1128 + 434.6(5) &
1129 + 411.4(5) &
1130 + 545.3(7) &
1131 + 459.6(6) \\
1132 + $\tau_2$ (fs) &
1133 + 1.136 &
1134 + 1.041 &
1135 + 1.064 &
1136 + 1.109 &
1137 + 1.211 &
1138 + 1.119 &
1139 + 1.039 &
1140 + 1.058 &
1141 + 1.21  &
1142 + 1.15  &
1143 + 1.172 &
1144 + 1.153 &
1145 + 1.125 \\
1146 + \end{tabular}
1147 + \end{table}
1148 +
1149 +
1150   \section{CONCLUSION}
1151 < We have generalized the charged neutralized potential energy
1152 < originally developed by the Wolf et al.\cite{Wolf:1999dn} for the
1153 < charge-charge interaction to the charge-multipole and
1154 < multipole-multipole interaction in the SP method for higher order
1155 < multipoles. Also, we have developed GSF and TSF methods by
1156 < implementing the modification purposed by Fennel and
1157 < Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the
1158 < higher order multipoles to ensure consistency and smooth truncation of
1159 < 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.
1151 > In the first paper in this series, we generalized the
1152 > charge-neutralized electrostatic energy originally developed by Wolf
1153 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1154 > up to quadrupolar order.  The SP method is essentially a
1155 > multipole-capable version of the Wolf model.  The SP method for
1156 > multipoles provides excellent agreement with Ewald-derived energies,
1157 > forces and torques, and is suitable for Monte Carlo simulations,
1158 > although the forces and torques retain discontinuities at the cutoff
1159 > distance that prevents its use in molecular dynamics.
1160  
1161 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1162 < that can be made extremely efficient using spline interpolations of
1163 < the radial functions.  They require no Fourier transforms or $k$-space
1164 < sums, and guarantee the smooth handling of energies, forces, and
1165 < torques as multipoles cross the real-space cutoff boundary.  
1161 > We also developed two natural extensions of the damped shifted-force
1162 > (DSF) model originally proposed by Fennel and
1163 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1164 > smooth truncation of energies, forces, and torques at the real-space
1165 > cutoff, and both converge to DSF electrostatics for point-charge
1166 > interactions.  The TSF model is based on a high-order truncated Taylor
1167 > expansion which can be relatively perturbative inside the cutoff
1168 > sphere.  The GSF model takes the gradient from an images of the
1169 > interacting multipole that has been projected onto the cutoff sphere
1170 > to derive shifted force and torque expressions, and is a significantly
1171 > more gentle approach.
1172  
1173 + Of the two newly-developed shifted force models, the GSF method
1174 + produced quantitative agreement with Ewald energy, force, and torques.
1175 + It also performs well in conserving energy in MD simulations.  The
1176 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1177 + place on a potential energy surface that is significantly perturbed
1178 + from Ewald-based electrostatics.  
1179 +
1180 + % The direct truncation of any electrostatic potential energy without
1181 + % multipole neutralization creates large fluctuations in molecular
1182 + % simulations.  This fluctuation in the energy is very large for the case
1183 + % of crystal because of long range of multipole ordering (Refer paper
1184 + % I).\cite{PaperI} This is also significant in the case of the liquid
1185 + % because of the local multipole ordering in the molecules. If the net
1186 + % multipole within cutoff radius neutralized within cutoff sphere by
1187 + % placing image multiples on the surface of the sphere, this fluctuation
1188 + % in the energy reduced significantly. Also, the multipole
1189 + % neutralization in the generalized SP method showed very good agreement
1190 + % with the Ewald as compared to direct truncation for the evaluation of
1191 + % the $\triangle E$ between the configurations.  In MD simulations, the
1192 + % energy conservation is very important. The conservation of the total
1193 + % energy can be ensured by i) enforcing the smooth truncation of the
1194 + % energy, force and torque in the cutoff radius and ii) making the
1195 + % energy, force and torque consistent with each other. The GSF and TSF
1196 + % methods ensure the consistency and smooth truncation of the energy,
1197 + % force and torque at the cutoff radius, as a result show very good
1198 + % total energy conservation. But the TSF method does not show good
1199 + % agreement in the absolute value of the electrostatic energy, force and
1200 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1201 + % energy and torque accurately and also conserved energy.
1202 +
1203 + The only cases we have found where the new GSF and SP real-space
1204 + methods can be problematic are those which retain a bulk dipole moment
1205 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1206 + materials, uniform weighting of the orientational contributions can be
1207 + important for converging the total energy.  In these cases, the
1208 + damping function which causes the non-uniform weighting can be
1209 + replaced by the bare electrostatic kernel, and the energies return to
1210 + the expected converged values.
1211 +
1212 + Based on the results of this work, the GSF method is a suitable and
1213 + efficient replacement for the Ewald sum for evaluating electrostatic
1214 + interactions in MD simulations.  Both methods retain excellent
1215 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1216 + energy drift and fluctuations from the GSF electrostatics are better
1217 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1218 + Because they use real-space cutoffs with moderate cutoff radii, the
1219 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1220 + increases.  Additionally, they can be made extremely efficient using
1221 + spline interpolations of the radial functions.  They require no
1222 + Fourier transforms or $k$-space sums, and guarantee the smooth
1223 + handling of energies, forces, and torques as multipoles cross the
1224 + real-space cutoff boundary.
1225 +
1226 + \begin{acknowledgments}
1227 +  JDG acknowledges helpful discussions with Christopher
1228 +  Fennell. Support for this project was provided by the National
1229 +  Science Foundation under grant CHE-1362211. Computational time was
1230 +  provided by the Center for Research Computing (CRC) at the
1231 +  University of Notre Dame.
1232 + \end{acknowledgments}
1233 +
1234   %\bibliographystyle{aip}
1235   \newpage
1236   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines