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

Comparing trunk/multipole/multipole_2/multipole2.tex (file contents):
Revision 4171 by gezelter, Wed Jun 4 19:31:06 2014 UTC vs.
Revision 4186 by gezelter, Sun Jun 15 02:31:11 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
49 < Sum. II. Comparison of Simulation Methodologies} % Force line breaks with \\
50 > \title{Real space alternatives to the Ewald Sum. II. Comparison of Methods}
51  
52   \author{Madan Lamichhane}
53 < \affiliation{Department of Physics, University
53 < of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
53 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
54  
55   \author{Kathie E. Newman}
56 < \affiliation{Department of Physics, University
57 < of Notre Dame, Notre Dame, IN 46556}
56 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
57  
58   \author{J. Daniel Gezelter}%
59   \email{gezelter@nd.edu.}
60 < \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
61 < }%
60 > \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
61 > }
62  
63 < \date{\today}% It is always \today, today,
65 <             %  but any date may be explicitly specified
63 > \date{\today}
64  
65   \begin{abstract}
66 < 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.
66 >  We report on tests of the real-space shifted potential (SP),
67 >  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
68 >  for multipole interactions developed in the first paper in this
69 >  series, using the multipolar Ewald sum as a reference method. The
70 >  tests were carried out in a variety of condensed-phase environments
71 >  designed to test up to quadrupole-quadrupole interactions.
72 >  Comparisons of the energy differences between configurations,
73 >  molecular forces, and torques were used to analyze how well the
74 >  real-space models perform relative to the more computationally
75 >  expensive Ewald treatment.  We have also investigated the energy
76 >  conservation properties of the new methods in molecular dynamics
77 >  simulations. The SP method shows excellent agreement with
78 >  configurational energy differences, forces, and torques, and would
79 >  be suitable for use in Monte Carlo calculations.  Of the two new
80 >  shifted-force methods, the GSF approach shows the best agreement
81 >  with Ewald-derived energies, forces, and torques and also exhibits
82 >  energy conservation properties that make it an excellent choice for
83 >  efficient computation of electrostatic interactions in molecular
84 >  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. BETTER CITATIONS\cite{Clarke:1986eu,Woodcock75}
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
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
118 > method.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
119 > Parry's extension of the three dimensional Ewald sum is appropriate
120 > for slab geometries.\cite{Parry:1975if} Modified Ewald methods that
121 > were developed to handle two-dimensional (2D) electrostatic
122 > interactions in interfacial systems have not seen similar
123 > particle-mesh treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
124 >  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
125 > with system size. The inherent periodicity in the Ewald’s method can
126 > also be problematic for interfacial molecular
127 > systems.\cite{Fennell:2006lq}
128  
129   \subsection{Real-space methods}
130   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
131   method for calculating electrostatic interactions between point
132 < charges. They argued that the effective Coulomb interaction in
133 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
134 < an ordered lattice (e.g. when computing the Madelung constant of an
135 < ionic solid), the material can be considered as a set of ions
136 < interacting with neutral dipolar or quadrupolar ``molecules'' giving
137 < an effective distance dependence for the electrostatic interactions of
138 < $r^{-5}$ (see figure \ref{fig:NaCl}.  For this reason, careful
139 < applications of Wolf's method are able to obtain accurate estimates of
140 < Madelung constants using relatively short cutoff radii.  Recently,
141 < Fukuda used neutralization of the higher order moments for the
142 < calculation of the electrostatic interaction of the point charges
143 < system.\cite{Fukuda:2013sf}
132 > charges. They argued that the effective Coulomb interaction in most
133 > condensed phase systems is effectively short
134 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
135 > computing the Madelung constant of an ionic solid), the material can
136 > be considered as a set of ions interacting with neutral dipolar or
137 > quadrupolar ``molecules'' giving an effective distance dependence for
138 > the electrostatic interactions of $r^{-5}$ (see figure
139 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
140 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
141 > electrostatic energy per ion converges more rapidly to the Madelung
142 > energy than the dipolar approximation.\cite{Wolf92} To find the
143 > correct Madelung constant, Lacman suggested that the NaCl structure
144 > could be constructed in a way that the finite crystal terminates with
145 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
146 > what is effectively a set of octupoles at large distances. These facts
147 > suggest that the Madelung constants are relatively short ranged for
148 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
149 > application of Wolf's method can provide accurate estimates of
150 > Madelung constants using relatively short cutoff radii.
151  
152 < \begin{figure}[h!]
152 > Direct truncation of interactions at a cutoff radius creates numerical
153 > errors.  Wolf \textit{et al.} suggest that truncation errors are due
154 > to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
155 > neutralize this charge they proposed placing an image charge on the
156 > surface of the cutoff sphere for every real charge inside the cutoff.
157 > These charges are present for the evaluation of both the pair
158 > interaction energy and the force, although the force expression
159 > maintains a discontinuity at the cutoff sphere.  In the original Wolf
160 > formulation, the total energy for the charge and image were not equal
161 > to the integral of the force expression, and as a result, the total
162 > energy would not be conserved in molecular dynamics (MD)
163 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
164 > Gezelter later proposed shifted force variants of the Wolf method with
165 > commensurate force and energy expressions that do not exhibit this
166 > problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
167 > were also proposed by Chen \textit{et
168 >  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
169 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
170 > used additional neutralization of higher order moments for systems of
171 > point charges.\cite{Fukuda:2013sf}
172 >
173 > \begin{figure}
174    \centering
175 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
176 <  \caption{Top: NaCl crystal showing how spherical truncation can
177 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
178 <    molecules interact with the central ion.  Bottom: A dipolar
179 <    crystal exhibiting similar behavior and illustrating how the
180 <    effective dipole-octupole interactions can be disrupted by
181 <    spherical truncation.}
182 <  \label{fig:NaCl}
175 >  \includegraphics[width=\linewidth]{schematic.pdf}
176 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
177 >    charges (in the smaller grey circle), so interactions are
178 >    effectively charge-multipole at longer distances.  With hard
179 >    cutoffs, motion of individual charges in and out of the cutoff
180 >    sphere can break the effective multipolar ordering.  Bottom:
181 >    dipolar crystals and fluids have a similar effective
182 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
183 >    orientational averaging helps to reduce the effective range of the
184 >    interactions in the fluid.  Placement of reversed image multipoles
185 >    on the surface of the cutoff sphere recovers the effective
186 >    higher-order multipole behavior.}
187 >  \label{fig:schematic}
188   \end{figure}
189  
190 < The direct truncation of interactions at a cutoff radius creates
191 < truncation defects. Wolf \textit{et al.} further argued that
192 < truncation errors are due to net charge remaining inside the cutoff
193 < sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
194 < placing an image charge on the surface of the cutoff sphere for every
195 < real charge inside the cutoff.  These charges are present for the
196 < evaluation of both the pair interaction energy and the force, although
197 < the force expression maintained a discontinuity at the cutoff sphere.
198 < In the original Wolf formulation, the total energy for the charge and
199 < 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
190 > One can make a similar effective range argument for crystals of point
191 > \textit{multipoles}. The Luttinger and Tisza treatment of energy
192 > constants for dipolar lattices utilizes 24 basis vectors that contain
193 > dipoles at the eight corners of a unit cube.\cite{LT} Only three of
194 > these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
195 > moments, while the rest have zero net dipole and retain contributions
196 > only from higher order multipoles.  The lowest-energy crystalline
197 > structures are built out of basis vectors that have only residual
198 > quadrupolar moments (e.g. the $Z_5$ array). In these low energy
199 > structures, the effective interaction between a dipole at the center
200   of a crystal and a group of eight dipoles farther away is
201   significantly shorter ranged than the $r^{-3}$ that one would expect
202   for raw dipole-dipole interactions.  Only in crystals which retain a
# Line 188 | Line 206 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
206   unstable.
207  
208   In ionic crystals, real-space truncation can break the effective
209 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
210 < swings in the electrostatic energy as the cutoff radius is increased
211 < (or as individual ions move back and forth across the boundary).  This
212 < is why the image charges were necessary for the Wolf sum to exhibit
213 < rapid convergence.  Similarly, the real-space truncation of point
214 < multipole interactions breaks higher order multipole arrangements, and
215 < image multipoles are required for real-space treatments of
198 < electrostatic energies.
209 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
210 > significant swings in the electrostatic energy as individual ions move
211 > back and forth across the boundary.  This is why the image charges are
212 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
213 > the real-space truncation of point multipole interactions breaks
214 > higher order multipole arrangements, and image multipoles are required
215 > for real-space treatments of electrostatic energies.
216  
217 + The shorter effective range of electrostatic interactions is not
218 + limited to perfect crystals, but can also apply in disordered fluids.
219 + Even at elevated temperatures, there is local charge balance in an
220 + ionic liquid, where each positive ion has surroundings dominated by
221 + negaitve ions and vice versa.  The reversed-charge images on the
222 + cutoff sphere that are integral to the Wolf and DSF approaches retain
223 + the effective multipolar interactions as the charges traverse the
224 + cutoff boundary.
225 +
226 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
227 + significant orientational averaging that additionally reduces the
228 + effect of long-range multipolar interactions.  The image multipoles
229 + that are introduced in the TSF, GSF, and SP methods mimic this effect
230 + and reduce the effective range of the multipolar interactions as
231 + interacting molecules traverse each other's cutoff boundaries.
232 +
233   % Because of this reason, although the nature of electrostatic
234   % interaction short ranged, the hard cutoff sphere creates very large
235   % fluctuation in the electrostatic energy for the perfect crystal. In
# Line 207 | Line 240 | The forces and torques acting on atomic sites are the
240   % to the non-neutralized value of the higher order moments within the
241   % cutoff sphere.
242  
243 < The forces and torques acting on atomic sites are the fundamental
244 < factors driving dynamics in molecular simulations. Fennell and
245 < Gezelter proposed the damped shifted force (DSF) energy kernel to
246 < obtain consistent energies and forces on the atoms within the cutoff
247 < sphere. Both the energy and the force go smoothly to zero as an atom
248 < aproaches the cutoff radius. The comparisons of the accuracy these
249 < quantities between the DSF kernel and SPME was surprisingly
250 < good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
251 < calculating electrostatic interactions in molecular systems with
252 < relatively uniform charge
220 < densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
243 > Forces and torques acting on atomic sites are fundamental in driving
244 > dynamics in molecular simulations, and the damped shifted force (DSF)
245 > energy kernel provides consistent energies and forces on charged atoms
246 > within the cutoff sphere. Both the energy and the force go smoothly to
247 > zero as an atom aproaches the cutoff radius. The comparisons of the
248 > accuracy these quantities between the DSF kernel and SPME was
249 > surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
250 > has seen increasing use in molecular systems with relatively uniform
251 > charge
252 > densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
253  
254   \subsection{The damping function}
255 < The damping function used in our research has been discussed in detail
256 < in the first paper of this series.\cite{PaperI} The radial kernel
257 < $1/r$ for the interactions between point charges can be replaced by
258 < the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
259 < accelerate the rate of convergence, where $\alpha$ is a damping
260 < parameter with units of inverse distance.  Altering the value of
261 < $\alpha$ is equivalent to changing the width of Gaussian charge
262 < distributions that replace each point charge -- Gaussian overlap
263 < integrals yield complementary error functions when truncated at a
264 < finite distance.
255 > The damping function has been discussed in detail in the first paper
256 > of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
257 > interactions between point charges can be replaced by the
258 > complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
259 > convergence, where $\alpha$ is a damping parameter with units of
260 > inverse distance.  Altering the value of $\alpha$ is equivalent to
261 > changing the width of Gaussian charge distributions that replace each
262 > point charge, as Coulomb integrals with Gaussian charge distributions
263 > produce complementary error functions when truncated at a finite
264 > distance.
265  
266 < By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
267 < cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
268 < agreement with SPME for the interaction energies, forces and torques
269 < for charge-charge interactions.\cite{Fennell:2006lq}
266 > With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
267 > produced very good agreement with SPME for interaction energies,
268 > forces and torques for charge-charge
269 > interactions.\cite{Fennell:2006lq}
270  
271   \subsection{Point multipoles in molecular modeling}
272   Coarse-graining approaches which treat entire molecular subsystems as
273   a single rigid body are now widely used. A common feature of many
274   coarse-graining approaches is simplification of the electrostatic
275   interactions between bodies so that fewer site-site interactions are
276 < required to compute configurational energies.  Many coarse-grained
277 < 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}
276 > required to compute configurational
277 > energies.\cite{Ren06,Essex10,Essex11}
278  
279 < Because electrons in a molecule are not localized at specific points,
280 < the assignment of partial charges to atomic centers is a relatively
281 < rough approximation.  Atomic sites can also be assigned point
282 < multipoles and polarizabilities to increase the accuracy of the
283 < molecular model.  Recently, water has been modeled with point
284 < multipoles up to octupolar
285 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
279 > Additionally, because electrons in a molecule are not localized at
280 > specific points, the assignment of partial charges to atomic centers
281 > is always an approximation.  For increased accuracy, atomic sites can
282 > also be assigned point multipoles and polarizabilities.  Recently,
283 > water has been modeled with point multipoles up to octupolar order
284 > using the soft sticky dipole-quadrupole-octupole (SSDQO)
285 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
286   multipoles up to quadrupolar order have also been coupled with point
287   polarizabilities in the high-quality AMOEBA and iAMOEBA water
288 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
289 < using point multipole with the real space truncation without
290 < accounting for multipolar neutrality will create energy conservation
291 < issues in molecular dynamics (MD) simulations.
288 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
289 > truncating point multipoles without smoothing the forces and torques
290 > can create energy conservation issues in molecular dynamics
291 > simulations.
292  
293   In this paper we test a set of real-space methods that were developed
294   for point multipolar interactions.  These methods extend the damped
295   shifted force (DSF) and Wolf methods originally developed for
296   charge-charge interactions and generalize them for higher order
297 < multipoles. The detailed mathematical development of these methods has
298 < been presented in the first paper in this series, while this work
299 < covers the testing the energies, forces, torques, and energy
297 > multipoles.  The detailed mathematical development of these methods
298 > has been presented in the first paper in this series, while this work
299 > covers the testing of energies, forces, torques, and energy
300   conservation properties of the methods in realistic simulation
301   environments.  In all cases, the methods are compared with the
302 < reference method, a full multipolar Ewald treatment.
302 > reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
303  
304  
305   %\subsection{Conservation of total energy }
# Line 297 | Line 325 | where the multipole operator for site $\bf a$,
325   \begin{equation}
326   U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
327   \end{equation}
328 < where the multipole operator for site $\bf a$,
329 < \begin{equation}
330 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
331 < +  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.
328 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
329 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
330 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
331 > $\bf a$, etc.
332  
333 < Interactions between multipoles can be expressed as higher derivatives
334 < of the bare Coulomb potential, so one way of ensuring that the forces
335 < and torques vanish at the cutoff distance is to include a larger
336 < number of terms in the truncated Taylor expansion, e.g.,
337 < %
338 < \begin{equation}
339 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
340 < \end{equation}
341 < %
342 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
343 < Thus, for $f(r)=1/r$, we find
344 < %
345 < \begin{equation}
346 < 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} .
347 < \end{equation}
348 < This function is an approximate electrostatic potential that has
349 < vanishing second derivatives at the cutoff radius, making it suitable
350 < for shifting the forces and torques of charge-dipole interactions.
333 > % Interactions between multipoles can be expressed as higher derivatives
334 > % of the bare Coulomb potential, so one way of ensuring that the forces
335 > % and torques vanish at the cutoff distance is to include a larger
336 > % number of terms in the truncated Taylor expansion, e.g.,
337 > % %
338 > % \begin{equation}
339 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
340 > % \end{equation}
341 > % %
342 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
343 > % Thus, for $f(r)=1/r$, we find
344 > % %
345 > % \begin{equation}
346 > % 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} .
347 > % \end{equation}
348 > % This function is an approximate electrostatic potential that has
349 > % vanishing second derivatives at the cutoff radius, making it suitable
350 > % for shifting the forces and torques of charge-dipole interactions.
351  
352 < In general, the TSF potential for any multipole-multipole interaction
353 < can be written
352 > The TSF potential for any multipole-multipole interaction can be
353 > written
354   \begin{equation}
355   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
356   \label{generic}
357   \end{equation}
358 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
359 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
360 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
361 < energy, force, and torques, the required number of terms from Taylor
362 < series expansion in $f_n(r)$ must be performed for different
363 < multipole-multipole interactions.
358 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
359 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
360 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
361 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
362 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
363 > force, and torques, a Taylor expansion with $n$ terms must be
364 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
365  
366 < To carry out the same procedure for a damped electrostatic kernel, we
367 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
368 < Many of the derivatives of the damped kernel are well known from
369 < Smith's early work on multipoles for the Ewald
370 < summation.\cite{Smith82,Smith98}
366 > % To carry out the same procedure for a damped electrostatic kernel, we
367 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
368 > % Many of the derivatives of the damped kernel are well known from
369 > % Smith's early work on multipoles for the Ewald
370 > % summation.\cite{Smith82,Smith98}
371  
372 < Note that increasing the value of $n$ will add additional terms to the
373 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
374 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
375 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
376 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
377 < for computing multipole energies, forces, and torques, and smooth
378 < cutoffs of these quantities can be guaranteed as long as the number of
379 < terms in the Taylor series exceeds the derivative order required.
372 > % Note that increasing the value of $n$ will add additional terms to the
373 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
374 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
375 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
376 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
377 > % for computing multipole energies, forces, and torques, and smooth
378 > % cutoffs of these quantities can be guaranteed as long as the number of
379 > % terms in the Taylor series exceeds the derivative order required.
380  
381   For multipole-multipole interactions, following this procedure results
382 < in separate radial functions for each distinct orientational
383 < contribution to the potential, and ensures that the forces and torques
384 < from {\it each} of these contributions will vanish at the cutoff
385 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
386 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
382 > in separate radial functions for each of the distinct orientational
383 > contributions to the potential, and ensures that the forces and
384 > torques from each of these contributions will vanish at the cutoff
385 > radius.  For example, the direct dipole dot product
386 > ($\mathbf{D}_{\bf a}
387 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
388   dot products:
389   \begin{equation}
390 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
391 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
392 < -\frac{1}{4\pi \epsilon_0}
393 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
394 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
390 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
391 >  \mathbf{D}_{\bf a} \cdot
392 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
393 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
394 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
395   \end{equation}
396  
397 < The electrostatic forces and torques acting on the central multipole
398 < site due to another site within cutoff sphere are derived from
397 > For the Taylor shifted (TSF) method with the undamped kernel,
398 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
399 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
400 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
401 > connection to unmodified electrostatics as well as the smooth
402 > transition to zero in both these functions as $r\rightarrow r_c$.  The
403 > electrostatic forces and torques acting on the central multipole due
404 > to another site within the cutoff sphere are derived from
405   Eq.~\ref{generic}, accounting for the appropriate number of
406   derivatives. Complete energy, force, and torque expressions are
407   presented in the first paper in this series (Reference
408 < \citep{PaperI}).
408 > \onlinecite{PaperI}).
409  
410   \subsection{Gradient-shifted force (GSF)}
411  
412 < A second (and significantly simpler) method involves shifting the
413 < gradient of the raw coulomb potential for each particular multipole
412 > A second (and conceptually simpler) method involves shifting the
413 > gradient of the raw Coulomb potential for each particular multipole
414   order.  For example, the raw dipole-dipole potential energy may be
415   shifted smoothly by finding the gradient for two interacting dipoles
416   which have been projected onto the surface of the cutoff sphere
417   without changing their relative orientation,
418 < \begin{displaymath}
419 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
420 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
421 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
422 < \end{displaymath}
423 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
424 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
425 < the signs are reversed for the dipole that has been projected onto the
426 < cutoff sphere).  In many ways, this simpler approach is closer in
427 < spirit to the original shifted force method, in that it projects a
428 < neutralizing multipole (and the resulting forces from this multipole)
429 < onto a cutoff sphere. The resulting functional forms for the
430 < potentials, forces, and torques turn out to be quite similar in form
431 < to the Taylor-shifted approach, although the radial contributions are
432 < significantly less perturbed by the Gradient-shifted approach than
433 < they are in the Taylor-shifted method.
418 > \begin{equation}
419 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
420 > U_{D_{\bf a} D_{\bf b}}(r_c)
421 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
422 >  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
423 > \end{equation}
424 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
425 >  a}$ and $\mathbf{D}_{\bf b}$, 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}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
447 >    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
448 >    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{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 $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
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}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
482 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{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 477 | Line 520 | arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} Thi
520   real-space cutoffs.  In the first paper of this series, we compared
521   the dipolar and quadrupolar energy expressions against analytic
522   expressions for ordered dipolar and quadrupolar
523 < arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
524 < multipolar Ewald sum as a reference method for comparing energies,
525 < forces, and torques for molecular models that mimic disordered and
526 < ordered condensed-phase systems.  These test-cases include:
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.  The parameters used
527 > in the test cases are given in table~\ref{tab:pars}.
528  
529 < \begin{itemize}
530 < \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
531 < \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
532 < \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
533 < \item Soft Quadrupolar solids  ($\sigma = , \epsilon = , Q_{xx} = ...$)
534 < \item A mixed multipole model for water
535 < \item A mixed multipole models for water with dissolved ions
536 < \end{itemize}
537 < This last test case exercises all levels of the multipole-multipole
538 < interactions we have derived so far and represents the most complete
539 < test of the new methods.
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   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 obtaine from the multipolar
573 < Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
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 imporant
576   indicators of the reliability of a method even if the absolute
# Line 510 | Line 581 | contributions to the forces and torques.
581   behavior of the simulation, so we also compute the electrostatic
582   contributions to the forces and torques.
583  
584 < \subsection{Model systems}
585 < To sample independent configurations of multipolar crystals, a body
586 < centered cubic (BCC) crystal which is a minimum energy structure for
587 < point dipoles was generated using 3,456 molecules.  The multipoles
588 < were translationally locked in their respective crystal sites for
589 < equilibration at a relatively low temperature (50K), so that dipoles
590 < or quadrupoles could freely explore all accessible orientations.  The
591 < translational constraints were removed, and the crystals were
521 < simulated for 10 ps in the microcanonical (NVE) ensemble with an
522 < average temperature of 50 K.  Configurations were sampled at equal
523 < time intervals for the comparison of the configurational energy
524 < differences.  The crystals were not simulated close to the melting
525 < points in order to avoid translational deformation away of the ideal
526 < lattice geometry.
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{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 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
594 < system was created with 2048 molecules oriented randomly.  These were
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 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
603 < temperature after equilibration.  We collected 250 different
604 < configurations in equal interval of time. For the ions mixed liquid
605 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
606 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
607 < at the same environment for 1ns and 250 configurations were
608 < collected. While comparing energies, forces, and torques with Ewald
609 < method, Lennard-Jones potentials were turned off and purely
610 < electrostatic interaction had been compared.
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 550 | 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 561 | 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 647 | Line 757 | model must allow for long simulation times with minima
757        
758   %        \label{fig:barGraph2}
759   %      \end{figure}
760 < %The correlation coefficient ($R^2$) and slope of the linear regression plots for the energy differences for all six different molecular systems is shown in figure 4a and 4b.The plot shows that the correlation coefficient improves for the SP cutoff method as compared to the undamped hard cutoff method in the case of SSDQC, SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar crystal and liquid, the correlation coefficient is almost unchanged and close to 1.  The correlation coefficient is smallest (0.696276 for $r_c$ = 9 $A^o$) for the SSDQC liquid because of the presence of charge-charge and charge-multipole interactions. Since the charge-charge and charge-multipole interaction is long ranged, there is huge deviation of correlation coefficient from 1. Similarly, the quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with compared to interactions in the other multipolar systems, thus the correlation coefficient very close to 1 even for hard cutoff method. The idea of placing image multipole on the surface of the cutoff sphere improves the correlation coefficient and makes it close to 1 for all types of multipolar systems. Similarly the slope is hugely deviated from the correct value for the lower order multipole-multipole interaction and slightly deviated for higher order multipole – multipole interaction. The SP method improves both correlation coefficient ($R^2$) and slope significantly in SSDQC and dipolar systems.  The Slope is found to be deviated more in dipolar crystal as compared to liquid which is associated with the large fluctuation in the electrostatic energy in crystal. The GSF also produced better values of correlation coefficient and slope with the proper selection of the damping alpha (Interested reader can consult accompanying supporting material). The TSF method gives good value of correlation coefficient for the dipolar crystal, dipolar liquid, SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the regression slopes are significantly deviated.
760 > %The correlation coefficient ($R^2$) and slope of the linear
761 > %regression plots for the energy differences for all six different
762 > %molecular systems is shown in figure 4a and 4b.The plot shows that
763 > %the correlation coefficient improves for the SP cutoff method as
764 > %compared to the undamped hard cutoff method in the case of SSDQC,
765 > %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
766 > %crystal and liquid, the correlation coefficient is almost unchanged
767 > %and close to 1.  The correlation coefficient is smallest (0.696276
768 > %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
769 > %charge-charge and charge-multipole interactions. Since the
770 > %charge-charge and charge-multipole interaction is long ranged, there
771 > %is huge deviation of correlation coefficient from 1. Similarly, the
772 > %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
773 > %compared to interactions in the other multipolar systems, thus the
774 > %correlation coefficient very close to 1 even for hard cutoff
775 > %method. The idea of placing image multipole on the surface of the
776 > %cutoff sphere improves the correlation coefficient and makes it close
777 > %to 1 for all types of multipolar systems. Similarly the slope is
778 > %hugely deviated from the correct value for the lower order
779 > %multipole-multipole interaction and slightly deviated for higher
780 > %order multipole – multipole interaction. The SP method improves both
781 > %correlation coefficient ($R^2$) and slope significantly in SSDQC and
782 > %dipolar systems.  The Slope is found to be deviated more in dipolar
783 > %crystal as compared to liquid which is associated with the large
784 > %fluctuation in the electrostatic energy in crystal. The GSF also
785 > %produced better values of correlation coefficient and slope with the
786 > %proper selection of the damping alpha (Interested reader can consult
787 > %accompanying supporting material). The TSF method gives good value of
788 > %correlation coefficient for the dipolar crystal, dipolar liquid,
789 > %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
790 > %regression slopes are significantly deviated.
791 >
792   \begin{figure}
793 <        \centering
794 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
795 <        \caption{The correlation coefficient and regression slope of configurational energy differences for a given method with compared with the reference Ewald method. The value of result equal to 1(dashed line) indicates energy difference is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle)}
796 <        \label{fig:slopeCorr_energy}
797 <    \end{figure}
798 < The combined correlation coefficient and slope for all six systems is shown in Figure ~\ref{fig:slopeCorr_energy}. The correlation coefficient for the undamped hard cutoff method is does not have good agreement with the Ewald because of the fluctuation of the electrostatic energy in the direct truncation method. This deviation in correlation coefficient is improved by using SP, GSF, and TSF method. But the TSF method worsens the regression slope stating that this method produces statistically more biased result as compared to Ewald. Also the GSF method slightly deviate slope but it can be alleviated by using proper value of damping alpha and cutoff radius. The SP method shows good agreement with Ewald method for all values of damping alpha and radii.
793 >  \centering
794 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
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
798 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
799 >    from those obtained using the multipolar Ewald sum.  Different
800 >    values of the cutoff radius are indicated with different symbols
801 >    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
802 >    triangles).}
803 >  \label{fig:slopeCorr_energy}
804 > \end{figure}
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 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 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
821 > systematic bias in the values of the energies, particularly for
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, 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 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 < The comparison of the magnitude of the combined forces and torques for the data accumulated from all system types are shown in Figure ~\ref{fig:slopeCorr_force}. The correlation and slope for the forces agree with the Ewald even for the hard cutoff method. For the system of molecules with higher order multipoles, the interaction is short ranged. Moreover, the force decays more rapidly than the electrostatic energy hence the hard cutoff method also produces good results. Although the pure cutoff gives the good match of the electrostatic force, the discontinuity in the force at the cutoff radius causes problem in the total energy conservation in MD simulations, which will be discussed in detail in subsection D. The correlation coefficient for GSF method also perfectly matches with Ewald but the slope is slightly deviated (due to extra term obtained from the angular differentiation). This deviation in the slope can be alleviated with proper selection of the damping alpha and radii ($\alpha = 0.2$ and $r_c = 12 A^o$ are good choice). The TSF method shows good agreement in the correlation coefficient but the slope is not good as compared to the Ewald.
835 >
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 cutoffs.
841 >
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 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
863 > forces is desired.
864 >
865   \begin{figure}
866 <        \centering
867 <        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
868 <        \caption{The correlation coefficient and regression slope of the magnitude of the force for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle). }
869 <        \label{fig:slopeCorr_force}
870 <    \end{figure}
871 < The torques appears to be very influenced because of extra term generated when the potential energy is modified to get consistent force and torque.  The result shows that the torque from the hard cutoff method has good agreement with Ewald. As the potential is modified to make it consistent with the force and torque, the correlation and slope is deviated as shown in Figure~\ref{fig:slopeCorr_torque} for SP, GSF and TSF cutoff methods.  But the proper value of the damping alpha and radius can improve the agreement of the GSF with the Ewald method. The TSF method shows worst agreement in the slope as compared to Ewald even for larger cutoff radii.
872 < \begin{figure}
873 <        \centering
874 <        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
875 <        \caption{The correlation coefficient and regression slope of the magnitude of the torque for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle).}
876 <        \label{fig:slopeCorr_torque}
877 <    \end{figure}
878 < \subsection{Directionality of the force and torque vectors}  
879 < The accurate evaluation of the direction of the force and torques are also important for the dynamic simulation.In our research, the direction data sets were computed from the purposed method and compared with Ewald using Fisher statistics and results are expressed in terms of circular variance ($Var(\theta$).The force and torque vectors from the purposed method followed Fisher probability distribution function expressed in equation~\ref{eq:pdf}. The circular variance for the force and torque vectors of each molecule in the 250 configurations for all system types is shown in Figure~\ref{fig:slopeCorr_circularVariance}. The direction of the force and torque vectors from hard and SP cutoff methods showed best directional agreement with the Ewald. The force and torque vectors from GSF method also showed good agreement with the Ewald method, which can also be improved by varying damping alpha and cutoff radius.For $\alpha = 0.2$ and $r_c = 12 A^o$, $ Var(\theta) $ for direction of the force was found to be 0.002061 and corresponding value of $\kappa $ was 485.20. Integration of equation ~\ref{eq:pdf} for that corresponding value of $\kappa$ showed that 95\% of force vectors are with in $6.37^o$. The TSF method is the poorest in evaluating accurate direction with compared to Hard, SP, and GSF methods. The circular variance for the direction of the torques is larger as compared to force. For same $\alpha = 0.2, r_c = 12 A^o$ and GSF method, the circular variance was 0.01415, which showed 95\% of torque vectors are within $16.75^o$.The direction of the force and torque vectors can be improved by varying $\alpha$ and $r_c$.
866 >  \centering
867 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
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
871 >    line) indicate force magnitude values indistinguishable from those
872 >    obtained using the multipolar Ewald sum.  Different values of the
873 >    cutoff radius are indicated with different symbols (9\AA\ =
874 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
875 >  \label{fig:slopeCorr_force}
876 > \end{figure}
877 >
878 >
879 > \begin{figure}
880 >  \centering
881 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
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
885 >    line) indicate force magnitude values indistinguishable from those
886 >    obtained using the multipolar Ewald sum.  Different values of the
887 >    cutoff radius are indicated with different symbols (9\AA\ =
888 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
889 >  \label{fig:slopeCorr_torque}
890 > \end{figure}
891  
892 + The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
893 + significantly influenced by the choice of real-space method.  The
894 + torque expressions have the same distance dependence as the energies,
895 + which are naturally longer-ranged expressions than the inter-site
896 + forces.  Torques are also quite sensitive to orientations of
897 + neighboring molecules, even those that are near the cutoff distance.
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 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 +
911 + The accurate evaluation of force and torque directions is just as
912 + important for molecular dynamics simulations as the magnitudes of
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
917 + \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
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 $r_c =
924 + 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
925 + to a distribution with 95\% of force vectors within $6.37^\circ$ of
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 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
934 + results. Again, the direction of the force and torque vectors can be
935 + systematically improved by varying $\alpha$ and $r_c$.
936 +
937   \begin{figure}
938 <        \centering
939 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
940 <        \caption{The circular variance of the data sets of the
941 <          direction of the  force and torque vectors obtained from a
942 <          given method about reference Ewald method. The result equal
943 <          to 0 (dashed line) indicates direction of the vectors are
944 <          indistinguishable from the Ewald method. Here different
945 <          symbols represent different value of the cutoff radius (9
946 <          \AA\ = circle, 12 \AA\ = square, 15 \AA\  = inverted triangle)}
947 <        \label{fig:slopeCorr_circularVariance}
948 <    \end{figure}
688 < \subsection{Total energy conservation}
689 < We have tested the conservation of energy in the SSDQC liquid system
690 < by running system for 1ns in the Hard, SP, GSF and TSF method. The
691 < Hard cutoff method shows very high energy drifts 433.53
692 < KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol
693 < (measured by the SD from the slope) for the undamped case, which makes
694 < it completely unusable in MD simulations. The SP method also shows
695 < large value of energy drift 1.289 Kcal/Mol/ns/particle and energy
696 < fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method
697 < is due to the non-vanishing nature of the torque and force at the
698 < cutoff radius. We can improve the energy conservation in some extent
699 < by the proper selection of the damping alpha but the improvement is
700 < not good enough, which can be observed in Figure 9a and 9b .The GSF
701 < and TSF shows very low value of energy drift 0.09016, 0.07371
702 < KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol
703 < respectively for the undamped case. Since the absolute value of the
704 < evaluated electrostatic energy, force and torque from TSF method are
705 < deviated from the Ewald, it does not mimic MD simulations
706 < appropriately. The electrostatic energy, force and torque from the GSF
707 < method have very good agreement with the Ewald. In addition, the
708 < energy drift and energy fluctuation from the GSF method is much better
709 < than Ewald’s method for reciprocal space vector value ($k_f$) equal to
710 < 7 as shown in Figure~\ref{fig:energyDrift} and
711 < ~\ref{fig:fluctuation}. We can improve the total energy fluctuation
712 < and drift for the Ewald’s method by increasing size of the reciprocal
713 < space, which extremely increseses the simulation time. In our current
714 < simulation, the simulation time for the Hard, SP, and GSF methods are
715 < about 5.5 times faster than the Ewald method.
938 >  \centering
939 >  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
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)
943 >    indicates direction of the force or torque vectors are
944 >    indistinguishable from those obtained from the Ewald sum. Here
945 >    different symbols represent different values of the cutoff radius
946 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
947 >  \label{fig:slopeCorr_circularVariance}
948 > \end{figure}
949  
950 < In Fig.~\ref{fig:energyDrift}, $\delta \mbox{E}_1$ is a measure of the
718 < linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
719 < over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
720 < standard deviation of the energy fluctuations in units of $\mbox{kcal
721 <  mol}^{-1}$ per particle. In the bottom plot, it is apparent that the
722 < energy drift is reduced by a significant amount (2 to 3 orders of
723 < magnitude improvement at all values of the damping coefficient) by
724 < chosing either of the shifted-force methods over the hard or SP
725 < methods.  We note that the two shifted-force method can give
726 < significantly better energy conservation than the multipolar Ewald sum
727 < with the same choice of real-space cutoffs.
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
954 + fraction of solvated ions. This is a test system which exercises all
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 12\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 all of the real space methods.  A
964 + sample was also run using the multipolar Ewald sum with the same
965 + real-space cutoff.
966 +
967 + In figure~\ref{fig:energyDrift} we show the both the linear drift in
968 + energy over time, $\delta E_1$, and the standard deviation of energy
969 + fluctuations around this drift $\delta E_0$.  Both of the
970 + shifted-force methods (GSF and TSF) provide excellent energy
971 + conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
972 + while the hard cutoff is essentially unusable for molecular dynamics.
973 + SP provides some benefit over the hard cutoff because the energetic
974 + jumps that happen as particles leave and enter the cutoff sphere are
975 + somewhat reduced, but like the Wolf method for charges, the SP method
976 + would not be as useful for molecular dynamics as either of the
977 + shifted-force methods.
978 +
979 + We note that for all tested values of the cutoff radius, the new
980 + real-space methods can provide better energy conservation behavior
981 + than the multipolar Ewald sum, even when utilizing a relatively large
982 + $k$-space cutoff values.
983 +
984   \begin{figure}
985    \centering
986 <  \includegraphics[width=\textwidth]{newDrift.pdf}
986 >  \includegraphics[width=\textwidth]{newDrift_12.pdf}
987   \label{fig:energyDrift}        
988 < \caption{Analysis of the energy conservation of the real space
988 > \caption{Analysis of the energy conservation of the real-space
989    electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
990 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
991 <  of energy fluctuations around this drift.  All simulations were of a
992 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
993 <  K starting from the same initial configuration.}
990 >  energy over time (in kcal / mol / particle / ns) and $\delta
991 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
992 >  around this drift (in kcal / mol / particle).  All simulations were
993 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
994 >  300 K starting from the same initial configuration. All runs
995 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
996   \end{figure}
997  
998 +
999   \section{CONCLUSION}
1000 < We have generalized the charged neutralized potential energy originally developed by the Wolf et al.\cite{Wolf:1999dn} for the charge-charge interaction to the charge-multipole and multipole-multipole interaction in the SP method for higher order multipoles. Also, we have developed GSF and TSF methods by implementing the modification purposed by Fennel and Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the higher order multipoles to ensure consistency and smooth truncation of the electrostatic energy, force, and torque for the spherical truncation. The SP methods for multipoles proved its suitability in MC simulations. On the other hand, the results from the GSF method produced good agreement with the Ewald's energy, force, and torque. Also, it shows very good energy conservation in MD simulations.
1001 < The direct truncation of any molecular system without multipole neutralization creates the fluctuation in the electrostatic energy. This fluctuation in the energy is very large for the case of crystal because of long range of multipole ordering (Refer paper I).\cite{PaperI} This is also significant in the case of the liquid because of the local multipole ordering in the molecules. If the net multipole within cutoff radius neutralized within cutoff sphere by placing image multiples on the surface of the sphere, this fluctuation in the energy reduced significantly. Also, the multipole neutralization in the generalized SP method showed very good agreement with the Ewald as compared to direct truncation for the evaluation of the $\triangle E$ between the configurations.
1002 < In MD simulations, the energy conservation is very important. The
1003 < conservation of the total energy can be ensured by  i) enforcing the
1004 < smooth truncation of the energy, force and torque in the cutoff radius
1005 < and ii) making the energy, force and torque consistent with each
1006 < other. The GSF and TSF methods ensure the consistency and smooth
1007 < truncation of the energy, force and torque at the cutoff radius, as a
1008 < result show very good total energy conservation. But the TSF method
751 < does not show good agreement in the absolute value of the
752 < electrostatic energy, force and torque with the Ewald.  The GSF method
753 < has mimicked Ewald’s force, energy and torque accurately and also
754 < conserved energy. Therefore, the GSF method is the suitable method for
755 < evaluating required force field in MD simulations. In addition, the
756 < energy drift and fluctuation from the GSF method is much better than
757 < Ewald’s method for finite-sized reciprocal space.
1000 > In the first paper in this series, we generalized the
1001 > charge-neutralized electrostatic energy originally developed by Wolf
1002 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1003 > up to quadrupolar order.  The SP method is essentially a
1004 > multipole-capable version of the Wolf model.  The SP method for
1005 > multipoles provides excellent agreement with Ewald-derived energies,
1006 > forces and torques, and is suitable for Monte Carlo simulations,
1007 > although the forces and torques retain discontinuities at the cutoff
1008 > distance that prevents its use in molecular dynamics.
1009  
1010 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1011 < that can be made extremely efficient using spline interpolations of
1012 < the radial functions.  They require no Fourier transforms or $k$-space
1013 < sums, and guarantee the smooth handling of energies, forces, and
1014 < torques as multipoles cross the real-space cutoff boundary.  
1010 > We also developed two natural extensions of the damped shifted-force
1011 > (DSF) model originally proposed by Fennel and
1012 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1013 > smooth truncation of energies, forces, and torques at the real-space
1014 > cutoff, and both converge to DSF electrostatics for point-charge
1015 > interactions.  The TSF model is based on a high-order truncated Taylor
1016 > expansion which can be relatively perturbative inside the cutoff
1017 > sphere.  The GSF model takes the gradient from an images of the
1018 > interacting multipole that has been projected onto the cutoff sphere
1019 > to derive shifted force and torque expressions, and is a significantly
1020 > more gentle approach.
1021  
1022 + Of the two newly-developed shifted force models, the GSF method
1023 + produced quantitative agreement with Ewald energy, force, and torques.
1024 + It also performs well in conserving energy in MD simulations.  The
1025 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1026 + place on a potential energy surface that is significantly perturbed
1027 + from Ewald-based electrostatics.  
1028 +
1029 + % The direct truncation of any electrostatic potential energy without
1030 + % multipole neutralization creates large fluctuations in molecular
1031 + % simulations.  This fluctuation in the energy is very large for the case
1032 + % of crystal because of long range of multipole ordering (Refer paper
1033 + % I).\cite{PaperI} This is also significant in the case of the liquid
1034 + % because of the local multipole ordering in the molecules. If the net
1035 + % multipole within cutoff radius neutralized within cutoff sphere by
1036 + % placing image multiples on the surface of the sphere, this fluctuation
1037 + % in the energy reduced significantly. Also, the multipole
1038 + % neutralization in the generalized SP method showed very good agreement
1039 + % with the Ewald as compared to direct truncation for the evaluation of
1040 + % the $\triangle E$ between the configurations.  In MD simulations, the
1041 + % energy conservation is very important. The conservation of the total
1042 + % energy can be ensured by i) enforcing the smooth truncation of the
1043 + % energy, force and torque in the cutoff radius and ii) making the
1044 + % energy, force and torque consistent with each other. The GSF and TSF
1045 + % methods ensure the consistency and smooth truncation of the energy,
1046 + % force and torque at the cutoff radius, as a result show very good
1047 + % total energy conservation. But the TSF method does not show good
1048 + % agreement in the absolute value of the electrostatic energy, force and
1049 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1050 + % energy and torque accurately and also conserved energy.
1051 +
1052 + The only cases we have found where the new GSF and SP real-space
1053 + methods can be problematic are those which retain a bulk dipole moment
1054 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1055 + materials, uniform weighting of the orientational contributions can be
1056 + important for converging the total energy.  In these cases, the
1057 + damping function which causes the non-uniform weighting can be
1058 + replaced by the bare electrostatic kernel, and the energies return to
1059 + the expected converged values.
1060 +
1061 + Based on the results of this work, the GSF method is a suitable and
1062 + efficient replacement for the Ewald sum for evaluating electrostatic
1063 + interactions in MD simulations.  Both methods retain excellent
1064 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1065 + energy drift and fluctuations from the GSF electrostatics are better
1066 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1067 + Because they use real-space cutoffs with moderate cutoff radii, the
1068 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1069 + increases.  Additionally, they can be made extremely efficient using
1070 + spline interpolations of the radial functions.  They require no
1071 + Fourier transforms or $k$-space sums, and guarantee the smooth
1072 + handling of energies, forces, and torques as multipoles cross the
1073 + real-space cutoff boundary.
1074 +
1075 + \begin{acknowledgments}
1076 +  JDG acknowledges helpful discussions with Christopher
1077 +  Fennell. Support for this project was provided by the National
1078 +  Science Foundation under grant CHE-1362211. Computational time was
1079 +  provided by the Center for Research Computing (CRC) at the
1080 +  University of Notre Dame.
1081 + \end{acknowledgments}
1082 +
1083   %\bibliographystyle{aip}
1084   \newpage
1085   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines