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 4184 by gezelter, Sat Jun 14 23:35:39 2014 UTC vs.
Revision 4203 by gezelter, Wed Aug 6 19:10:04 2014 UTC

# Line 47 | Line 47 | preprint,
47  
48   %\preprint{AIP/123-QED}
49  
50 < \title{Real space alternatives to the Ewald
51 < Sum. II. Comparison of Methods} % 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
55 < 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
59 < 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,
67 <             %  but any date may be explicitly specified
64 > \date{\today}
65  
66   \begin{abstract}
67 <  We have tested the real-space shifted potential (SP),
68 <  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
69 <  for multipoles that were developed in the first paper in this series
70 <  against the multipolar Ewald sum as a reference method. The tests
71 <  were carried out in a variety of condensed-phase environments which
72 <  were designed to test all levels of the multipole-multipole
73 <  interactions.  Comparisons of the energy differences between
74 <  configurations, molecular forces, and torques were used to analyze
75 <  how well the real-space models perform relative to the more
76 <  computationally expensive Ewald treatment.  We have also investigated the
77 <  energy conservation properties of the new methods in molecular
81 <  dynamics simulations using all of these methods. The SP method shows
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 exhibits energy conservation properties that
83 <  make it an excellent choice for efficiently computing electrostatic
84 <  interactions in molecular dynamics simulations.
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
# Line 94 | Line 90 | of Notre Dame, Notre Dame, IN 46556}
90  
91   \maketitle
92  
97
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
120 < 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 method can also be problematic for molecular
121 < interfaces.\cite{Fennell:2006lq} Modified Ewald methods that were
122 < developed to handle two-dimensional (2D) electrostatic interactions in
123 < interfacial systems have not seen similar particle-mesh
124 < treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
125 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
126 < with system size.
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 phase systems is actually short ranged.\cite{Wolf92,Wolf95}
136 < For an ordered lattice (e.g., when computing the Madelung constant of
137 < an 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:schematic}).  For this reason, careful
141 < application of Wolf's method can obtain accurate estimates of Madelung
142 < constants using relatively short cutoff radii.  Recently, Fukuda used
143 < neutralization of the higher order moments for calculation of the
144 < electrostatic interactions in point charge
145 < systems.\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 + 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=\linewidth]{schematic.pdf}
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
# Line 163 | Line 189 | The direct truncation of interactions at a cutoff radi
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.}  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
176 < a result, the total energy would not be conserved in molecular
177 < dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
178 < Fennel and Gezelter later proposed shifted force variants of the Wolf
179 < method with commensurate force and energy expressions that do not
180 < exhibit this problem.\cite{Fennell:2006lq}   Related real-space
181 < methods were also proposed by Chen \textit{et
182 <  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
183 < and by Wu and Brooks.\cite{Wu:044107}
184 <
185 < Considering the interaction of one central ion in an ionic crystal
186 < with a portion of the crystal at some distance, the effective Columbic
187 < potential is found to decrease as $r^{-5}$. If one views the \ce{NaCl}
188 < crystal as a simple cubic (SC) structure with an octupolar
189 < \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
190 < rapidly to the Madelung energy than the dipolar
191 < approximation.\cite{Wolf92} To find the correct Madelung constant,
192 < Lacman suggested that the NaCl structure could be constructed in a way
193 < that the finite crystal terminates with complete \ce{(NaCl)4}
194 < molecules.\cite{Lacman65} The central ion sees what is effectively a
195 < set of octupoles at large distances. These facts suggest that the
196 < Madelung constants are relatively short ranged for perfect ionic
197 < crystals.\cite{Wolf:1999dn}
198 <
199 < One can make a similar argument for crystals of point multipoles. The
200 < Luttinger and Tisza treatment of energy constants for dipolar lattices
201 < utilizes 24 basis vectors that contain dipoles at the eight corners of
202 < a unit cube.  Only three of these basis vectors, $X_1, Y_1,
203 < \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
204 < zero net dipole and retain contributions only from higher order
205 < 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 222 | Line 218 | Even at elevated temperatures, there is, on average, l
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, on average, local charge
222 < balance in an ionic liquid, where each positive ion has surroundings
223 < dominated by negaitve ions and vice versa.  The reversed-charge images
224 < on the cutoff sphere that are integral to the Wolf and DSF approaches
225 < retain the effective multipolar interactions as the charges traverse
226 < the cutoff boundary.
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
# Line 246 | 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
259 < 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
285 < negative charges, and rather than use multiple site-site interactions,
286 < the interaction between higher order multipoles can also be used to
287 < evaluate a single molecule-molecule
288 < 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 order using the soft sticky
296 < dipole-quadrupole-octupole (SSDQO)
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 584 | Line 574 | the simulation proceeds. These differences are the mos
574   and have been compared with the values obtained from the multipolar
575   Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
576   between two configurations is the primary quantity that governs how
577 < the simulation proceeds. These differences are the most imporant
577 > the simulation proceeds. These differences are the most important
578   indicators of the reliability of a method even if the absolute
579   energies are not exact.  For each of the multipolar systems listed
580   above, we have compared the change in electrostatic potential energy
# Line 596 | Line 586 | program, OpenMD,\cite{openmd} which was used for all c
586   \subsection{Implementation}
587   The real-space methods developed in the first paper in this series
588   have been implemented in our group's open source molecular simulation
589 < program, OpenMD,\cite{openmd} which was used for all calculations in
589 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
590   this work.  The complementary error function can be a relatively slow
591   function on some processors, so all of the radial functions are
592   precomputed on a fine grid and are spline-interpolated to provide
# Line 803 | Line 793 | model must allow for long simulation times with minima
793  
794   \begin{figure}
795    \centering
796 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
796 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
797    \caption{Statistical analysis of the quality of configurational
798      energy differences for the real-space electrostatic methods
799      compared with the reference Ewald sum.  Results with a value equal
# Line 876 | Line 866 | forces is desired.
866  
867   \begin{figure}
868    \centering
869 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
869 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
870    \caption{Statistical analysis of the quality of the force vector
871      magnitudes for the real-space electrostatic methods compared with
872      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 890 | Line 880 | forces is desired.
880  
881   \begin{figure}
882    \centering
883 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
883 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
884    \caption{Statistical analysis of the quality of the torque vector
885      magnitudes for the real-space electrostatic methods compared with
886      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 948 | Line 938 | systematically improved by varying $\alpha$ and $r_c$.
938  
939   \begin{figure}
940    \centering
941 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
941 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
942    \caption{The circular variance of the direction of the force and
943      torque vectors obtained from the real-space methods around the
944      reference Ewald vectors. A variance equal to 0 (dashed line)
# Line 995 | Line 985 | $k$-space cutoff values.
985  
986   \begin{figure}
987    \centering
988 <  \includegraphics[width=\textwidth]{newDrift_12.pdf}
988 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
989   \label{fig:energyDrift}        
990   \caption{Analysis of the energy conservation of the real-space
991    electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
# Line 1005 | Line 995 | $k$-space cutoff values.
995    of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
996    300 K starting from the same initial configuration. All runs
997    utilized the same real-space cutoff, $r_c = 12$\AA.}
998 + \end{figure}
999 +
1000 + \subsection{Reproduction of Structural Features\label{sec:structure}}
1001 + One of the best tests of modified interaction potentials is the
1002 + fidelity with which they can reproduce structural features in a
1003 + liquid.  One commonly-utilized measure of structural ordering is the
1004 + pair distribution function, $g(r)$, which measures local density
1005 + deviations in relation to the bulk density.  In the electrostatic
1006 + approaches studied here, the short-range repulsion from the
1007 + Lennard-Jones potential is identical for the various electrostatic
1008 + methods, and since short range repulsion determines much of the local
1009 + liquid ordering, one would not expect to see any differences in
1010 + $g(r)$.  Indeed, the pair distributions are essentially identical for
1011 + all of the electrostatic methods studied (for each of the different
1012 + systems under investigation).  Interested readers may consult the
1013 + supplementary information for plots of these pair distribution
1014 + functions.
1015 +
1016 + A direct measure of the structural features that is a more
1017 + enlightening test of the modified electrostatic methods is the average
1018 + value of the electrostatic energy $\langle U_\mathrm{elect} \rangle$
1019 + which is obtained by sampling the liquid-state configurations
1020 + experienced by a liquid evolving entirely under the influence of the
1021 + methods being investigated.  In figure \ref{fig:Uelect} we show how
1022 + $\langle U_\mathrm{elect} \rangle$ for varies with the damping parameter,
1023 + $\alpha$, for each of the methods.
1024 +
1025 + \begin{figure}
1026 +  \centering
1027 +  \includegraphics[width=\textwidth]{averagePotentialEnergy_r9_12.eps}
1028 + \label{fig:Uelect}        
1029 + \caption{The average electrostatic potential energy,
1030 +  $\langle U_\mathrm{elect} \rangle$ for the SSDQ water with ions as a function
1031 +  of the damping parameter, $\alpha$, for each of the real-space
1032 +  electrostatic methods. Top panel: simulations run with a real-space
1033 +  cutoff, $r_c = 9$\AA.  Bottom panel: the same quantity, but with a
1034 +  larger cutoff, $r_c = 12$\AA.}
1035   \end{figure}
1036  
1037 + It is clear that moderate damping is important for converging the mean
1038 + potential energy values, particularly for the two shifted force
1039 + methods (GSF and TSF).  A value of $\alpha \approx 0.18$ \AA$^{-1}$ is
1040 + sufficient to converge the SP and GSF energies with a cutoff of 12
1041 + \AA, while shorter cutoffs require more dramatic damping ($\alpha
1042 + \approx 0.36$ \AA$^{-1}$ for $r_c = 9$ \AA).  It is also clear from
1043 + fig. \ref{fig:Uelect} that it is possible to overdamp the real-space
1044 + electrostatic methods, causing the estimate of the energy to drop
1045 + below the Ewald results.
1046  
1047 + These ``optimal'' values of the damping coefficient are slightly
1048 + larger than what were observed for DSF electrostatics for purely
1049 + point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1050 + $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1051 + multipole systems.
1052 +
1053 + \subsection{Reproduction of Dynamic Properties\label{sec:structure}}
1054 + To test the fidelity of the electrostatic methods at reproducing
1055 + dynamics in a multipolar liquid, it is also useful to look at
1056 + transport properties, particularly the diffusion constant,
1057 + \begin{equation}
1058 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1059 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1060 + \label{eq:diff}
1061 + \end{equation}
1062 + which measures long-time behavior and is sensitive to the forces on
1063 + the multipoles.  For the soft dipolar fluid, and the SSDQ liquid
1064 + systems, the self-diffusion constants (D) were calculated from linear
1065 + fits to the long-time portion of the mean square displacement
1066 + ($\langle r^{2}(t) \rangle$).\cite{Allen87}
1067 +
1068 + In addition to translational diffusion, orientational relaxation times
1069 + were calculated for comparisons with the Ewald simulations and with
1070 + experiments. These values were determined from the same 1~ns $NVE$
1071 + trajectories used for translational diffusion by calculating the
1072 + orientational time correlation function,
1073 + \begin{equation}
1074 + C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\gamma(t)
1075 +                \cdot\hat{\mathbf{u}}_i^\gamma(0)\right]\right\rangle,
1076 + \label{eq:OrientCorr}
1077 + \end{equation}
1078 + where $P_l$ is the Legendre polynomial of order $l$ and
1079 + $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
1080 + axis $\gamma$.  The body-fixed reference frame used for our
1081 + orientational correlation functions has the $z$-axis running along the
1082 + dipoles, and for the SSDQ water model, the $y$-axis connects the two
1083 + implied hydrogen atoms.
1084 +
1085 + From the orientation autocorrelation functions, we can obtain time
1086 + constants for rotational relaxation either by fitting an exponential
1087 + function or by integrating the entire correlation function.  These
1088 + decay times are directly comparable to water orientational relaxation
1089 + times from nuclear magnetic resonance (NMR). The relaxation constant
1090 + obtained from $C_2^y(t)$ is normally of experimental interest because
1091 + it describes the relaxation of the principle axis connecting the
1092 + hydrogen atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular
1093 + portion of the dipole-dipole relaxation from a proton NMR signal and
1094 + should provide an estimate of the NMR relaxation time
1095 + constant.\cite{Impey82}
1096 +
1097 + Results for the diffusion constants and orientational relaxation times
1098 + are shown in figure \ref{fig:dynamics}. From this data, it is apparent
1099 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1100 + reproduced with high fidelity by the GSF method.
1101 +
1102 + The $\tau_2$ results in \ref{fig:dynamics} show a much greater
1103 + difference between the real-space and the Ewald results.
1104 +
1105 +
1106   \section{CONCLUSION}
1107   In the first paper in this series, we generalized the
1108   charge-neutralized electrostatic energy originally developed by Wolf

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines