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 4180 by gezelter, Wed Jun 11 21:03:11 2014 UTC vs.
Revision 4206 by gezelter, Thu Aug 7 20:53:27 2014 UTC

# Line 35 | Line 35 | preprint,
35   %\linenumbers\relax % Commence numbering lines
36   \usepackage{amsmath}
37   \usepackage{times}
38 < \usepackage{mathptm}
38 > \usepackage{mathptmx}
39   \usepackage{tabularx}
40   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
41   \usepackage{url}
# 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 a reference method. The tests were carried out in a variety
71 <  of condensed-phase environments which were designed to test all
72 <  levels of the multipole-multipole interactions.  Comparisons of the
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 sum.  We have
76 <  also investigated the energy conservation properties of the new
77 <  methods in molecular dynamics simulations using all of these
78 <  methods. The SP method shows excellent agreement with
79 <  configurational energy differences, forces, and torques, and would
80 <  be suitable for use in Monte Carlo calculations.  Of the two new
81 <  shifted-force methods, the GSF approach shows the best agreement
82 <  with Ewald-derived energies, forces, and torques and exhibits energy
83 <  conservation properties that make it an excellent choice for
84 <  efficiently computing electrostatic interactions in molecular
88 <  dynamics simulations.
75 >  relative to the more computationally expensive Ewald treatment.  We
76 >  have also investigated the energy conservation properties of the new
77 >  methods in molecular dynamics simulations. The SP method shows
78 >  excellent agreement with configurational energy differences, forces,
79 >  and torques, and would be suitable for use in Monte Carlo
80 >  calculations.  Of the two new shifted-force methods, the GSF
81 >  approach shows the best agreement with Ewald-derived energies,
82 >  forces, and torques and also exhibits energy conservation properties
83 >  that make it an excellent choice for efficient computation of
84 >  electrostatic interactions in molecular dynamics simulations.
85   \end{abstract}
86  
87   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
88                               % Classification Scheme.
89 < \keywords{Electrostatics, Multipoles, Real-space}
89 > %\keywords{Electrostatics, Multipoles, Real-space}
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’s method can also be problematic for
121 < interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
122 < methods that were developed to handle two-dimensional (2D)
123 < electrostatic interactions in interfacial systems have not had similar
124 < particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
125 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
117 > interfaces require modifications to the method.  Parry's extension of
118 > the three dimensional Ewald sum is appropriate for slab
119 > geometries.\cite{Parry:1975if} Modified Ewald methods that were
120 > developed to handle two-dimensional (2-D) electrostatic
121 > interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
122 > These methods were originally quite computationally
123 > expensive.\cite{Spohr97,Yeh99} There have been several successful
124 > efforts that reduced the computational cost of 2-D lattice summations,
125 > bringing them more in line with the scaling for the full 3-D
126 > treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
127 > inherent periodicity required by the Ewald method can also be
128 > problematic in a number of protein/solvent and ionic solution
129 > environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
130  
131   \subsection{Real-space methods}
132   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
133   method for calculating electrostatic interactions between point
134 < charges. They argued that the effective Coulomb interaction in
135 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
136 < an ordered lattice (e.g., when computing the Madelung constant of an
137 < ionic solid), the material can be considered as a set of ions
138 < interacting with neutral dipolar or quadrupolar ``molecules'' giving
139 < an effective distance dependence for the electrostatic interactions of
140 < $r^{-5}$ (see figure \ref{fig:NaCl}).  For this reason, careful
141 < applications of Wolf's method are able to obtain accurate estimates of
142 < Madelung constants using relatively short cutoff radii.  Recently,
143 < Fukuda used neutralization of the higher order moments for the
144 < calculation of the electrostatic interaction of the point charges
145 < system.\cite{Fukuda:2013sf}
134 > charges. They argued that the effective Coulomb interaction in most
135 > condensed phase systems is effectively short
136 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
137 > computing the Madelung constant of an ionic solid), the material can
138 > be considered as a set of ions interacting with neutral dipolar or
139 > quadrupolar ``molecules'' giving an effective distance dependence for
140 > the electrostatic interactions of $r^{-5}$ (see figure
141 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
142 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
143 > electrostatic energy per ion converges more rapidly to the Madelung
144 > energy than the dipolar approximation.\cite{Wolf92} To find the
145 > correct Madelung constant, Lacman suggested that the NaCl structure
146 > could be constructed in a way that the finite crystal terminates with
147 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
148 > what is effectively a set of octupoles at large distances. These facts
149 > suggest that the Madelung constants are relatively short ranged for
150 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
151 > application of Wolf's method can provide accurate estimates of
152 > Madelung constants using relatively short cutoff radii.
153  
154 < \begin{figure}[h!]
154 > Direct truncation of interactions at a cutoff radius creates numerical
155 > errors.  Wolf \textit{et al.} suggest that truncation errors are due
156 > to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
157 > neutralize this charge they proposed placing an image charge on the
158 > surface of the cutoff sphere for every real charge inside the cutoff.
159 > These charges are present for the evaluation of both the pair
160 > interaction energy and the force, although the force expression
161 > maintains a discontinuity at the cutoff sphere.  In the original Wolf
162 > formulation, the total energy for the charge and image were not equal
163 > to the integral of the force expression, and as a result, the total
164 > energy would not be conserved in molecular dynamics (MD)
165 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
166 > Gezelter later proposed shifted force variants of the Wolf method with
167 > commensurate force and energy expressions that do not exhibit this
168 > problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
169 > were also proposed by Chen \textit{et
170 >  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
171 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
172 > used additional neutralization of higher order moments for systems of
173 > point charges.\cite{Fukuda:2013sf}
174 >
175 > \begin{figure}
176    \centering
177 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
178 <  \caption{Top: NaCl crystal showing how spherical truncation can
179 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
180 <    molecules interact with the central ion.  Bottom: A dipolar
181 <    crystal exhibiting similar behavior and illustrating how the
182 <    effective dipole-octupole interactions can be disrupted by
183 <    spherical truncation.}
184 <  \label{fig:NaCl}
177 >  \includegraphics[width=\linewidth]{schematic.eps}
178 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
179 >    charges (in the smaller grey circle), so interactions are
180 >    effectively charge-multipole at longer distances.  With hard
181 >    cutoffs, motion of individual charges in and out of the cutoff
182 >    sphere can break the effective multipolar ordering.  Bottom:
183 >    dipolar crystals and fluids have a similar effective
184 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
185 >    orientational averaging helps to reduce the effective range of the
186 >    interactions in the fluid.  Placement of reversed image multipoles
187 >    on the surface of the cutoff sphere recovers the effective
188 >    higher-order multipole behavior.}
189 >  \label{fig:schematic}
190   \end{figure}
191  
192 < The direct truncation of interactions at a cutoff radius creates
193 < truncation defects. Wolf \textit{et al.} further argued that
194 < truncation errors are due to net charge remaining inside the cutoff
195 < sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
196 < placing an image charge on the surface of the cutoff sphere for every
197 < real charge inside the cutoff.  These charges are present for the
198 < evaluation of both the pair interaction energy and the force, although
199 < the force expression maintained a discontinuity at the cutoff sphere.
200 < In the original Wolf formulation, the total energy for the charge and
201 < image were not equal to the integral of their force expression, and as
170 < a result, the total energy would not be conserved in molecular
171 < dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
172 < Fennel and Gezelter later proposed shifted force variants of the Wolf
173 < method with commensurate force and energy expressions that do not
174 < exhibit this problem.\cite{Fennell:2006lq}   Related real-space
175 < methods were also proposed by Chen \textit{et
176 <  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
177 < and by Wu and Brooks.\cite{Wu:044107}
178 <
179 < Considering the interaction of one central ion in an ionic crystal
180 < with a portion of the crystal at some distance, the effective Columbic
181 < potential is found to be decreasing as $r^{-5}$. If one views the
182 < \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
183 < \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
184 < rapidly to the Madelung energy than the dipolar
185 < approximation.\cite{Wolf92} To find the correct Madelung constant,
186 < Lacman suggested that the NaCl structure could be constructed in a way
187 < that the finite crystal terminates with complete \ce{(NaCl)4}
188 < molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
189 < by opposite charges. Similarly for each pair of charges, there is an
190 < opposite pair of charge adjacent to it.  The central ion sees what is
191 < effectively a set of octupoles at large distances. These facts suggest
192 < that the Madelung constants are relatively short ranged for perfect
193 < ionic crystals.\cite{Wolf:1999dn}
194 <
195 < One can make a similar argument for crystals of point multipoles. The
196 < Luttinger and Tisza treatment of energy constants for dipolar lattices
197 < utilizes 24 basis vectors that contain dipoles at the eight corners of
198 < a unit cube.  Only three of these basis vectors, $X_1, Y_1,
199 < \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
200 < zero net dipole and retain contributions only from higher order
201 < 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 208 | Line 208 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
208   unstable.
209  
210   In ionic crystals, real-space truncation can break the effective
211 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
212 < swings in the electrostatic energy as individual ions move back and
213 < forth across the boundary.  This is why the image charges are
211 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
212 > significant swings in the electrostatic energy as individual ions move
213 > back and forth across the boundary.  This is why the image charges are
214   necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
215   the real-space truncation of point multipole interactions breaks
216   higher order multipole arrangements, and image multipoles are required
217   for real-space treatments of electrostatic energies.
218  
219 + The shorter effective range of electrostatic interactions is not
220 + limited to perfect crystals, but can also apply in disordered fluids.
221 + Even at elevated temperatures, there is local charge balance in an
222 + ionic liquid, where each positive ion has surroundings dominated by
223 + negaitve ions and vice versa.  The reversed-charge images on the
224 + cutoff sphere that are integral to the Wolf and DSF approaches retain
225 + the effective multipolar interactions as the charges traverse the
226 + cutoff boundary.
227 +
228 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
229 + significant orientational averaging that additionally reduces the
230 + effect of long-range multipolar interactions.  The image multipoles
231 + that are introduced in the TSF, GSF, and SP methods mimic this effect
232 + and reduce the effective range of the multipolar interactions as
233 + interacting molecules traverse each other's cutoff boundaries.
234 +
235   % Because of this reason, although the nature of electrostatic
236   % interaction short ranged, the hard cutoff sphere creates very large
237   % fluctuation in the electrostatic energy for the perfect crystal. In
# Line 226 | 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
239 < 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
265 < negative charges, and rather than use multiple site-site interactions,
266 < the interaction between higher order multipoles can also be used to
267 < evaluate a single molecule-molecule
268 < 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
276 < 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 315 | Line 325 | U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1
325   expressed as the product of two multipole operators and a Coulombic
326   kernel,
327   \begin{equation}
328 < U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
328 > U_{ab}(r)= M_{a} M_{b} \frac{1}{r}  \label{kernel}.
329   \end{equation}
330 < where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
331 < expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
332 <    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
323 < $\bf a$.
330 > where the multipole operator for site $a$, $M_{a}$, is
331 > expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
332 > $a$, etc.
333  
334   % Interactions between multipoles can be expressed as higher derivatives
335   % of the bare Coulomb potential, so one way of ensuring that the forces
# Line 348 | Line 357 | of the interaction, with $n=0$ for charge-charge, $n=1
357   \label{generic}
358   \end{equation}
359   where $f_n(r)$ is a shifted kernel that is appropriate for the order
360 < of the interaction, with $n=0$ for charge-charge, $n=1$ for
361 < charge-dipole, $n=2$ for charge-quadrupole and dipole-dipole, $n=3$
362 < for dipole-quadrupole, and $n=4$ for quadrupole-quadrupole.  To ensure
363 < smooth convergence of the energy, force, and torques, a Taylor
364 < expansion with $n$ terms must be performed at cutoff radius ($r_c$) to
365 < obtain $f_n(r)$.
360 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
361 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
362 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
363 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
364 > force, and torques, a Taylor expansion with $n$ terms must be
365 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
366  
367   % To carry out the same procedure for a damped electrostatic kernel, we
368   % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
# Line 375 | Line 384 | radius.  For example, the direct dipole dot product
384   contributions to the potential, and ensures that the forces and
385   torques from each of these contributions will vanish at the cutoff
386   radius.  For example, the direct dipole dot product
387 < ($\mathbf{D}_{\bf a}
388 < \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
387 > ($\mathbf{D}_{a}
388 > \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
389   dot products:
390   \begin{equation}
391 < U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
392 <  \mathbf{D}_{\bf a} \cdot
393 < \mathbf{D}_{\bf b} \right) v_{21}(r) +
394 < \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
395 < \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
391 > U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
392 >  \mathbf{D}_{a} \cdot
393 > \mathbf{D}_{b} \right) v_{21}(r) +
394 > \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
395 > \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
396   \end{equation}
397  
398   For the Taylor shifted (TSF) method with the undamped kernel,
# Line 393 | Line 402 | to another site within cutoff sphere are derived from
402   connection to unmodified electrostatics as well as the smooth
403   transition to zero in both these functions as $r\rightarrow r_c$.  The
404   electrostatic forces and torques acting on the central multipole due
405 < to another site within cutoff sphere are derived from
405 > to another site within the cutoff sphere are derived from
406   Eq.~\ref{generic}, accounting for the appropriate number of
407   derivatives. Complete energy, force, and torque expressions are
408   presented in the first paper in this series (Reference
# Line 407 | Line 416 | without changing their relative orientation,
416   shifted smoothly by finding the gradient for two interacting dipoles
417   which have been projected onto the surface of the cutoff sphere
418   without changing their relative orientation,
419 < \begin{displaymath}
420 < U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
421 < U_{D_{\bf a} D_{\bf b}}(r_c)
422 <   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
423 <  \vec{\nabla} U_{D_{\bf a}D_{\bf b}}(r) \Big \lvert _{r_c}
424 < \end{displaymath}
425 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
417 <  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
419 > \begin{equation}
420 > U_{D_{a}D_{b}}(r)  = U_{D_{a}D_{b}}(r) -
421 > U_{D_{a}D_{b}}(r_c)
422 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
423 >  \nabla U_{D_{a}D_{b}}(r_c).
424 > \end{equation}
425 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
426   (although the signs are reversed for the dipole that has been
427   projected onto the cutoff sphere).  In many ways, this simpler
428   approach is closer in spirit to the original shifted force method, in
# Line 435 | Line 443 | U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathb
443   In general, the gradient shifted potential between a central multipole
444   and any multipolar site inside the cutoff radius is given by,
445   \begin{equation}
446 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
447 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
448 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
446 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
447 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
448 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
449   \label{generic2}
450   \end{equation}
451   where the sum describes a separate force-shifting that is applied to
452 < each orientational contribution to the energy.
452 > each orientational contribution to the energy.  In this expression,
453 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
454 > ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
455 > represent the orientations the multipoles.
456  
457   The third term converges more rapidly than the first two terms as a
458   function of radius, hence the contribution of the third term is very
459   small for large cutoff radii.  The force and torque derived from
460 < equation \ref{generic2} are consistent with the energy expression and
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
# Line 454 | Line 465 | GSF potential are presented in the first paper in this
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~\onlinecite{PaperI})
468 > (Reference~\onlinecite{PaperI}).
469  
470  
471   \subsection{Shifted potential (SP) }
# Line 467 | Line 478 | U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf
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^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
482 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
481 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
482 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
483   \label{eq:SP}
484   \end{equation}          
485   where the sum describes separate potential shifting that is done for
# Line 561 | Line 572 | the simulation proceeds. These differences are the mos
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
575 > the simulation proceeds. These differences are the most important
576   indicators of the reliability of a method even if the absolute
577   energies are not exact.  For each of the multipolar systems listed
578   above, we have compared the change in electrostatic potential energy
# Line 573 | Line 584 | program, OpenMD,\cite{openmd} which was used for all c
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
587 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
588   this work.  The complementary error function can be a relatively slow
589   function on some processors, so all of the radial functions are
590   precomputed on a fine grid and are spline-interpolated to provide
# Line 780 | Line 791 | model must allow for long simulation times with minima
791  
792   \begin{figure}
793    \centering
794 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
794 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
795    \caption{Statistical analysis of the quality of configurational
796      energy differences for the real-space electrostatic methods
797      compared with the reference Ewald sum.  Results with a value equal
# Line 853 | Line 864 | forces is desired.
864  
865   \begin{figure}
866    \centering
867 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
867 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
868    \caption{Statistical analysis of the quality of the force vector
869      magnitudes for the real-space electrostatic methods compared with
870      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 867 | Line 878 | forces is desired.
878  
879   \begin{figure}
880    \centering
881 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
881 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
882    \caption{Statistical analysis of the quality of the torque vector
883      magnitudes for the real-space electrostatic methods compared with
884      the reference Ewald sum. Results with a value equal to 1 (dashed
# Line 925 | Line 936 | systematically improved by varying $\alpha$ and $r_c$.
936  
937   \begin{figure}
938    \centering
939 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
939 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
940    \caption{The circular variance of the direction of the force and
941      torque vectors obtained from the real-space methods around the
942      reference Ewald vectors. A variance equal to 0 (dashed line)
# Line 945 | Line 956 | temperature of 300K.  After equilibration, this liquid
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.
959 > temperature of 300K.  After equilibration in the canonical (NVT)
960 > ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
961 > was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
962 > Hard, SP, GSF, and TSF methods with a cutoff radius of 12\AA.  The
963 > value of the damping coefficient was also varied from the undamped
964 > case ($\alpha = 0$) to a heavily damped case ($\alpha = 0.3$
965 > \AA$^{-1}$) for all of the real space methods.  A sample was also run
966 > using the multipolar Ewald sum with the same real-space cutoff.
967  
968   In figure~\ref{fig:energyDrift} we show the both the linear drift in
969   energy over time, $\delta E_1$, and the standard deviation of energy
970   fluctuations around this drift $\delta E_0$.  Both of the
971   shifted-force methods (GSF and TSF) provide excellent energy
972 < conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
972 > conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
973   while the hard cutoff is essentially unusable for molecular dynamics.
974   SP provides some benefit over the hard cutoff because the energetic
975   jumps that happen as particles leave and enter the cutoff sphere are
# Line 972 | Line 984 | $k$-space cutoff values.
984  
985   \begin{figure}
986    \centering
987 <  \includegraphics[width=\textwidth]{newDrift_12.pdf}
987 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
988   \label{fig:energyDrift}        
989   \caption{Analysis of the energy conservation of the real-space
990 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
991 <  energy over time (in kcal / mol / particle / ns) and $\delta
992 <  \mathrm{E}_0$ is the standard deviation of energy fluctuations
993 <  around this drift (in kcal / mol / particle).  All simulations were
994 <  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
990 >  methods. $\delta \mathrm{E}_1$ is the linear drift in energy over
991 >  time (in kcal / mol / particle / ns) and $\delta \mathrm{E}_0$ is
992 >  the standard deviation of energy fluctuations around this drift (in
993 >  kcal / mol / particle).  Points that appear below the dashed grey
994 >  (Ewald) lines exhibit better energy conservation than commonly-used
995 >  parameters for Ewald-based electrostatics.  All simulations were of
996 >  a 2000-molecule simulation of SSDQ water with 48 ionic charges at
997    300 K starting from the same initial configuration. All runs
998    utilized the same real-space cutoff, $r_c = 12$\AA.}
999   \end{figure}
1000  
1001 + \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
1002 + The most important test of the modified interaction potentials is the
1003 + fidelity with which they can reproduce structural features and
1004 + dynamical properties in a liquid.  One commonly-utilized measure of
1005 + structural ordering is the pair distribution function, $g(r)$, which
1006 + measures local density deviations in relation to the bulk density.  In
1007 + the electrostatic approaches studied here, the short-range repulsion
1008 + from the Lennard-Jones potential is identical for the various
1009 + electrostatic methods, and since short range repulsion determines much
1010 + of the local liquid ordering, one would not expect to see many
1011 + differences in $g(r)$.  Indeed, the pair distributions are essentially
1012 + identical for all of the electrostatic methods studied (for each of
1013 + the different systems under investigation).  An example of this
1014 + agreement for the SSDQ water/ion system is shown in
1015 + Fig. \ref{fig:gofr}.
1016  
1017 + \begin{figure}
1018 +  \centering
1019 +  \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
1020 + \label{fig:gofr}        
1021 + \caption{The pair distribution functions, $g(r)$, for the SSDQ
1022 +  water/ion system obtained using the different real-space methods are
1023 +  essentially identical with the result from the Ewald
1024 +  treatment.}
1025 + \end{figure}
1026 +
1027 + There is a very slight overstructuring of the first solvation shell
1028 + when using when using TSF at lower values of the damping coefficient
1029 + ($\alpha \le 0.1$) or when using undamped GSF.  With moderate damping,
1030 + GSF and SP produce pair distributions that are identical (within
1031 + numerical noise) to their Ewald counterparts.
1032 +
1033 + A structural property that is a more demanding test of modified
1034 + electrostatics is the mean value of the electrostatic energy $\langle
1035 + U_\mathrm{elect} \rangle / N$ which is obtained by sampling the
1036 + liquid-state configurations experienced by a liquid evolving entirely
1037 + under the influence of each of the methods.  In table \ref{tab:Props}
1038 + we demonstrate how $\langle U_\mathrm{elect} \rangle / N$ varies with
1039 + the damping parameter, $\alpha$, for each of the methods.
1040 +
1041 + As in the crystals studied in the first paper, damping is important
1042 + for converging the mean electrostatic energy values, particularly for
1043 + the two shifted force methods (GSF and TSF).  A value of $\alpha
1044 + \approx 0.2$ \AA$^{-1}$ is sufficient to converge the SP and GSF
1045 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
1046 + dramatic damping ($\alpha \approx 0.3$ \AA$^{-1}$ for $r_c = 9$ \AA).
1047 + Overdamping the real-space electrostatic methods occurs with $\alpha >
1048 + 0.4$, causing the estimate of the energy to drop below the Ewald
1049 + results.
1050 +
1051 + These ``optimal'' values of the damping coefficient are slightly
1052 + larger than what were observed for DSF electrostatics for purely
1053 + point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1054 + $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1055 + multipole systems.
1056 +
1057 + To test the fidelity of the electrostatic methods at reproducing
1058 + dynamics in a multipolar liquid, it is also useful to look at
1059 + transport properties, particularly the diffusion constant,
1060 + \begin{equation}
1061 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1062 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1063 + \label{eq:diff}
1064 + \end{equation}
1065 + which measures long-time behavior and is sensitive to the forces on
1066 + the multipoles.  For the soft dipolar fluid and the SSDQ liquid
1067 + systems, the self-diffusion constants (D) were calculated from linear
1068 + fits to the long-time portion of the mean square displacement,
1069 + $\langle r^{2}(t) \rangle$.\cite{Allen87}
1070 +
1071 + In addition to translational diffusion, orientational relaxation times
1072 + were calculated for comparisons with the Ewald simulations and with
1073 + experiments. These values were determined from the same 1~ns
1074 + microcanonical trajectories used for translational diffusion by
1075 + calculating the orientational time correlation function,
1076 + \begin{equation}
1077 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1078 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1079 + \label{eq:OrientCorr}
1080 + \end{equation}
1081 + where $P_l$ is the Legendre polynomial of order $l$ and
1082 + $\hat{\mathbf{A}}_\gamma$ is the space-frame unit vector for body axis
1083 + $\gamma$ on a molecule..  Th body-fixed reference frame used for our
1084 + models has the $z$-axis running along the dipoles, and for the SSDQ
1085 + water model, the $y$-axis connects the two implied hydrogen atom
1086 + positions.  From the orientation autocorrelation functions, we can
1087 + obtain time constants for rotational relaxation either by fitting an
1088 + exponential function or by integrating the entire correlation
1089 + function.  In a good water model, these decay times would be
1090 + comparable to water orientational relaxation times from nuclear
1091 + magnetic resonance (NMR). The relaxation constant obtained from
1092 + $C_2^y(t)$ is normally of experimental interest because it describes
1093 + the relaxation of the principle axis connecting the hydrogen
1094 + atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1095 + of the dipole-dipole relaxation from a proton NMR signal and should
1096 + provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1097 +
1098 + Results for the diffusion constants and orientational relaxation times
1099 + are shown in figure \ref{tab:Props}. From this data, it is apparent
1100 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1101 + reproduced with reasonable fidelity by the GSF method.
1102 +
1103 + The $\tau_2$ results in \ref{tab:Props} show a much greater difference
1104 + between the real-space and the Ewald results.
1105 +
1106 + \begin{table}
1107 + \label{tab:Props}
1108 + \caption{Comparison of the structural and dynamic properties for the
1109 +  soft dipolar liquid test for all of the real-space methods.}
1110 + \begin{tabular}{l|c|cccc|cccc|cccc}
1111 +         & Ewald & \multicolumn{4}{c|}{SP} & \multicolumn{4}{c|}{GSF} & \multicolumn{4}{c|}{TSF} \\
1112 + $\alpha$ (\AA$^{-1}$) & &      
1113 + 0.0 & 0.1 & 0.2 & 0.3 &
1114 + 0.0 & 0.1 & 0.2 & 0.3 &
1115 + 0.0 & 0.1 & 0.2 & 0.3 \\ \cline{2-6}\cline{6-10}\cline{10-14}
1116 +
1117 + $\langle U_\mathrm{elect} \rangle /N$ &&&&&&&&&&&&&\\
1118 + D ($10^{-4}~\mathrm{cm}^2/\mathrm{s}$)&
1119 + 470.2(6) &
1120 + 416.6(5) &
1121 + 379.6(5) &
1122 + 438.6(5) &
1123 + 476.0(6) &
1124 + 412.8(5) &
1125 + 421.1(5) &
1126 + 400.5(5) &
1127 + 437.5(6) &
1128 + 434.6(5) &
1129 + 411.4(5) &
1130 + 545.3(7) &
1131 + 459.6(6) \\
1132 + $\tau_2$ (fs) &
1133 + 1.136 &
1134 + 1.041 &
1135 + 1.064 &
1136 + 1.109 &
1137 + 1.211 &
1138 + 1.119 &
1139 + 1.039 &
1140 + 1.058 &
1141 + 1.21  &
1142 + 1.15  &
1143 + 1.172 &
1144 + 1.153 &
1145 + 1.125 \\
1146 + \end{tabular}
1147 + \end{table}
1148 +
1149 +
1150   \section{CONCLUSION}
1151   In the first paper in this series, we generalized the
1152   charge-neutralized electrostatic energy originally developed by Wolf

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines