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 4187 by gezelter, Sun Jun 15 16:25:42 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 alternatives to the Ewald Sum. II. Comparison of Methods}
51  
52   \author{Madan Lamichhane}
53 < \affiliation{Department of Physics, University
55 < 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
59 < 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,
67 <             %  but any date may be explicitly specified
63 > \date{\today}
64  
65   \begin{abstract}
66 <  We have tested the real-space shifted potential (SP),
67 <  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
68 <  for multipoles that were developed in the first paper in this series
69 <  against the multipolar Ewald sum as a reference method. The tests
70 <  were carried out in a variety of condensed-phase environments which
71 <  were designed to test all levels of the multipole-multipole
72 <  interactions.  Comparisons of the energy differences between
73 <  configurations, molecular forces, and torques were used to analyze
74 <  how well the real-space models perform relative to the more
75 <  computationally expensive Ewald treatment.  We have also investigated the
76 <  energy conservation properties of the new methods in molecular
81 <  dynamics simulations using all of these methods. The SP method shows
66 >  We report on tests of the shifted potential (SP), gradient shifted
67 >  force (GSF), and Taylor shifted force (TSF) real-space methods for
68 >  multipole interactions developed in the first paper in this series,
69 >  using the multipolar Ewald sum as a reference method. The tests were
70 >  carried out in a variety of condensed-phase environments designed to
71 >  test up to quadrupole-quadrupole interactions.  Comparisons of the
72 >  energy differences between configurations, molecular forces, and
73 >  torques were used to analyze how well the real-space models perform
74 >  relative to the more computationally expensive Ewald treatment.  We
75 >  have also investigated the energy conservation properties of the new
76 >  methods in molecular dynamics simulations. The SP method shows
77    excellent agreement with configurational energy differences, forces,
78    and torques, and would be suitable for use in Monte Carlo
79    calculations.  Of the two new shifted-force methods, the GSF
80    approach shows the best agreement with Ewald-derived energies,
81 <  forces, and torques and exhibits energy conservation properties that
82 <  make it an excellent choice for efficiently computing electrostatic
83 <  interactions in molecular dynamics simulations.
81 >  forces, and torques and also exhibits energy conservation properties
82 >  that make it an excellent choice for efficient computation of
83 >  electrostatic interactions in molecular dynamics simulations.
84   \end{abstract}
85  
86   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
# Line 94 | Line 89 | of Notre Dame, Notre Dame, IN 46556}
89  
90   \maketitle
91  
97
92   \section{\label{sec:intro}Introduction}
93   Computing the interactions between electrostatic sites is one of the
94 < most expensive aspects of molecular simulations, which is why there
95 < have been significant efforts to develop practical, efficient and
96 < convergent methods for handling these interactions. Ewald's method is
97 < perhaps the best known and most accurate method for evaluating
98 < energies, forces, and torques in explicitly-periodic simulation
99 < cells. In this approach, the conditionally convergent electrostatic
100 < energy is converted into two absolutely convergent contributions, one
101 < which is carried out in real space with a cutoff radius, and one in
102 < reciprocal space.\cite{Clarke:1986eu,Woodcock75}
94 > most expensive aspects of molecular simulations. There have been
95 > significant efforts to develop practical, efficient and convergent
96 > methods for handling these interactions. Ewald's method is perhaps the
97 > best known and most accurate method for evaluating energies, forces,
98 > and torques in explicitly-periodic simulation cells. In this approach,
99 > the conditionally convergent electrostatic energy is converted into
100 > two absolutely convergent contributions, one which is carried out in
101 > real space with a cutoff radius, and one in reciprocal
102 > space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
103  
104   When carried out as originally formulated, the reciprocal-space
105   portion of the Ewald sum exhibits relatively poor computational
106 < scaling, making it prohibitive for large systems. By utilizing
107 < particle meshes and three dimensional fast Fourier transforms (FFT),
108 < the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
109 < (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
110 < the computational cost from $O(N^2)$ down to $O(N \log
111 < N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
106 > scaling, making it prohibitive for large systems. By utilizing a
107 > particle mesh and three dimensional fast Fourier transforms (FFT), the
108 > particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
109 > (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
110 > methods can decrease the computational cost from $O(N^2)$ down to $O(N
111 > \log
112 > N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
113  
114 < Because of the artificial periodicity required for the Ewald sum, the
120 < method may require modification to compute interactions for
114 > Because of the artificial periodicity required for the Ewald sum,
115   interfacial molecular systems such as membranes and liquid-vapor
116 < interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
117 < To simulate interfacial systems, Parry's extension of the 3D Ewald sum
118 < is appropriate for slab geometries.\cite{Parry:1975if} The inherent
119 < periodicity in the Ewald method can also be problematic for molecular
120 < interfaces.\cite{Fennell:2006lq} Modified Ewald methods that were
121 < developed to handle two-dimensional (2D) electrostatic interactions in
122 < interfacial systems have not seen similar particle-mesh
123 < treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
124 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
125 < with system size.
116 > interfaces require modifications to the method.  Parry's extension of
117 > the three dimensional Ewald sum is appropriate for slab
118 > geometries.\cite{Parry:1975if} Modified Ewald methods that were
119 > developed to handle two-dimensional (2-D) electrostatic
120 > interactions,\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
121 > but these methods were originally quite computationally
122 > expensive.\cite{Spohr97,Yeh99} There have been several successful
123 > efforts that reduced the computational cost of 2-D lattice
124 > summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
125 > bringing them more in line with the scaling for the full 3-D
126 > treatments.  The inherent periodicity in the Ewald’s method can also
127 > be problematic for interfacial molecular
128 > systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
129  
130   \subsection{Real-space methods}
131   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
132   method for calculating electrostatic interactions between point
133 < charges. They argued that the effective Coulomb interaction in
134 < condensed phase systems is actually short ranged.\cite{Wolf92,Wolf95}
135 < For an ordered lattice (e.g., when computing the Madelung constant of
136 < an ionic solid), the material can be considered as a set of ions
137 < interacting with neutral dipolar or quadrupolar ``molecules'' giving
138 < an effective distance dependence for the electrostatic interactions of
139 < $r^{-5}$ (see figure \ref{fig:schematic}).  For this reason, careful
140 < application of Wolf's method can obtain accurate estimates of Madelung
141 < constants using relatively short cutoff radii.  Recently, Fukuda used
142 < neutralization of the higher order moments for calculation of the
143 < electrostatic interactions in point charge
144 < systems.\cite{Fukuda:2013sf}
133 > charges. They argued that the effective Coulomb interaction in most
134 > condensed phase systems is effectively short
135 > ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
136 > computing the Madelung constant of an ionic solid), the material can
137 > be considered as a set of ions interacting with neutral dipolar or
138 > quadrupolar ``molecules'' giving an effective distance dependence for
139 > the electrostatic interactions of $r^{-5}$ (see figure
140 > \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
141 > cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
142 > electrostatic energy per ion converges more rapidly to the Madelung
143 > energy than the dipolar approximation.\cite{Wolf92} To find the
144 > correct Madelung constant, Lacman suggested that the NaCl structure
145 > could be constructed in a way that the finite crystal terminates with
146 > complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
147 > what is effectively a set of octupoles at large distances. These facts
148 > suggest that the Madelung constants are relatively short ranged for
149 > perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
150 > application of Wolf's method can provide accurate estimates of
151 > Madelung constants using relatively short cutoff radii.
152  
153 + Direct truncation of interactions at a cutoff radius creates numerical
154 + errors.  Wolf \textit{et al.} suggest that truncation errors are due
155 + to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
156 + neutralize this charge they proposed placing an image charge on the
157 + surface of the cutoff sphere for every real charge inside the cutoff.
158 + These charges are present for the evaluation of both the pair
159 + interaction energy and the force, although the force expression
160 + maintains a discontinuity at the cutoff sphere.  In the original Wolf
161 + formulation, the total energy for the charge and image were not equal
162 + to the integral of the force expression, and as a result, the total
163 + energy would not be conserved in molecular dynamics (MD)
164 + simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
165 + Gezelter later proposed shifted force variants of the Wolf method with
166 + commensurate force and energy expressions that do not exhibit this
167 + problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
168 + were also proposed by Chen \textit{et
169 +  al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
170 + and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
171 + used additional neutralization of higher order moments for systems of
172 + point charges.\cite{Fukuda:2013sf}
173 +
174   \begin{figure}
175    \centering
176    \includegraphics[width=\linewidth]{schematic.pdf}
# Line 163 | Line 188 | The direct truncation of interactions at a cutoff radi
188    \label{fig:schematic}
189   \end{figure}
190  
191 < The direct truncation of interactions at a cutoff radius creates
192 < truncation defects. Wolf \textit{et al.}  argued that
193 < truncation errors are due to net charge remaining inside the cutoff
194 < sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
195 < placing an image charge on the surface of the cutoff sphere for every
196 < real charge inside the cutoff.  These charges are present for the
197 < evaluation of both the pair interaction energy and the force, although
198 < the force expression maintained a discontinuity at the cutoff sphere.
199 < In the original Wolf formulation, the total energy for the charge and
200 < 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
191 > One can make a similar effective range argument for crystals of point
192 > \textit{multipoles}. The Luttinger and Tisza treatment of energy
193 > constants for dipolar lattices utilizes 24 basis vectors that contain
194 > dipoles at the eight corners of a unit cube.\cite{LT} Only three of
195 > these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
196 > moments, while the rest have zero net dipole and retain contributions
197 > only from higher order multipoles.  The lowest-energy crystalline
198 > structures are built out of basis vectors that have only residual
199 > quadrupolar moments (e.g. the $Z_5$ array). In these low energy
200 > structures, the effective interaction between a dipole at the center
201   of a crystal and a group of eight dipoles farther away is
202   significantly shorter ranged than the $r^{-3}$ that one would expect
203   for raw dipole-dipole interactions.  Only in crystals which retain a
# Line 222 | Line 217 | Even at elevated temperatures, there is, on average, l
217  
218   The shorter effective range of electrostatic interactions is not
219   limited to perfect crystals, but can also apply in disordered fluids.
220 < Even at elevated temperatures, there is, on average, local charge
221 < balance in an ionic liquid, where each positive ion has surroundings
222 < dominated by negaitve ions and vice versa.  The reversed-charge images
223 < on the cutoff sphere that are integral to the Wolf and DSF approaches
224 < retain the effective multipolar interactions as the charges traverse
225 < the cutoff boundary.
220 > Even at elevated temperatures, there is local charge balance in an
221 > ionic liquid, where each positive ion has surroundings dominated by
222 > negaitve ions and vice versa.  The reversed-charge images on the
223 > cutoff sphere that are integral to the Wolf and DSF approaches retain
224 > the effective multipolar interactions as the charges traverse the
225 > cutoff boundary.
226  
227   In multipolar fluids (see Fig. \ref{fig:schematic}) there is
228   significant orientational averaging that additionally reduces the
# Line 246 | Line 241 | The forces and torques acting on atomic sites are the
241   % to the non-neutralized value of the higher order moments within the
242   % cutoff sphere.
243  
244 < The forces and torques acting on atomic sites are the fundamental
245 < factors driving dynamics in molecular simulations. Fennell and
246 < Gezelter proposed the damped shifted force (DSF) energy kernel to
247 < obtain consistent energies and forces on the atoms within the cutoff
248 < sphere. Both the energy and the force go smoothly to zero as an atom
249 < aproaches the cutoff radius. The comparisons of the accuracy these
250 < quantities between the DSF kernel and SPME was surprisingly
251 < good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
252 < calculating electrostatic interactions in molecular systems with
253 < relatively uniform charge
259 < densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
244 > Forces and torques acting on atomic sites are fundamental in driving
245 > dynamics in molecular simulations, and the damped shifted force (DSF)
246 > energy kernel provides consistent energies and forces on charged atoms
247 > within the cutoff sphere. Both the energy and the force go smoothly to
248 > zero as an atom aproaches the cutoff radius. The comparisons of the
249 > accuracy these quantities between the DSF kernel and SPME was
250 > surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
251 > has seen increasing use in molecular systems with relatively uniform
252 > charge
253 > densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
254  
255   \subsection{The damping function}
256 < The damping function used in our research has been discussed in detail
257 < in the first paper of this series.\cite{PaperI} The radial kernel
258 < $1/r$ for the interactions between point charges can be replaced by
259 < the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
260 < accelerate the rate of convergence, where $\alpha$ is a damping
261 < parameter with units of inverse distance.  Altering the value of
262 < $\alpha$ is equivalent to changing the width of Gaussian charge
263 < distributions that replace each point charge -- Gaussian overlap
264 < integrals yield complementary error functions when truncated at a
265 < finite distance.
256 > The damping function has been discussed in detail in the first paper
257 > of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
258 > interactions between point charges can be replaced by the
259 > complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
260 > convergence, where $\alpha$ is a damping parameter with units of
261 > inverse distance.  Altering the value of $\alpha$ is equivalent to
262 > changing the width of Gaussian charge distributions that replace each
263 > point charge, as Coulomb integrals with Gaussian charge distributions
264 > produce complementary error functions when truncated at a finite
265 > distance.
266  
267 < By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
268 < cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
269 < agreement with SPME for the interaction energies, forces and torques
270 < for charge-charge interactions.\cite{Fennell:2006lq}
267 > With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
268 > produced very good agreement with SPME for interaction energies,
269 > forces and torques for charge-charge
270 > interactions.\cite{Fennell:2006lq}
271  
272   \subsection{Point multipoles in molecular modeling}
273   Coarse-graining approaches which treat entire molecular subsystems as
274   a single rigid body are now widely used. A common feature of many
275   coarse-graining approaches is simplification of the electrostatic
276   interactions between bodies so that fewer site-site interactions are
277 < required to compute configurational energies.  Many coarse-grained
278 < 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}
277 > required to compute configurational
278 > energies.\cite{Ren06,Essex10,Essex11}
279  
280 < Because electrons in a molecule are not localized at specific points,
281 < the assignment of partial charges to atomic centers is a relatively
282 < rough approximation.  Atomic sites can also be assigned point
283 < multipoles and polarizabilities to increase the accuracy of the
284 < molecular model.  Recently, water has been modeled with point
285 < multipoles up to octupolar order using the soft sticky
296 < dipole-quadrupole-octupole (SSDQO)
280 > Additionally, because electrons in a molecule are not localized at
281 > specific points, the assignment of partial charges to atomic centers
282 > is always an approximation.  For increased accuracy, atomic sites can
283 > also be assigned point multipoles and polarizabilities.  Recently,
284 > water has been modeled with point multipoles up to octupolar order
285 > using the soft sticky dipole-quadrupole-octupole (SSDQO)
286   model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
287   multipoles up to quadrupolar order have also been coupled with point
288   polarizabilities in the high-quality AMOEBA and iAMOEBA water
289 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
290 < using point multipole with the real space truncation without
291 < accounting for multipolar neutrality will create energy conservation
292 < issues in molecular dynamics (MD) simulations.
289 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
290 > truncating point multipoles without smoothing the forces and torques
291 > can create energy conservation issues in molecular dynamics
292 > simulations.
293  
294   In this paper we test a set of real-space methods that were developed
295   for point multipolar interactions.  These methods extend the damped
296   shifted force (DSF) and Wolf methods originally developed for
297   charge-charge interactions and generalize them for higher order
298 < multipoles. The detailed mathematical development of these methods has
299 < been presented in the first paper in this series, while this work
300 < covers the testing the energies, forces, torques, and energy
298 > multipoles.  The detailed mathematical development of these methods
299 > has been presented in the first paper in this series, while this work
300 > covers the testing of energies, forces, torques, and energy
301   conservation properties of the methods in realistic simulation
302   environments.  In all cases, the methods are compared with the
303 < reference method, a full multipolar Ewald treatment.
303 > reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
304  
305  
306   %\subsection{Conservation of total energy }
# Line 596 | Line 585 | program, OpenMD,\cite{openmd} which was used for all c
585   \subsection{Implementation}
586   The real-space methods developed in the first paper in this series
587   have been implemented in our group's open source molecular simulation
588 < program, OpenMD,\cite{openmd} which was used for all calculations in
588 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
589   this work.  The complementary error function can be a relatively slow
590   function on some processors, so all of the radial functions are
591   precomputed on a fine grid and are spline-interpolated to provide

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines