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 4184 by gezelter, Sat Jun 14 23:35:39 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 70 | Line 70 | of Notre Dame, Notre Dame, IN 46556}
70    We have tested the real-space shifted potential (SP),
71    gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
72    for multipoles that were developed in the first paper in this series
73 <  against a reference method. The tests were carried out in a variety
74 <  of condensed-phase environments which were designed to test all
75 <  levels of the multipole-multipole interactions.  Comparisons of the
76 <  energy differences between configurations, molecular forces, and
77 <  torques were used to analyze how well the real-space models perform
78 <  relative to the more computationally expensive Ewald sum.  We have
79 <  also investigated the energy conservation properties of the new
80 <  methods in molecular dynamics simulations using all of these
81 <  methods. The SP method shows excellent agreement with
82 <  configurational energy differences, forces, and torques, and would
83 <  be suitable for use in Monte Carlo calculations.  Of the two new
84 <  shifted-force methods, the GSF approach shows the best agreement
85 <  with Ewald-derived energies, forces, and torques and exhibits energy
86 <  conservation properties that make it an excellent choice for
87 <  efficiently computing electrostatic interactions in molecular
88 <  dynamics simulations.
73 >  against the multipolar Ewald sum as a reference method. The tests
74 >  were carried out in a variety of condensed-phase environments which
75 >  were designed to test all levels of the multipole-multipole
76 >  interactions.  Comparisons of the energy differences between
77 >  configurations, molecular forces, and torques were used to analyze
78 >  how well the real-space models perform relative to the more
79 >  computationally expensive Ewald treatment.  We have also investigated the
80 >  energy conservation properties of the new methods in molecular
81 >  dynamics simulations using all of these methods. The SP method shows
82 >  excellent agreement with configurational energy differences, forces,
83 >  and torques, and would be suitable for use in Monte Carlo
84 >  calculations.  Of the two new shifted-force methods, the GSF
85 >  approach shows the best agreement with Ewald-derived energies,
86 >  forces, and torques and exhibits energy conservation properties that
87 >  make it an excellent choice for efficiently computing electrostatic
88 >  interactions in molecular dynamics simulations.
89   \end{abstract}
90  
91   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92                               % Classification Scheme.
93 < \keywords{Electrostatics, Multipoles, Real-space}
93 > %\keywords{Electrostatics, Multipoles, Real-space}
94  
95   \maketitle
96  
# Line 122 | Line 122 | periodicity in the Ewald’s method can also be proble
122   interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
123   To simulate interfacial systems, Parry's extension of the 3D Ewald sum
124   is appropriate for slab geometries.\cite{Parry:1975if} The inherent
125 < periodicity in the Ewald’s method can also be problematic for
126 < interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
127 < methods that were developed to handle two-dimensional (2D)
128 < electrostatic interactions in interfacial systems have not had similar
129 < particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
125 > periodicity in the Ewald method can also be problematic for molecular
126 > interfaces.\cite{Fennell:2006lq} Modified Ewald methods that were
127 > developed to handle two-dimensional (2D) electrostatic interactions in
128 > interfacial systems have not seen similar particle-mesh
129 > treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 >  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
131 > with system size.
132  
133   \subsection{Real-space methods}
134   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
135   method for calculating electrostatic interactions between point
136   charges. They argued that the effective Coulomb interaction in
137 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
138 < an ordered lattice (e.g., when computing the Madelung constant of an
139 < ionic solid), the material can be considered as a set of ions
137 > condensed phase systems is actually short ranged.\cite{Wolf92,Wolf95}
138 > For an ordered lattice (e.g., when computing the Madelung constant of
139 > an ionic solid), the material can be considered as a set of ions
140   interacting with neutral dipolar or quadrupolar ``molecules'' giving
141   an effective distance dependence for the electrostatic interactions of
142 < $r^{-5}$ (see figure \ref{fig:NaCl}).  For this reason, careful
143 < applications of Wolf's method are able to obtain accurate estimates of
144 < Madelung constants using relatively short cutoff radii.  Recently,
145 < Fukuda used neutralization of the higher order moments for the
146 < calculation of the electrostatic interaction of the point charges
147 < system.\cite{Fukuda:2013sf}
142 > $r^{-5}$ (see figure \ref{fig:schematic}).  For this reason, careful
143 > application of Wolf's method can obtain accurate estimates of Madelung
144 > constants using relatively short cutoff radii.  Recently, Fukuda used
145 > neutralization of the higher order moments for calculation of the
146 > electrostatic interactions in point charge
147 > systems.\cite{Fukuda:2013sf}
148  
149 < \begin{figure}[h!]
149 > \begin{figure}
150    \centering
151 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
152 <  \caption{Top: NaCl crystal showing how spherical truncation can
153 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
154 <    molecules interact with the central ion.  Bottom: A dipolar
155 <    crystal exhibiting similar behavior and illustrating how the
156 <    effective dipole-octupole interactions can be disrupted by
157 <    spherical truncation.}
158 <  \label{fig:NaCl}
151 >  \includegraphics[width=\linewidth]{schematic.pdf}
152 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
153 >    charges (in the smaller grey circle), so interactions are
154 >    effectively charge-multipole at longer distances.  With hard
155 >    cutoffs, motion of individual charges in and out of the cutoff
156 >    sphere can break the effective multipolar ordering.  Bottom:
157 >    dipolar crystals and fluids have a similar effective
158 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
159 >    orientational averaging helps to reduce the effective range of the
160 >    interactions in the fluid.  Placement of reversed image multipoles
161 >    on the surface of the cutoff sphere recovers the effective
162 >    higher-order multipole behavior.}
163 >  \label{fig:schematic}
164   \end{figure}
165  
166   The direct truncation of interactions at a cutoff radius creates
167 < truncation defects. Wolf \textit{et al.} further argued that
167 > truncation defects. Wolf \textit{et al.}  argued that
168   truncation errors are due to net charge remaining inside the cutoff
169   sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
170   placing an image charge on the surface of the cutoff sphere for every
# Line 178 | Line 184 | potential is found to be decreasing as $r^{-5}$. If on
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 be decreasing as $r^{-5}$. If one views the
188 < \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
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} Any charge in a NaCl crystal is surrounded
195 < by opposite charges. Similarly for each pair of charges, there is an
196 < opposite pair of charge adjacent to it.  The central ion sees what is
197 < 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 > 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
# Line 208 | Line 212 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
212   unstable.
213  
214   In ionic crystals, real-space truncation can break the effective
215 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
216 < swings in the electrostatic energy as individual ions move back and
217 < forth across the boundary.  This is why the image charges are
215 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
216 > significant swings in the electrostatic energy as individual ions move
217 > back and forth across the boundary.  This is why the image charges are
218   necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
219   the real-space truncation of point multipole interactions breaks
220   higher order multipole arrangements, and image multipoles are required
221   for real-space treatments of electrostatic energies.
222 +
223 + The shorter effective range of electrostatic interactions is not
224 + limited to perfect crystals, but can also apply in disordered fluids.
225 + Even at elevated temperatures, there is, on average, local charge
226 + balance in an ionic liquid, where each positive ion has surroundings
227 + dominated by negaitve ions and vice versa.  The reversed-charge images
228 + on the cutoff sphere that are integral to the Wolf and DSF approaches
229 + retain the effective multipolar interactions as the charges traverse
230 + the cutoff boundary.
231 +
232 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
233 + significant orientational averaging that additionally reduces the
234 + effect of long-range multipolar interactions.  The image multipoles
235 + that are introduced in the TSF, GSF, and SP methods mimic this effect
236 + and reduce the effective range of the multipolar interactions as
237 + interacting molecules traverse each other's cutoff boundaries.
238  
239   % Because of this reason, although the nature of electrostatic
240   % interaction short ranged, the hard cutoff sphere creates very large
# Line 320 | Line 340 | $\bf a$.
340   where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
341   expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
342      a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
343 < $\bf a$.
343 > $\bf a$, etc.
344  
345   % Interactions between multipoles can be expressed as higher derivatives
346   % of the bare Coulomb potential, so one way of ensuring that the forces
# Line 348 | Line 368 | of the interaction, with $n=0$ for charge-charge, $n=1
368   \label{generic}
369   \end{equation}
370   where $f_n(r)$ is a shifted kernel that is appropriate for the order
371 < of the interaction, with $n=0$ for charge-charge, $n=1$ for
372 < charge-dipole, $n=2$ for charge-quadrupole and dipole-dipole, $n=3$
373 < for dipole-quadrupole, and $n=4$ for quadrupole-quadrupole.  To ensure
374 < smooth convergence of the energy, force, and torques, a Taylor
375 < expansion with $n$ terms must be performed at cutoff radius ($r_c$) to
376 < obtain $f_n(r)$.
371 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
372 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
373 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
374 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
375 > force, and torques, a Taylor expansion with $n$ terms must be
376 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
377  
378   % To carry out the same procedure for a damped electrostatic kernel, we
379   % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
# Line 393 | Line 413 | to another site within cutoff sphere are derived from
413   connection to unmodified electrostatics as well as the smooth
414   transition to zero in both these functions as $r\rightarrow r_c$.  The
415   electrostatic forces and torques acting on the central multipole due
416 < to another site within cutoff sphere are derived from
416 > to another site within the cutoff sphere are derived from
417   Eq.~\ref{generic}, accounting for the appropriate number of
418   derivatives. Complete energy, force, and torque expressions are
419   presented in the first paper in this series (Reference
# Line 407 | Line 427 | without changing their relative orientation,
427   shifted smoothly by finding the gradient for two interacting dipoles
428   which have been projected onto the surface of the cutoff sphere
429   without changing their relative orientation,
430 < \begin{displaymath}
430 > \begin{equation}
431   U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
432   U_{D_{\bf a} D_{\bf b}}(r_c)
433     - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
434 <  \vec{\nabla} U_{D_{\bf a}D_{\bf b}}(r) \Big \lvert _{r_c}
435 < \end{displaymath}
434 >  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
435 > \end{equation}
436   Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
437    a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
438   (although the signs are reversed for the dipole that has been
# Line 435 | Line 455 | U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathb
455   In general, the gradient shifted potential between a central multipole
456   and any multipolar site inside the cutoff radius is given by,
457   \begin{equation}
458 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
459 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
460 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
458 >  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
459 >    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
460 >    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
461   \label{generic2}
462   \end{equation}
463   where the sum describes a separate force-shifting that is applied to
464 < each orientational contribution to the energy.
464 > each orientational contribution to the energy.  In this expression,
465 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
466 > ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
467 > represent the orientations the multipoles.
468  
469   The third term converges more rapidly than the first two terms as a
470   function of radius, hence the contribution of the third term is very
471   small for large cutoff radii.  The force and torque derived from
472 < equation \ref{generic2} are consistent with the energy expression and
472 > Eq. \ref{generic2} are consistent with the energy expression and
473   approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
474   can be considered generalizations of the original DSF method for
475   higher order multipole interactions. GSF and TSF are also identical up
# Line 454 | Line 477 | GSF potential are presented in the first paper in this
477   the energy, force and torque for higher order multipole-multipole
478   interactions. Complete energy, force, and torque expressions for the
479   GSF potential are presented in the first paper in this series
480 < (Reference~\onlinecite{PaperI})
480 > (Reference~\onlinecite{PaperI}).
481  
482  
483   \subsection{Shifted potential (SP) }
# Line 468 | Line 491 | U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \ri
491   effectively shifts the total potential to zero at the cutoff radius,
492   \begin{equation}
493   U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
494 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
494 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
495   \label{eq:SP}
496   \end{equation}          
497   where the sum describes separate potential shifting that is done for
# Line 957 | Line 980 | conservation (drift less than $10^{-6}$ kcal / mol / n
980   energy over time, $\delta E_1$, and the standard deviation of energy
981   fluctuations around this drift $\delta E_0$.  Both of the
982   shifted-force methods (GSF and TSF) provide excellent energy
983 < conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
983 > conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
984   while the hard cutoff is essentially unusable for molecular dynamics.
985   SP provides some benefit over the hard cutoff because the energetic
986   jumps that happen as particles leave and enter the cutoff sphere are

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines