ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
(Generate patch)

Comparing trunk/multipole/multipole1.tex (file contents):
Revision 3980 by gezelter, Fri Dec 6 18:50:14 2013 UTC vs.
Revision 3982 by gezelter, Fri Dec 27 17:41:17 2013 UTC

# Line 31 | Line 31 | preprint,%
31   \usepackage{graphicx}% Include figure files
32   \usepackage{dcolumn}% Align table columns on decimal point
33   %\usepackage{bm}% bold math
34 + \usepackage{times}
35   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
36   \usepackage{url}
37  
# Line 82 | Line 83 | of Notre Dame, Notre Dame, IN 46556}
83   \maketitle
84  
85   \section{Introduction}
86 <
87 < There is increasing interest in real-space methods for calculating
88 < electrostatic interactions in computer simulations of condensed
88 < molecular
86 > There has been increasing interest in real-space methods for
87 > calculating electrostatic interactions in computer simulations of
88 > condensed molecular
89   systems.\cite{Wolf99,Zahn02,Kast03,BeckD.A.C._bi0486381,Ma05,Fennell:2006zl,Chen:2004du,Chen:2006ii,Rodgers:2006nw,Denesyuk:2008ez,Izvekov:2008wo}
90   The simplest of these techniques was developed by Wolf {\it et al.}
91   in their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
92 < Fennell and Gezelter showed that a simple damped shifted force (DSF)
93 < modification to Wolf's method could give nearly quantitative agreement
94 < with smooth particle mesh Ewald (SPME)\cite{Essmann95} for quantities
95 < of interest in Monte Carlo (i.e., configurational energy differences)
96 < and Molecular Dynamics (i.e., atomic force and molecular torque
97 < vectors).\cite{Fennell:2006zl}
92 > For systems of point charges, Fennell and Gezelter showed that a
93 > simple damped shifted force (DSF) modification to Wolf's method could
94 > give nearly quantitative agreement with smooth particle mesh Ewald
95 > (SPME)\cite{Essmann95} configurational energy differences as well as
96 > atomic force and molecular torque vectors.\cite{Fennell:2006zl}
97  
98   The computational efficiency and the accuracy of the DSF method are
99   surprisingly good, particularly for systems with uniform charge
# Line 112 | Line 111 | One common feature of many coarse-graining approaches,
111   compromise between the computational speed of real-space cutoffs and
112   the accuracy of fully-periodic Ewald treatments.
113  
114 + \subsection{Coarse Graining using Point Multipoles}
115   One common feature of many coarse-graining approaches, which treat
116   entire molecular subsystems as a single rigid body, is simplification
117   of the electrostatic interactions between these bodies so that fewer
# Line 154 | Line 154 | kernel to point multipoles.  We present in this paper
154   a different orientation can cause energy discontinuities.
155  
156   This paper outlines an extension of the original DSF electrostatic
157 < kernel to point multipoles.  We present in this paper two distinct
158 < real-space interaction models for higher-order multipoles based on two
159 < truncated Taylor expansions that are carried out at the cutoff radius.
160 < We are calling these models {\bf Taylor-shifted} and {\bf
161 <  Gradient-shifted} electrostatics.  Because of differences in the
162 < initial assumptions, the two methods yield related, but different
163 < expressions for energies, forces, and torques.
157 > kernel to point multipoles.  We have developed two distinct real-space
158 > interaction models for higher-order multipoles based on two truncated
159 > Taylor expansions that are carried out at the cutoff radius.  We are
160 > calling these models {\bf Taylor-shifted} and {\bf Gradient-shifted}
161 > electrostatics.  Because of differences in the initial assumptions,
162 > the two methods yield related, but different expressions for energies,
163 > forces, and torques.  
164  
165 + In this paper we outline the new methodology and give functional forms
166 + for the energies, forces, and torques up to quadrupole-quadrupole
167 + order.  We also compare the new methods to analytic energy constants
168 + for periodic arrays of point multipoles.  In the following paper, we
169 + provide extensive numerical comparisons to Ewald-based electrostatics
170 + in common simulation enviornments.  
171 +
172   \section{Methodology}
173  
174   \subsection{Self-neutralization, damping, and force-shifting}
168
175   The DSF and Wolf methods operate by neutralizing the total charge
176   contained within the cutoff sphere surrounding each particle.  This is
177   accomplished by shifting the potential functions to generate image
178   charges on the surface of the cutoff sphere for each pair interaction
179 < computed within $R_\textrm{c}$. An Ewald-like complementary error
180 < function damping is applied to the potential to accelerate
181 < convergence.  The potential for the DSF method, where $\alpha$ is the
182 < adjustable damping parameter, is given by
179 > computed within $R_\textrm{c}$. Damping using a complementary error
180 > function is applied to the potential to accelerate convergence. The
181 > potential for the DSF method, where $\alpha$ is the adjustable damping
182 > parameter, is given by
183   \begin{equation*}
184   V_\mathrm{DSF}(r) = C_a C_b \Biggr{[}
185   \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
# Line 195 | Line 201 | damping.\cite{deLeeuw80,Wolf99}
201   utilization of the complimentary error function for electrostatic
202   damping.\cite{deLeeuw80,Wolf99}
203  
204 + There have been recent efforts to extend the Wolf self-neutralization
205 + method to zero out the dipole and higher order multipoles contained
206 + within the cutoff
207 + sphere.\cite{Fukuda:2011jk,Fukuda:2012yu,Fukuda:2013qv}  
208 +
209 + In this work, we will be extending the idea of self-neutralization for
210 + the point multipoles in a similar way.  In Figure
211 + \ref{fig:shiftedMultipoles}, the central dipolar site $\mathbf{D}_i$
212 + is interacting with point dipole $\mathbf{D}_j$ and point quadrupole,
213 + $\mathbf{Q}_k$.  The self-neutralization scheme for point multipoles
214 + involves projecting opposing multipoles for sites $j$ and $k$ on the
215 + surface of the cutoff sphere.
216 +
217   \begin{figure}
218 < \includegraphics[width=\linewidth]{SM}
218 > \includegraphics[width=3in]{SM}
219   \caption{Reversed multipoles are projected onto the surface of the
220    cutoff sphere. The forces, torques, and potential are then smoothly
221    shifted to zero as the sites leave the cutoff region.}
222   \label{fig:shiftedMultipoles}
223   \end{figure}
224  
225 < \subsection{Multipole Expansion}
225 > As in the point-charge approach, there is a contribution from
226 > self-neutralization of site $i$.  The self term for multipoles is
227 > described in section \ref{sec:selfTerm}.
228  
229 + \subsection{The multipole expansion}
230 +
231   Consider two discrete rigid collections of point charges, denoted as
232 < objects $\bf a$ and $\bf b$.  In the following, we assume that the two
233 < objects only interact via electrostatics and describe those
234 < interactions in terms of a multipole expansion.  Putting the origin of
235 < the coordinate system at the center of mass of $\bf a$, we use vectors
232 > $\bf a$ and $\bf b$.  In the following, we assume that the two objects
233 > interact via electrostatics only and describe those interactions in
234 > terms of a standard multipole expansion.  Putting the origin of the
235 > coordinate system at the center of mass of $\bf a$, we use vectors
236   $\mathbf{r}_k$ to denote the positions of all charges $q_k$ in $\bf
237   a$.  Then the electrostatic potential of object $\bf a$ at
238   $\mathbf{r}$ is given by
# Line 217 | Line 240 | We write the Taylor series expansion in $r$ using an i
240   V_a(\mathbf r) = \frac{1}{4\pi\epsilon_0}
241   \sum_{k \, \text{in \bf a}} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}.
242   \end{equation}
243 < We write the Taylor series expansion in $r$ using an implied summation
244 < notation, Greek indices are used to indicate space coordinates ($x$,
245 < $y$, $z$) and the subscripts $k$ and $j$ are reserved for labelling
246 < specific charges in $\bf a$ and $\bf b$ respectively, and find:
243 > The Taylor expansion in $r$ can be written using an implied summation
244 > notation.  Here Greek indices are used to indicate space coordinates
245 > ($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for
246 > labelling specific charges in $\bf a$ and $\bf b$ respectively.  The
247 > Taylor expansion,
248   \begin{equation}
249   \frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} =
250   \left( 1
251   -  r_{k\alpha} \frac{\partial}{\partial r_{\alpha}}
252   + \frac{1}{2}  r_{k\alpha} r_{k\beta} \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} +\dots
253   \right)
254 < \frac{1}{r}  .
254 > \frac{1}{r}  ,
255   \end{equation}
256 < The electrostatic potential on $\bf a$ can then be expressed in terms
257 < of multipole operators,
256 > can then be used to express the electrostatic potential on $\bf a$ in
257 > terms of multipole operators,
258   \begin{equation}
259   V_{\bf a}(\mathbf{r}) = \frac{1}{4\pi\epsilon_0}\hat{M}_{\bf a} \frac{1}{r}
260   \end{equation}
# Line 242 | Line 266 | given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\b
266   \end{equation}
267   Here, the point charge, dipole, and quadrupole for object $\bf a$ are
268   given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
269 <    a}\alpha\beta}$, respectively.  These can be tied to distribution
270 < of charges
271 < \begin{equation}
272 < C_{\bf a}=\sum_{k \, \text{in \bf a}} q_k ,
273 < \end{equation}
274 < \begin{equation}
275 < D_{{\bf a}\alpha} = \sum_{k \, \text{in \bf a}} q_k r_{k\alpha} ,
276 < \end{equation}
277 < \begin{equation}
278 < Q_{{\bf a}\alpha\beta} = \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k r_{k\alpha}  r_{k\beta} .
255 < \end{equation}
269 >    a}\alpha\beta}$, respectively.  These are the primitive multipoles
270 > which can be expressed as a distribution of charges,
271 > \begin{align}
272 > C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \\
273 > D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha} ,\\
274 > Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k r_{k\alpha}  r_{k\beta} .
275 > \end{align}
276 > Note that the definition of the primitive quadrupole here differs from
277 > the standard traceless form, and contains an additional Taylor-series
278 > based factor of $1/2$.
279  
280   It is convenient to locate charges $q_j$ relative to the center of mass of  $\bf b$.  Then with $\bf{r}$ pointing from
281   $\bf a$ to $\bf b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_b $), the interaction energy is given by
259 \begin{eqnarray}
260 U_{\bf{ab}}(r) =&& \frac{1}{4\pi \epsilon_0}
261 \sum_{k \, \text{in \bf a}} \sum_{j \, \text{in \bf b}}
262 \frac{q_k q_j}{\vert \bf{r}_k - (\bf{r}+\bf{r}_j) \vert} \nonumber\\
263 =&& \frac{1}{4\pi \epsilon_0}
264 \sum_{k \, \text{in \bf a}} \sum_{j \, \text{in \bf b}}
265 \frac{q_k q_j}{\vert \bf{r}+ (\bf{r}_j-\bf{r}_k) \vert} \nonumber\\
266 =&&\frac{1}{4\pi \epsilon_0}  \sum_{j \, \text{in \bf b}}  q_j V_a(\bf{r}+\bf{r}_j) \nonumber\\
267 =&&\frac{1}{4\pi \epsilon_0} \hat{M}_a \sum_{j \, \text{in \bf b}} \frac {q_j}{\vert \bf{r}+\bf{r}_j \vert} .
268 \end{eqnarray}
269 The last expression can also be expanded as a Taylor series in $r$.  Using a notation similar to before, we define
282   \begin{equation}
283 + U_{\bf{ab}}(r)
284 + = \frac{1}{4\pi \epsilon_0} \hat{M}_a \sum_{j \, \text{in \bf b}} \frac {q_j}{\vert \bf{r}+\bf{r}_j \vert} .
285 + \end{equation}
286 + This can also be expanded as a Taylor series in $r$.  Using a notation
287 + similar to before to define the multipoles on object {\bf b},
288 + \begin{equation}
289   \hat{M}_{\bf b} = C_{\bf b} + D_{{\bf b}\alpha} \frac{\partial}{\partial r_{\alpha}}
290   +   Q_{{\bf b}\alpha\beta}
291   \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
292   \end{equation}
293 < and
293 > we arrive at the multipole expression for the total interaction energy.
294   \begin{equation}
295   U_{\bf{ab}}(r)=\frac{\hat{M}_{\bf a} \hat{M}_{\bf b}}{4\pi \epsilon_0} \frac{1}{r}  \label{kernel}.
296   \end{equation}
297 < Note the ease of separting out the respective energies of interaction of the charge, dipole, and quadrupole of $\bf a$  from those of $\bf b$.
297 > This form has the benefit of separating out the energies of
298 > interaction into contributions from the charge, dipole, and quadrupole
299 > of $\bf a$ interacting with the same multipoles $\bf b$.
300  
301 < \subsection{Bare Coulomb versus smeared charge}
302 <
303 < With the four types of methods being considered here, it is desirable to list the approximations in as transparent a form
304 < as possible.  First, one may use the bare Coulomb potential, with radial dependence $1/r$,
305 < as shown in Eq.~(\ref{kernel}).  Alternatively, one may use
306 < a smeared charge distribution, then the``kernel'' $1/r$ of the expansion is replaced with a function:
301 > \subsection{Damped Coulomb interactions}
302 > In the standard multipole expansion, one typically uses the bare
303 > Coulomb potential, with radial dependence $1/r$, as shown in
304 > Eq.~(\ref{kernel}).  It is also quite common to use a damped Coulomb
305 > interaction, which results from replacing point charges with Gaussian
306 > distributions of charge with width $\alpha$.  In damped multipole
307 > electrostatics, the kernel ($1/r$) of the expansion is replaced with
308 > the function:
309   \begin{equation}
310   B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}
311   \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
312   \end{equation}
313 < We develop equations using a function $f(r)$ to represent either $1/r$ or $B_0(r)$, dependent on the the type of approach being considered.
314 < Smith's convenient functions $B_l(r)$ are summarized in Appendix A.
313 > We develop equations below using the function $f(r)$ to represent
314 > either $1/r$ or $B_0(r)$, and all of the techniques can be applied
315 > either to bare or damped Coulomb kernels as long as derivatives of
316 > these functions are known.  Smith's convenient functions $B_l(r)$ are
317 > summarized in Appendix A.
318  
319 < \subsection{Shifting the force, method 1}
319 > \subsection{Taylor-shifted force (TSF) electrostatics}
320  
321 < As discussed in the introduction, it is desirable to cutoff the electrostatic energy at a radius
322 < $r_c$.  For consistency in approximation, we want the interaction energy as well as the force and
323 < torque to go to zero at $r=r_c$.  
324 < To describe how this goal may be met using a radial approximation, we use two examples, charge-charge
325 < and charge-dipole, using the bare Coulomb kernel $f(r)=1/r$ to explain the idea.
321 > The main goal of this work is to smoothly cut off the interaction
322 > energy as well as forces and torques as $r\rightarrow r_c$.  To
323 > describe how this goal may be met, we use two examples, charge-charge
324 > and charge-dipole, using the bare Coulomb kernel $f(r)=1/r$ to explain
325 > the idea.
326  
327 < In the shifted-force approximation, the interaction energy $U_{\bf{ab}}(r_c)=0$
328 < for two charges $C_{\bf a}$ and $C_{\bf b}$ separated by a distance $r$ is written:
327 > In the shifted-force approximation, the interaction energy for two
328 > charges $C_{\bf a}$ and $C_{\bf b}$ separated by a distance $r$ is
329 > written:
330   \begin{equation}
331   U_{C_{\bf a}C_{\bf b}}(r)=\frac{1}{4\pi \epsilon_0} C_{\bf a} C_{\bf b}
332   \left({ \frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2}  }
333   \right) .
334   \end{equation}
335 < Two shifting terms appear in this equations because we want the force to
336 < also be shifted due to an image charge located at a distance $r_c$.  
337 < Since one derivative of the interaction energy is needed for the force, we want a term
338 < linear in $(r-r_c)$ in the interaction energy, that is:
335 > Two shifting terms appear in this equations, one from the
336 > neutralization procedure ($-1/r_c$), and one that will make the first
337 > derivative also vanish at the cutoff radius.  
338 >
339 > Since one derivative of the interaction energy is needed for the
340 > force, the minimal perturbation is a term linear in $(r-r_c)$ in the
341 > interaction energy, that is:
342   \begin{equation}
343   \frac{d\,}{dr}
344   \left( {\frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2}  }
345   \right) = \left(- \frac{1}{r^2} + \frac{1}{r_c^2}
346   \right) .
347   \end{equation}
348 < This demonstrates the need of the third term in the brackets of the energy expression, but  leads to the question, how does this idea generalize for higher-order multipole energies?
348 > There are a number of ways to generalize this derivative shift for
349 > higher-order multipoles.
350  
351 < In method 1, the procedure that we follow is based on the number of derivatives need for each energy, force, or torque.  That is,
352 < a quadrupole-quadrupole interaction energy will have four derivatives,
353 < $\partial^4/\partial r_\alpha \partial r_\beta \partial r_\gamma \partial r_\delta$,
354 < and the force or torque will bring in yet another derivative.  
355 < We thus want shifted energy expressions to include terms so that all energies, forces, and torques
356 < are zero at $r=r_c$.  In each case, we will subtract off a function $f_n^{\text{shift}}(r)$ from the
357 < kernel $f(r)=1/r$.  The index $n$ indicates the number of derivatives to be taken when
358 < deriving a given multipole energy.
359 < We choose a function with guaranteed smooth derivatives --- a truncated Taylor series of the function
360 < $f(r)$, e.g.,
351 > In the Taylor-shifted force (TSF) method, the procedure that we follow
352 > is based on a Taylor expansion containing the same number of
353 > derivatives required for each force term to vanish at the cutoff.  For
354 > example, the quadrupole-quadrupole interaction energy requires four
355 > derivatives of the kernel, and the force requires one additional
356 > derivative. We therefore require shifted energy expressions that
357 > include enough terms so that all energies, forces, and torques are
358 > zero as $r \rightarrow r_c$.  In each case, we will subtract off a
359 > function $f_n^{\text{shift}}(r)$ from the kernel $f(r)=1/r$.  The
360 > index $n$ indicates the number of derivatives to be taken when
361 > deriving a given multipole energy.  We choose a function with
362 > guaranteed smooth derivatives --- a truncated Taylor series of the
363 > function $f(r)$, e.g.,
364   %
365   \begin{equation}
366   f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
# Line 340 | Line 373 | Continuing with the example of a charge  $\bf a$  inte
373   f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
374   \end{equation}
375   %
376 < Continuing with the example of a charge  $\bf a$  interacting with a dipole  $\bf b$, we write
376 > Continuing with the example of a charge $\bf a$ interacting with a
377 > dipole $\bf b$, we write
378   %
379   \begin{equation}
380   U_{C_{\bf a}D_{\bf b}}(r)=
# Line 391 | Line 425 | in Appendices A and B.  
425   of the many derivatives that are taken, the algebra is tedious and are summarized
426   in Appendices A and B.  
427  
428 < \subsection{Shifting the force, method 2}
428 > \subsection{Gradient-shifted force (GSF) electrostatics}
429  
430   Note the method used in the previous subsection to shift a force is basically that of using
431   a truncated Taylor Series in the radius $r$.  An alternate method exists, best explained by
# Line 494 | Line 528 | be discussed below.  
528   be discussed below.  
529  
530  
531 < \section{The Self-Interaction}
531 > \subsection{The Self-Interaction \label{sec:selfTerm}}
532 >
533   The Wolf summation~\cite{Wolf99} and the later damped shifted force
534   (DSF) extension~\cite{Fennell06} included self-interactions that are
535   handled separately from the pairwise interactions between sites. The

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines