1 |
% ****** Start of file aipsamp.tex ****** |
2 |
% |
3 |
% This file is part of the AIP files in the AIP distribution for REVTeX 4. |
4 |
% Version 4.1 of REVTeX, October 2009 |
5 |
% |
6 |
% Copyright (c) 2009 American Institute of Physics. |
7 |
% |
8 |
% See the AIP README file for restrictions and more information. |
9 |
% |
10 |
% TeX'ing this file requires that you have AMS-LaTeX 2.0 installed |
11 |
% as well as the rest of the prerequisites for REVTeX 4.1 |
12 |
% |
13 |
% It also requires running BibTeX. The commands are as follows: |
14 |
% |
15 |
% 1) latex aipsamp |
16 |
% 2) bibtex aipsamp |
17 |
% 3) latex aipsamp |
18 |
% 4) latex aipsamp |
19 |
% |
20 |
% Use this file as a source of example code for your aip document. |
21 |
% Use the file aiptemplate.tex as a template for your document. |
22 |
\documentclass[% |
23 |
aip,jcp, |
24 |
amsmath,amssymb, |
25 |
preprint,% |
26 |
% reprint,% |
27 |
%author-year,% |
28 |
%author-numerical,% |
29 |
jcp]{revtex4-1} |
30 |
|
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 |
\usepackage{rotating} |
38 |
|
39 |
%\usepackage[mathlines]{lineno}% Enable numbering of text and display math |
40 |
%\linenumbers\relax % Commence numbering lines |
41 |
|
42 |
\begin{document} |
43 |
|
44 |
%\preprint{AIP/123-QED} |
45 |
|
46 |
\title{Real space electrostatics for multipoles. I. Development of methods} |
47 |
|
48 |
\author{Madan Lamichhane} |
49 |
\affiliation{Department of Physics, University |
50 |
of Notre Dame, Notre Dame, IN 46556} |
51 |
|
52 |
\author{J. Daniel Gezelter} |
53 |
\email{gezelter@nd.edu.} |
54 |
\affiliation{Department of Chemistry and Biochemistry, University |
55 |
of Notre Dame, Notre Dame, IN 46556} |
56 |
|
57 |
\author{Kathie E. Newman} |
58 |
\affiliation{Department of Physics, University |
59 |
of Notre Dame, Notre Dame, IN 46556} |
60 |
|
61 |
|
62 |
\date{\today}% It is always \today, today, |
63 |
% but any date may be explicitly specified |
64 |
|
65 |
\begin{abstract} |
66 |
We have extended the original damped-shifted force (DSF) |
67 |
electrostatic kernel and have been able to derive three new |
68 |
electrostatic potentials for higher-order multipoles that are based |
69 |
on truncated Taylor expansions around the cutoff radius. These |
70 |
include a shifted potential (SP) that generalizes the Wolf method |
71 |
for point multipoles, and Taylor-shifted force (TSF) and |
72 |
gradient-shifted force (GSF) potentials that are both |
73 |
generalizations of DSF electrostatics for multipoles. We find that |
74 |
each of the distinct orientational contributions requires a separate |
75 |
radial function to ensure that pairwise energies, forces and torques |
76 |
all vanish at the cutoff radius. In this paper, we present energy, |
77 |
force, and torque expressions for the new models, and compare these |
78 |
real-space interaction models to exact results for ordered arrays of |
79 |
multipoles. We find that the GSF and SP methods converge rapidly to |
80 |
the correct lattice energies for ordered dipolar and quadrupolar |
81 |
arrays, while the Taylor-Shifted Force (TSF) is too severe an |
82 |
approximation to provide accurate convergence to lattice energies. |
83 |
Because real-space methods can be made to scale linearly with system |
84 |
size, SP and GSF are attractive options for large Monte |
85 |
Carlo and molecular dynamics simulations, respectively. |
86 |
\end{abstract} |
87 |
|
88 |
%\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy |
89 |
% Classification Scheme. |
90 |
%\keywords{Suggested keywords}%Use showkeys class option if keyword |
91 |
%display desired |
92 |
\maketitle |
93 |
|
94 |
\section{Introduction} |
95 |
There has been increasing interest in real-space methods for |
96 |
calculating electrostatic interactions in computer simulations of |
97 |
condensed molecular |
98 |
systems.\cite{Wolf99,Zahn02,Kast03,BeckD.A.C._bi0486381,Ma05,Fennell:2006zl,Chen:2004du,Chen:2006ii,Rodgers:2006nw,Denesyuk:2008ez,Izvekov:2008wo} |
99 |
The simplest of these techniques was developed by Wolf {\it et al.} |
100 |
in their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99} |
101 |
For systems of point charges, Fennell and Gezelter showed that a |
102 |
simple damped shifted force (DSF) modification to Wolf's method could |
103 |
give nearly quantitative agreement with smooth particle mesh Ewald |
104 |
(SPME)\cite{Essmann95} configurational energy differences as well as |
105 |
atomic force and molecular torque vectors.\cite{Fennell:2006zl} |
106 |
|
107 |
The computational efficiency and the accuracy of the DSF method are |
108 |
surprisingly good, particularly for systems with uniform charge |
109 |
density. Additionally, dielectric constants obtained using DSF and |
110 |
similar methods where the force vanishes at $r_{c}$ are |
111 |
essentially quantitative.\cite{Izvekov:2008wo} The DSF and other |
112 |
related methods have now been widely investigated,\cite{Hansen:2012uq} |
113 |
and DSF is now used routinely in a diverse set of chemical |
114 |
environments.\cite{doi:10.1021/la400226g,McCann:2013fk,kannam:094701,Forrest:2012ly,English:2008kx,Louden:2013ve,Tokumasu:2013zr} |
115 |
DSF electrostatics provides a compromise between the computational |
116 |
speed of real-space cutoffs and the accuracy of fully-periodic Ewald |
117 |
treatments. |
118 |
|
119 |
One common feature of many coarse-graining approaches, which treat |
120 |
entire molecular subsystems as a single rigid body, is simplification |
121 |
of the electrostatic interactions between these bodies so that fewer |
122 |
site-site interactions are required to compute configurational |
123 |
energies. To do this, the interactions between coarse-grained sites |
124 |
are typically taken to be point |
125 |
multipoles.\cite{Golubkov06,ISI:000276097500009,ISI:000298664400012} |
126 |
|
127 |
Water, in particular, has been modeled recently with point multipoles |
128 |
up to octupolar |
129 |
order.\cite{Chowdhuri:2006lr,Te:2010rt,Te:2010ys,Te:2010vn} For |
130 |
maximum efficiency, these models require the use of an approximate |
131 |
multipole expansion as the exact multipole expansion can become quite |
132 |
expensive (particularly when handled via the Ewald |
133 |
sum).\cite{Ichiye:2006qy} Point multipoles and multipole |
134 |
polarizability have also been utilized in the AMOEBA water model and |
135 |
related force fields.\cite{Ponder:2010fk,schnieders:124114,Ren:2011uq} |
136 |
|
137 |
Higher-order multipoles present a peculiar issue for molecular |
138 |
dynamics. Multipolar interactions are inherently short-ranged, and |
139 |
should not need the relatively expensive Ewald treatment. However, |
140 |
real-space cutoff methods are normally applied in an orientation-blind |
141 |
fashion so multipoles which leave and then re-enter a cutoff sphere in |
142 |
a different orientation can cause energy discontinuities. |
143 |
|
144 |
This paper outlines an extension of the original DSF electrostatic |
145 |
kernel to point multipoles. We describe three distinct real-space |
146 |
interaction models for higher-order multipoles based on truncated |
147 |
Taylor expansions that are carried out at the cutoff radius. We are |
148 |
calling these models {\bf Taylor-shifted} (TSF), {\bf |
149 |
gradient-shifted} (GSF) and {\bf shifted potential} (SP) |
150 |
electrostatics. Because of differences in the initial assumptions, |
151 |
the two methods yield related, but distinct expressions for energies, |
152 |
forces, and torques. |
153 |
|
154 |
In this paper we outline the new methodology and give functional forms |
155 |
for the energies, forces, and torques up to quadrupole-quadrupole |
156 |
order. We also compare the new methods to analytic energy constants |
157 |
for periodic arrays of point multipoles. In the following paper, we |
158 |
provide numerical comparisons to Ewald-based electrostatics in common |
159 |
simulation enviornments. |
160 |
|
161 |
\section{Methodology} |
162 |
An efficient real-space electrostatic method involves the use of a |
163 |
pair-wise functional form, |
164 |
\begin{equation} |
165 |
U = \sum_i \sum_{j>i} U_\mathrm{pair}(\mathbf{r}_{ij}, \mathsf{A}_i, \mathsf{B}_j) + |
166 |
\sum_i U_i^\mathrm{self} |
167 |
\end{equation} |
168 |
that is short-ranged and easily truncated at a cutoff radius, |
169 |
\begin{equation} |
170 |
U_\mathrm{pair}(\mathbf{r}_{ij},\mathsf{A}_i, \mathsf{B}_j) = \left\{ |
171 |
\begin{array}{ll} |
172 |
U_\mathrm{approx} (\mathbf{r}_{ij}, \mathsf{A}_i, \mathsf{B}_j) & \quad \left| \mathbf{r}_{ij} \right| \le r_c \\ |
173 |
0 & \quad \left| \mathbf{r}_{ij} \right| > r_c , |
174 |
\end{array} |
175 |
\right. |
176 |
\end{equation} |
177 |
along with an easily computed self-interaction term ($\sum_i |
178 |
U_i^\mathrm{self}$) which scales linearly with the number of |
179 |
particles. Here $\mathsf{A}_i$ and $\mathsf{B}_j$ represent orientational |
180 |
coordinates of the two sites (expressed as rotation matrices), and $\mathbf{r}_{ij}$ is the vector |
181 |
between the two sites. The computational efficiency, energy |
182 |
conservation, and even some physical properties of a simulation can |
183 |
depend dramatically on how the $U_\mathrm{approx}$ function behaves at |
184 |
the cutoff radius. The goal of any approximation method should be to |
185 |
mimic the real behavior of the electrostatic interactions as closely |
186 |
as possible without sacrificing the near-linear scaling of a cutoff |
187 |
method. |
188 |
|
189 |
\subsection{Self-neutralization, damping, and force-shifting} |
190 |
The DSF and Wolf methods operate by neutralizing the total charge |
191 |
contained within the cutoff sphere surrounding each particle. This is |
192 |
accomplished by shifting the potential functions to generate image |
193 |
charges on the surface of the cutoff sphere for each pair interaction |
194 |
computed within $r_c$. Damping using a complementary error function is |
195 |
applied to the potential to accelerate convergence. The interaction |
196 |
for a pair of charges ($C_i$ and $C_j$) in the DSF method, |
197 |
\begin{equation*} |
198 |
U_\mathrm{DSF}(r_{ij}) = C_i C_j \Biggr{[} |
199 |
\frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}} |
200 |
- \frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c} + \left(\frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c^2} |
201 |
+ \frac{2\alpha}{\pi^{1/2}} |
202 |
\frac{\exp\left(-\alpha^2r_c^2\right)}{r_c} |
203 |
\right)\left(r_{ij}-r_c\right)\ \Biggr{]} |
204 |
\label{eq:DSFPot} |
205 |
\end{equation*} |
206 |
where $\alpha$ is the adjustable damping parameter. Note that in this |
207 |
potential and in all electrostatic quantities that follow, the |
208 |
standard $1/4 \pi \epsilon_{0}$ has been omitted for clarity. |
209 |
|
210 |
To insure net charge neutrality within each cutoff sphere, an |
211 |
additional ``self'' term is added to the potential. This term is |
212 |
constant (as long as the charges and cutoff radius do not change), and |
213 |
exists outside the normal pair-loop for molecular simulations. It can |
214 |
be thought of as a contribution from a charge opposite in sign, but |
215 |
equal in magnitude, to the central charge, which has been spread out |
216 |
over the surface of the cutoff sphere. A portion of the self term is |
217 |
identical to the self term in the Ewald summation, and comes from the |
218 |
utilization of the complimentary error function for electrostatic |
219 |
damping.\cite{deLeeuw80,Wolf99} There have also been recent efforts to |
220 |
extend the Wolf self-neutralization method to zero out the dipole and |
221 |
higher order multipoles contained within the cutoff |
222 |
sphere.\cite{Fukuda:2011jk,Fukuda:2012yu,Fukuda:2013qv} |
223 |
|
224 |
In this work, we extend the idea of self-neutralization for the point |
225 |
multipoles by insuring net charge-neutrality and net-zero moments |
226 |
within each cutoff sphere. In Figure \ref{fig:shiftedMultipoles}, point dipole |
227 |
$\mathbf{D}_i$ at the center site $i$ is interacting with point dipole |
228 |
$\mathbf{D}_j$ and point quadrupole $\mathsf{Q}_k$. The |
229 |
self-neutralization scheme for point multipoles involves projecting |
230 |
opposing multipoles for sites $j$ and $k$ on the surface of the cutoff |
231 |
sphere. There are also significant modifications made to make the |
232 |
forces and torques go smoothly to zero at the cutoff distance. |
233 |
|
234 |
\begin{figure} |
235 |
\includegraphics[width=3in]{SM.eps} |
236 |
\caption{Reversed multipoles are projected onto the surface of the |
237 |
cutoff sphere. The forces, torques, and potential are then smoothly |
238 |
shifted to zero as the sites leave the cutoff region.} |
239 |
\label{fig:shiftedMultipoles} |
240 |
\end{figure} |
241 |
|
242 |
As in the point-charge approach, there is an additional contribution |
243 |
from self-neutralization of site $i$. The self term for multipoles is |
244 |
described in section \ref{sec:selfTerm}. |
245 |
|
246 |
\subsection{The multipole expansion} |
247 |
|
248 |
Consider two discrete rigid collections of point charges, denoted as objects |
249 |
$a$ and $b$. In the following, we assume that the two objects |
250 |
interact via electrostatics only and describe those interactions in |
251 |
terms of a standard multipole expansion. Putting the origin of the |
252 |
coordinate system at the center of mass of $a$, we use vectors |
253 |
$\mathbf{r}_k$ to denote the positions of all charges $q_k$ in |
254 |
$a$. Then the electrostatic potential of object $a$ at |
255 |
$\mathbf{r}$ is given by |
256 |
\begin{equation} |
257 |
\phi_a(\mathbf r) = |
258 |
\sum_{k \, \text{in } a} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}. |
259 |
\end{equation} |
260 |
The Taylor expansion in $r$ can be written using an implied summation |
261 |
notation. Here Greek indices are used to indicate space coordinates |
262 |
($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for |
263 |
labeling specific sites for charges in $a$ and $b$ respectively. The |
264 |
Taylor expansion, |
265 |
\begin{equation} |
266 |
\frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} = |
267 |
\left( 1 |
268 |
- r_{k\alpha} \frac{\partial}{\partial r_{\alpha}} |
269 |
+ \frac{1}{2} r_{k\alpha} r_{k\beta} \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} +\dots |
270 |
\right) |
271 |
\frac{1}{r} , |
272 |
\end{equation} |
273 |
can then be used to express the electrostatic potential on $a$ in |
274 |
terms of multipole operators, |
275 |
\begin{equation} |
276 |
\phi_a(\mathbf{r}) =M_a \frac{1}{r} |
277 |
\end{equation} |
278 |
where |
279 |
\begin{equation} |
280 |
M_a = C_a - D_{a\alpha} \frac{\partial}{\partial r_{\alpha}} |
281 |
+ Q_{a\alpha\beta} |
282 |
\frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots |
283 |
\end{equation} |
284 |
Here, the point charge, dipole, and quadrupole for object $a$ are |
285 |
given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively. These are the primitive multipoles |
286 |
which can be expressed as a distribution of charges, |
287 |
\begin{align} |
288 |
C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\ |
289 |
D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\ |
290 |
Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k |
291 |
r_{k\alpha} r_{k\beta} . \label{eq:quadrupole} |
292 |
\end{align} |
293 |
Note that the definition of the primitive quadrupole here differs from |
294 |
the standard traceless form, and contains an additional Taylor-series |
295 |
based factor of $1/2$. We are essentially treating the mass |
296 |
distribution with higher priority; the moment of inertia tensor, |
297 |
$\mathsf I$, is diagonalized to obtain body-fixed |
298 |
axes, and the charge distribution may result in a quadrupole tensor |
299 |
that is not necessarily diagonal in the body frame. Additional |
300 |
reasons for utilizing the primitive quadrupole are discussed in |
301 |
section \ref{sec:damped}. |
302 |
|
303 |
It is convenient to locate charges $q_j$ relative to the center of mass of $b$. Then with $\bf{r}$ pointing from |
304 |
the center of mass of $a$ to the center of mass of $b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $), the interaction energy is given by |
305 |
\begin{equation} |
306 |
U_{ab}(r) |
307 |
= M_a \sum_{j \, \text{in } b} \frac {q_j}{\vert \mathbf{r}+\mathbf{r}_j \vert} . |
308 |
\end{equation} |
309 |
This can also be expanded as a Taylor series in $r$. Using a notation |
310 |
similar to before to define the multipoles in object $b$, |
311 |
\begin{equation} |
312 |
M_b = C_b + D_{b\alpha} \frac{\partial}{\partial r_{\alpha}} |
313 |
+ Q_{b\alpha\beta} |
314 |
\frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots |
315 |
\end{equation} |
316 |
we arrive at the multipole expression for the total interaction energy. |
317 |
\begin{equation} |
318 |
U_{ab}(r)=M_a M_b \frac{1}{r} \label{kernel}. |
319 |
\end{equation} |
320 |
This form has the benefit of separating out the energies of |
321 |
interaction into contributions from the charge, dipole, and quadrupole |
322 |
of $a$ interacting with the same types of multipoles in $b$. |
323 |
|
324 |
\subsection{Damped Coulomb interactions} |
325 |
\label{sec:damped} |
326 |
In the standard multipole expansion, one typically uses the bare |
327 |
Coulomb potential, with radial dependence $1/r$, as shown in |
328 |
Eq.~(\ref{kernel}). It is also quite common to use a damped Coulomb |
329 |
interaction, which results from replacing point charges with Gaussian |
330 |
distributions of charge with width $\alpha$. In damped multipole |
331 |
electrostatics, the kernel ($1/r$) of the expansion is replaced with |
332 |
the function: |
333 |
\begin{equation} |
334 |
B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r} |
335 |
\int_{\alpha r}^{\infty} \text{e}^{-s^2} ds . |
336 |
\end{equation} |
337 |
We develop equations below using the function $f(r)$ to represent |
338 |
either $1/r$ or $B_0(r)$, and all of the techniques can be applied to |
339 |
bare or damped Coulomb kernels (or any other function) as long as |
340 |
derivatives of these functions are known. Smith's convenient |
341 |
functions $B_l(r)$, which are used for derivatives of the damped |
342 |
kernel, are summarized in Appendix A. (N.B. there is one important |
343 |
distinction between the two kernels, which is the behavior of |
344 |
$\nabla^2 \frac{1}{r}$ compared with $\nabla^2 B_0(r)$. The former is |
345 |
zero everywhere except for a delta function evaluated at the origin. |
346 |
The latter also has delta function behavior, but is non-zero for $r |
347 |
\neq 0$. Thus the standard justification for using a traceless |
348 |
quadrupole tensor fails for the damped case.) |
349 |
|
350 |
The main goal of this work is to smoothly cut off the interaction |
351 |
energy as well as forces and torques as $r\rightarrow r_c$. To |
352 |
describe how this goal may be met, we use two examples, charge-charge |
353 |
and charge-dipole, using the bare Coulomb kernel, $f(r)=1/r$, to |
354 |
explain the idea. |
355 |
|
356 |
\subsection{Shifted-force methods} |
357 |
In the shifted-force approximation, the interaction energy for two |
358 |
charges $C_a$ and $C_b$ separated by a distance $r$ is |
359 |
written: |
360 |
\begin{equation} |
361 |
U_{C_aC_b}(r)= C_a C_b |
362 |
\left({ \frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} } |
363 |
\right) . |
364 |
\end{equation} |
365 |
Two shifting terms appear in this equations, one from the |
366 |
neutralization procedure ($-1/r_c$), and one that causes the first |
367 |
derivative to vanish at the cutoff radius. |
368 |
|
369 |
Since one derivative of the interaction energy is needed for the |
370 |
force, the minimal perturbation is a term linear in $(r-r_c)$ in the |
371 |
interaction energy, that is: |
372 |
\begin{equation} |
373 |
\frac{d\,}{dr} |
374 |
\left( {\frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} } |
375 |
\right) = \left(- \frac{1}{r^2} + \frac{1}{r_c^2} |
376 |
\right) . |
377 |
\end{equation} |
378 |
which clearly vanishes as the $r$ approaches the cutoff radius. There |
379 |
are a number of ways to generalize this derivative shift for |
380 |
higher-order multipoles. Below, we present two methods, one based on |
381 |
higher-order Taylor series for $r$ near $r_c$, and the other based on |
382 |
linear shift of the kernel gradients at the cutoff itself. |
383 |
|
384 |
\subsection{Taylor-shifted force (TSF) electrostatics} |
385 |
In the Taylor-shifted force (TSF) method, the procedure that we follow |
386 |
is based on a Taylor expansion containing the same number of |
387 |
derivatives required for each force term to vanish at the cutoff. For |
388 |
example, the quadrupole-quadrupole interaction energy requires four |
389 |
derivatives of the kernel, and the force requires one additional |
390 |
derivative. For quadrupole-quadrupole interactions, we therefore |
391 |
require shifted energy expressions that include up to $(r-r_c)^5$ so |
392 |
that all energies, forces, and torques are zero as $r \rightarrow |
393 |
r_c$. In each case, we subtract off a function $f_n^{\text{shift}}(r)$ |
394 |
from the kernel $f(r)=1/r$. The subscript $n$ indicates the number of |
395 |
derivatives to be taken when deriving a given multipole energy. We |
396 |
choose a function with guaranteed smooth derivatives -- a truncated |
397 |
Taylor series of the function $f(r)$, e.g., |
398 |
% |
399 |
\begin{equation} |
400 |
f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)}(r_c) . |
401 |
\end{equation} |
402 |
% |
403 |
The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$. |
404 |
Thus, for $f(r)=1/r$, we find |
405 |
% |
406 |
\begin{equation} |
407 |
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} . |
408 |
\end{equation} |
409 |
% |
410 |
Continuing with the example of a charge $a$ interacting with a |
411 |
dipole $b$, we write |
412 |
% |
413 |
\begin{equation} |
414 |
U_{C_a\mathbf{D}_b}(r)= |
415 |
C_a D_{b\alpha} \frac {\partial f_1(r) }{\partial r_\alpha} |
416 |
= C_a D_{b\alpha} |
417 |
\frac {r_\alpha}{r} \frac {\partial f_1(r)}{\partial r} . |
418 |
\end{equation} |
419 |
% |
420 |
The force that dipole $ b$ exerts on charge $a$ is |
421 |
% |
422 |
\begin{equation} |
423 |
F_{C_a \mathbf{D}_b \beta} = C_a D_{b \alpha} |
424 |
\left[ \frac{\delta_{\alpha\beta}}{r} \frac {\partial}{\partial r} + |
425 |
\frac{r_\alpha r_\beta}{r^2} |
426 |
\left( -\frac{1}{r} \frac {\partial} {\partial r} |
427 |
+ \frac {\partial ^2} {\partial r^2} \right) \right] f_1(r) . |
428 |
\end{equation} |
429 |
% |
430 |
For undamped coulombic interactions, $f(r)=1/r$, we find |
431 |
% |
432 |
\begin{equation} |
433 |
F_{C_a \mathbf{D}_b \beta} = |
434 |
\frac{C_a D_{b\beta}}{r} |
435 |
\left[ -\frac{1}{r^2}+\frac{1}{r_c^2}-\frac{2(r-r_c)}{r_c^3} \right] |
436 |
+C_a D_{b \alpha}r_\alpha r_\beta |
437 |
\left[ \frac{3}{r^5}-\frac{3}{r^3r_c^2} \right] . |
438 |
\end{equation} |
439 |
% |
440 |
This expansion shows the expected $1/r^3$ dependence of the force. |
441 |
|
442 |
In general, we can write |
443 |
% |
444 |
\begin{equation} |
445 |
U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r) |
446 |
\label{generic} |
447 |
\end{equation} |
448 |
% |
449 |
with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for |
450 |
charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and |
451 |
$n=4$ for quadrupole-quadrupole. For example, in |
452 |
quadrupole-quadrupole interactions for which the $\text{prefactor}$ is |
453 |
$Q_{a \alpha\beta}Q_{b \gamma\delta}$, the derivatives are |
454 |
$\partial^4/\partial r_\alpha \partial r_\beta \partial |
455 |
r_\gamma \partial r_\delta$, with implied summation combining the |
456 |
space indices. Appendix \ref{radialTSF} contains details on the |
457 |
radial functions. |
458 |
|
459 |
In the formulas presented in the tables below, the placeholder |
460 |
function $f(r)$ is used to represent the electrostatic kernel (either |
461 |
damped or undamped). The main functions that go into the force and |
462 |
torque terms, $g_n(r), h_n(r), s_n(r), \mathrm{~and~} t_n(r)$ are |
463 |
successive derivatives of the shifted electrostatic kernel, $f_n(r)$ |
464 |
of the same index $n$. The algebra required to evaluate energies, |
465 |
forces and torques is somewhat tedious, so only the final forms are |
466 |
presented in tables \ref{tab:tableenergy} and \ref{tab:tableFORCE}. |
467 |
One of the principal findings of our work is that the individual |
468 |
orientational contributions to the various multipole-multipole |
469 |
interactions must be treated with distinct radial functions, but each |
470 |
of these contributions is independently force shifted at the cutoff |
471 |
radius. |
472 |
|
473 |
\subsection{Gradient-shifted force (GSF) electrostatics} |
474 |
The second, and conceptually simpler approach to force-shifting |
475 |
maintains only the linear $(r-r_c)$ term in the truncated Taylor |
476 |
expansion, and has a similar interaction energy for all multipole |
477 |
orders: |
478 |
\begin{equation} |
479 |
U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) - |
480 |
U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c) |
481 |
\hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right] |
482 |
\label{generic2} |
483 |
\end{equation} |
484 |
where $\hat{\mathbf{r}}$ is the unit vector pointing between the two |
485 |
multipoles, and the sum describes a separate force-shifting that is |
486 |
applied to each orientational contribution to the energy. Both the |
487 |
potential and the gradient for force shifting are evaluated for an |
488 |
image multipole projected onto the surface of the cutoff sphere (see |
489 |
fig \ref{fig:shiftedMultipoles}). The image multipole retains the |
490 |
orientation (rotation matrix $\mathsf{B}$) of the interacting multipole. No |
491 |
higher order terms $(r-r_c)^n$ appear. The primary difference between |
492 |
the TSF and GSF methods is the stage at which the Taylor Series is |
493 |
applied; in the Taylor-shifted approach, it is applied to the kernel |
494 |
itself. In the Gradient-shifted approach, it is applied to individual |
495 |
radial interaction terms in the multipole expansion. Energies from |
496 |
this method thus have the general form: |
497 |
\begin{equation} |
498 |
U= \sum (\text{angular factor}) (\text{radial factor}). |
499 |
\label{generic3} |
500 |
\end{equation} |
501 |
|
502 |
Functional forms for both methods (TSF and GSF) can both be summarized |
503 |
using the form of Eq.~\ref{generic3}). The basic forms for the |
504 |
energy, force, and torque expressions are tabulated for both shifting |
505 |
approaches below -- for each separate orientational contribution, only |
506 |
the radial factors differ between the two methods. |
507 |
|
508 |
\subsection{Generalization of the Wolf shifted potential (SP)} |
509 |
It is also possible to formulate an extension of the Wolf approach for |
510 |
multipoles by simply projecting the image multipole onto the surface |
511 |
of the cutoff sphere, and including the interactions with the central |
512 |
multipole and the image. This effectively shifts the pair potential |
513 |
to zero at the cutoff radius, |
514 |
\begin{equation} |
515 |
U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) - |
516 |
U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right] |
517 |
\label{eq:SP} |
518 |
\end{equation} |
519 |
independent of the orientations of the two multipoles. The sum again |
520 |
describes separate potential shifting that is applied to each |
521 |
orientational contribution to the energy. |
522 |
|
523 |
The shifted potential (SP) method is a simple truncation of the GSF |
524 |
method for each orientational contribution, leaving out the $(r-r_c)$ |
525 |
terms that multiply the gradient. Functional forms for the |
526 |
shifted-potential (SP) method can also be summarized using the form of |
527 |
Eq.~\ref{generic3}. The energy, force, and torque expressions are |
528 |
tabulated below for all three methods. As in the GSF and TSF methods, |
529 |
for each separate orientational contribution, only the radial factors |
530 |
differ between the SP, GSF, and TSF methods. |
531 |
|
532 |
|
533 |
\subsection{\label{sec:level2}Body and space axes} |
534 |
Although objects $a$ and $b$ rotate during a molecular |
535 |
dynamics (MD) simulation, their multipole tensors remain fixed in |
536 |
body-frame coordinates. While deriving force and torque expressions, |
537 |
it is therefore convenient to write the energies, forces, and torques |
538 |
in intermediate forms involving the vectors of the rotation matrices. |
539 |
We denote body axes for objects $a$ and $b$ using unit vectors |
540 |
$\hat{\mathbf{A}}_m$ and $\hat{\mathbf{B}}_m$, respectively, with the index $m=(123)$. |
541 |
In a typical simulation, the initial axes are obtained by |
542 |
diagonalizing the moment of inertia tensors for the objects. (N.B., |
543 |
the body axes are generally {\it not} the same as those for which the |
544 |
quadrupole moment is diagonal.) The rotation matrices are then |
545 |
propagated during the simulation. |
546 |
|
547 |
The rotation matrices $\mathsf {A}$ and $\mathsf {B}$ can be |
548 |
expressed using these unit vectors: |
549 |
\begin{eqnarray} |
550 |
\mathsf {A} = |
551 |
\begin{pmatrix} |
552 |
\hat{\mathbf{A}}_1 \\ |
553 |
\hat{\mathbf{A}}_2 \\ |
554 |
\hat{\mathbf{A}}_3 |
555 |
\end{pmatrix}, \qquad |
556 |
\mathsf {B} = |
557 |
\begin{pmatrix} |
558 |
\hat{\mathbf{B}}_1 \\ |
559 |
\hat{\mathbf{B}}_2 \\ |
560 |
\hat{\mathbf{B}}_3 |
561 |
\end{pmatrix} |
562 |
\end{eqnarray} |
563 |
% |
564 |
These matrices convert from space-fixed $(xyz)$ to body-fixed $(123)$ |
565 |
coordinates. |
566 |
|
567 |
Allen and Germano,\cite{Allen:2006fk} following earlier work by Price |
568 |
{\em et al.},\cite{Price:1984fk} showed that if the interaction |
569 |
energies are written explicitly in terms of $\hat{\mathbf{r}}$ and the body |
570 |
axes ($\hat{\mathbf{A}}_m$, $\hat{\mathbf{B}}_n$) : |
571 |
% |
572 |
\begin{equation} |
573 |
U(r, \{\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}} \}, |
574 |
\{\hat{\mathbf{B}}_n\cdot \hat{\mathbf{r}} \}, |
575 |
\{\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n \}) . |
576 |
\label{ugeneral} |
577 |
\end{equation} |
578 |
% |
579 |
the forces come out relatively cleanly, |
580 |
% |
581 |
\begin{equation} |
582 |
\mathbf{F}_a=-\mathbf{F}_b = \nabla U = |
583 |
\frac{\partial U}{\partial \mathbf{r}} |
584 |
= \frac{\partial U}{\partial r} \hat{\mathbf{r}} |
585 |
+ \sum_m \left[ |
586 |
\frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})} |
587 |
\frac { \partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}} |
588 |
+ \frac{\partial U}{\partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})} |
589 |
\frac { \partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}} |
590 |
\right] \label{forceequation}. |
591 |
\end{equation} |
592 |
|
593 |
The torques can also be found in a relatively similar |
594 |
manner, |
595 |
% |
596 |
\begin{eqnarray} |
597 |
\mathbf{\tau}_a = |
598 |
\sum_m |
599 |
\frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})} |
600 |
( \hat{\mathbf{r}} \times \hat{\mathbf{A}}_m ) |
601 |
-\sum_{mn} |
602 |
\frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n)} |
603 |
(\hat{\mathbf{A}}_m \times \hat{\mathbf{B}}_n) \\ |
604 |
% |
605 |
\mathbf{\tau}_b = |
606 |
\sum_m |
607 |
\frac{\partial U}{\partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})} |
608 |
( \hat{\mathbf{r}} \times \hat{\mathbf{B}}_m) |
609 |
+\sum_{mn} |
610 |
\frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n)} |
611 |
(\hat{\mathbf{A}}_m \times \hat{\mathbf{B}}_n) . |
612 |
\end{eqnarray} |
613 |
|
614 |
Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $ |
615 |
is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk} |
616 |
We also made use of the identities, |
617 |
% |
618 |
\begin{align} |
619 |
\frac { \partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}} |
620 |
=& \frac{1}{r} \left( \hat{\mathbf{A}}_m - (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})\hat{\mathbf{r}} |
621 |
\right) \\ |
622 |
\frac { \partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}} |
623 |
=& \frac{1}{r} \left( \hat{\mathbf{B}}_m - (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})\hat{\mathbf{r}} |
624 |
\right). |
625 |
\end{align} |
626 |
|
627 |
Many of the multipole contractions required can be written in one of |
628 |
three equivalent forms using the unit vectors $\hat{\mathbf{r}}$, $\hat{\mathbf{A}}_m$, |
629 |
and $\hat{\mathbf{B}}_n$. In the torque expressions, it is useful to have the |
630 |
angular-dependent terms available in all three fashions, e.g. for the |
631 |
dipole-dipole contraction: |
632 |
% |
633 |
\begin{equation} |
634 |
\mathbf{D}_a \cdot \mathbf{D}_b |
635 |
= D_{a \alpha} D_{b \alpha} = |
636 |
\sum_{mn} D_{am} \hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n D_{bn}. |
637 |
\end{equation} |
638 |
% |
639 |
The first two forms are written using space coordinates. The first |
640 |
form is standard in the chemistry literature, while the second is |
641 |
expressed using implied summation notation. The third form shows |
642 |
explicit sums over body indices and the dot products now indicate |
643 |
contractions using space indices. |
644 |
|
645 |
In computing our force and torque expressions, we carried out most of |
646 |
the work in body coordinates, and have transformed the expressions |
647 |
back to space-frame coordinates, which are reported below. Interested |
648 |
readers may consult the supplemental information for this paper for |
649 |
the intermediate body-frame expressions. |
650 |
|
651 |
\subsection{The Self-Interaction \label{sec:selfTerm}} |
652 |
|
653 |
In addition to cutoff-sphere neutralization, the Wolf |
654 |
summation~\cite{Wolf99} and the damped shifted force (DSF) |
655 |
extension~\cite{Fennell:2006zl} also include self-interactions that |
656 |
are handled separately from the pairwise interactions between |
657 |
sites. The self-term is normally calculated via a single loop over all |
658 |
sites in the system, and is relatively cheap to evaluate. The |
659 |
self-interaction has contributions from two sources. |
660 |
|
661 |
First, the neutralization procedure within the cutoff radius requires |
662 |
a contribution from a charge opposite in sign, but equal in magnitude, |
663 |
to the central charge, which has been spread out over the surface of |
664 |
the cutoff sphere. For a system of undamped charges, the total |
665 |
self-term is |
666 |
\begin{equation} |
667 |
U_\textrm{self} = - \frac{1}{r_c} \sum_{a=1}^N C_a^{2}. |
668 |
\label{eq:selfTerm} |
669 |
\end{equation} |
670 |
|
671 |
Second, charge damping with the complementary error function is a |
672 |
partial analogy to the Ewald procedure which splits the interaction |
673 |
into real and reciprocal space sums. The real space sum is retained |
674 |
in the Wolf and DSF methods. The reciprocal space sum is first |
675 |
minimized by folding the largest contribution (the self-interaction) |
676 |
into the self-interaction from charge neutralization of the damped |
677 |
potential. The remainder of the reciprocal space portion is then |
678 |
discarded (as this contributes the largest computational cost and |
679 |
complexity to the Ewald sum). For a system containing only damped |
680 |
charges, the complete self-interaction can be written as |
681 |
\begin{equation} |
682 |
U_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} + |
683 |
\frac{\alpha}{\sqrt{\pi}} \right) |
684 |
\sum_{a=1}^N |
685 |
C_a^{2}. |
686 |
\label{eq:dampSelfTerm} |
687 |
\end{equation} |
688 |
|
689 |
The extension of DSF electrostatics to point multipoles requires |
690 |
treatment of the self-neutralization \textit{and} reciprocal |
691 |
contributions to the self-interaction for higher order multipoles. In |
692 |
this section we give formulae for these interactions up to quadrupolar |
693 |
order. |
694 |
|
695 |
The self-neutralization term is computed by taking the {\it |
696 |
non-shifted} kernel for each interaction, placing a multipole of |
697 |
equal magnitude (but opposite in polarization) on the surface of the |
698 |
cutoff sphere, and averaging over the surface of the cutoff sphere. |
699 |
Because the self term is carried out as a single sum over sites, the |
700 |
reciprocal-space portion is identical to half of the self-term |
701 |
obtained by Smith, and also by Aguado and Madden for the application |
702 |
of the Ewald sum to multipoles.\cite{Smith82,Smith98,Aguado03} For a |
703 |
given site which posesses a charge, dipole, and quadrupole, both types |
704 |
of contribution are given in table \ref{tab:tableSelf}. |
705 |
|
706 |
\begin{table*} |
707 |
\caption{\label{tab:tableSelf} Self-interaction contributions for |
708 |
site ($a$) that has a charge $(C_a)$, dipole |
709 |
$(\mathbf{D}_a)$, and quadrupole $(\mathsf{Q}_a)$}. |
710 |
\begin{ruledtabular} |
711 |
\begin{tabular}{lccc} |
712 |
Multipole order & Summed Quantity & Self-neutralization & Reciprocal \\ \hline |
713 |
Charge & $C_a^2$ & $-f(r_c)$ & $-\frac{\alpha}{\sqrt{\pi}}$ \\ |
714 |
Dipole & $|\mathbf{D}_a|^2$ & $\frac{1}{3} \left( h(r_c) + |
715 |
\frac{2 g(r_c)}{r_c} \right)$ & $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$\\ |
716 |
Quadrupole & $2 \mathsf{Q}_a:\mathsf{Q}_a + \text{Tr}(\mathsf{Q}_a)^2$ & |
717 |
$- \frac{1}{15} \left( t(r_c)+ \frac{4 s(r_c)}{r_c} \right)$ & |
718 |
$-\frac{4 \alpha^5}{5 \sqrt{\pi}}$ \\ |
719 |
Charge-Quadrupole & $-2 C_a \text{Tr}(\mathsf{Q}_a)$ & $\frac{1}{3} \left( |
720 |
h(r_c) + \frac{2 g(r_c)}{r_c} \right)$& $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$ \\ |
721 |
\end{tabular} |
722 |
\end{ruledtabular} |
723 |
\end{table*} |
724 |
|
725 |
For sites which simultaneously contain charges and quadrupoles, the |
726 |
self-interaction includes a cross-interaction between these two |
727 |
multipole orders. Symmetry prevents the charge-dipole and |
728 |
dipole-quadrupole interactions from contributing to the |
729 |
self-interaction. The functions that go into the self-neutralization |
730 |
terms, $g(r), h(r), s(r), \mathrm{~and~} t(r)$ are successive |
731 |
derivatives of the electrostatic kernel, $f(r)$ (either the undamped |
732 |
$1/r$ or the damped $B_0(r)=\mathrm{erfc}(\alpha r)/r$ function) that |
733 |
have been evaluated at the cutoff distance. For undamped |
734 |
interactions, $f(r_c) = 1/r_c$, $g(r_c) = -1/r_c^{2}$, and so on. For |
735 |
damped interactions, $f(r_c) = B_0(r_c)$, $g(r_c) = B_0'(r_c)$, and so |
736 |
on. Appendix \ref{SmithFunc} contains recursion relations that allow |
737 |
rapid evaluation of these derivatives. |
738 |
|
739 |
\section{Interaction energies, forces, and torques} |
740 |
The main result of this paper is a set of expressions for the |
741 |
energies, forces and torques (up to quadrupole-quadrupole order) that |
742 |
work for the Taylor-shifted, gradient-shifted, and shifted potential |
743 |
approximations. These expressions were derived using a set of generic |
744 |
radial functions. Without using the shifting approximations mentioned |
745 |
above, some of these radial functions would be identical, and the |
746 |
expressions coalesce into the familiar forms for unmodified |
747 |
multipole-multipole interactions. Table \ref{tab:tableenergy} maps |
748 |
between the generic functions and the radial functions derived for the |
749 |
three methods. The energy equations are written in terms of lab-frame |
750 |
representations of the dipoles, quadrupoles, and the unit vector |
751 |
connecting the two objects, |
752 |
|
753 |
% Energy in space coordinate form ---------------------------------------------------------------------------------------------- |
754 |
% |
755 |
% |
756 |
% u ca cb |
757 |
% |
758 |
\begin{align} |
759 |
U_{C_a C_b}(r)=& |
760 |
C_a C_b v_{01}(r) \label{uchch} |
761 |
\\ |
762 |
% |
763 |
% u ca db |
764 |
% |
765 |
U_{C_a \mathbf{D}_b}(r)=& |
766 |
C_a \left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right) v_{11}(r) |
767 |
\label{uchdip} |
768 |
\\ |
769 |
% |
770 |
% u ca qb |
771 |
% |
772 |
U_{C_a \mathsf{Q}_b}(r)=& C_a \Bigl[ \text{Tr}\mathsf{Q}_b |
773 |
v_{21}(r) + \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot |
774 |
\hat{\mathbf{r}} \right) v_{22}(r) \Bigr] |
775 |
\label{uchquad} |
776 |
\\ |
777 |
% |
778 |
% u da cb |
779 |
% |
780 |
%U_{D_{\bf a}C_{\bf b}}(r)=& |
781 |
%-\frac{C_{\bf b}}{4\pi \epsilon_0} |
782 |
%\left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right) v_{11}(r) \label{udipch} |
783 |
%\\ |
784 |
% |
785 |
% u da db |
786 |
% |
787 |
U_{\mathbf{D}_a \mathbf{D}_b}(r)=& |
788 |
-\Bigr[ \left( \mathbf{D}_a \cdot |
789 |
\mathbf{D}_b \right) v_{21}(r) |
790 |
+\left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right) |
791 |
\left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right) |
792 |
v_{22}(r) \Bigr] |
793 |
\label{udipdip} |
794 |
\\ |
795 |
% |
796 |
% u da qb |
797 |
% |
798 |
\begin{split} |
799 |
% 1 |
800 |
U_{\mathbf{D}_a \mathsf{Q}_b}(r) =& |
801 |
-\Bigl[ |
802 |
\text{Tr}\mathsf{Q}_b |
803 |
\left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right) |
804 |
+2 ( \mathbf{D}_a \cdot |
805 |
\mathsf{Q}_b \cdot \hat{\mathbf{r}} ) \Bigr] v_{31}(r) \\ |
806 |
% 2 |
807 |
&- \left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right) |
808 |
\left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}} \right) v_{32}(r) |
809 |
\label{udipquad} |
810 |
\end{split} |
811 |
\\ |
812 |
% |
813 |
% u qa cb |
814 |
% |
815 |
%U_{Q_{\bf a}C_{\bf b}}(r)=& |
816 |
%\frac{C_{\bf b }}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\bf a} v_{21}(r) |
817 |
%\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{22}(r) \Bigr] |
818 |
%\label{uquadch} |
819 |
%\\ |
820 |
% |
821 |
% u qa db |
822 |
% |
823 |
%\begin{split} |
824 |
%1 |
825 |
%U_{Q_{\bf a}D_{\bf b}}(r)=& |
826 |
%\frac{1}{4\pi \epsilon_0} \Bigl[ |
827 |
%\text{Tr}\mathbf{Q}_{\mathbf{a}} |
828 |
%\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right) |
829 |
%+2 ( \mathbf{D}_{\mathbf{b}} \cdot |
830 |
%\mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r)\\ |
831 |
% 2 |
832 |
%&+\frac{1}{4\pi \epsilon_0} |
833 |
%\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right) |
834 |
%\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{32}(r) |
835 |
%\label{uquaddip} |
836 |
%\end{split} |
837 |
%\\ |
838 |
% |
839 |
% u qa qb |
840 |
% |
841 |
\begin{split} |
842 |
%1 |
843 |
U_{\mathsf{Q}_a \mathsf{Q}_b}(r)=& |
844 |
\Bigl[ |
845 |
\text{Tr} \mathsf{Q}_a \text{Tr} \mathsf{Q}_b |
846 |
+2 |
847 |
\mathsf{Q}_a : \mathsf{Q}_b \Bigr] v_{41}(r) |
848 |
\\ |
849 |
% 2 |
850 |
&+\Bigl[ \text{Tr}\mathsf{Q}_a |
851 |
\left( \hat{\mathbf{r}} \cdot |
852 |
\mathsf{Q}_b \cdot \hat{\mathbf{r}} \right) |
853 |
+\text{Tr}\mathsf{Q}_b |
854 |
\left( \hat{\mathbf{r}} \cdot \mathsf{Q}_a |
855 |
\cdot \hat{\mathbf{r}} \right) +4 (\hat{\mathbf{r}} \cdot |
856 |
\mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
857 |
\Bigr] v_{42}(r) |
858 |
\\ |
859 |
% 4 |
860 |
&+ |
861 |
\left( \hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}} \right) |
862 |
\left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}} \right) v_{43}(r). |
863 |
\label{uquadquad} |
864 |
\end{split} |
865 |
\end{align} |
866 |
% |
867 |
Note that the energies of multipoles on site $b$ interacting |
868 |
with those on site $a$ can be obtained by swapping indices |
869 |
along with the sign of the intersite vector, $\hat{\mathbf{r}}$. |
870 |
|
871 |
% |
872 |
% |
873 |
% TABLE of radial functions ---------------------------------------------------------------------------------------------------------------- |
874 |
% |
875 |
|
876 |
\begin{sidewaystable} |
877 |
\caption{\label{tab:tableenergy}Radial functions used in the energy |
878 |
and torque equations. The $f, g, h, s, t, \mathrm{and~} u$ |
879 |
functions used in this table are defined in Appendices |
880 |
\ref{radialTSF} and \ref{radialGSF}. The gradient shifted (GSF) |
881 |
functions include the shifted potential (SP) |
882 |
contributions (\textit{cf.} Eqs. \ref{generic2} and |
883 |
\ref{eq:SP}).} |
884 |
\begin{tabular}{|c|c|l|l|l|} \hline |
885 |
Generic&Bare Coulomb&Taylor-Shifted (TSF)&Shifted Potential (SP)&Gradient-Shifted (GSF) |
886 |
\\ \hline |
887 |
% |
888 |
% |
889 |
% |
890 |
%Ch-Ch& |
891 |
$v_{01}(r)$ & |
892 |
$\frac{1}{r}$ & |
893 |
$f_0(r)$ & |
894 |
$f(r)-f(r_c)$ & |
895 |
SP $-(r-r_c)g(r_c)$ |
896 |
\\ |
897 |
% |
898 |
% |
899 |
% |
900 |
%Ch-Di& |
901 |
$v_{11}(r)$ & |
902 |
$-\frac{1}{r^2}$ & |
903 |
$g_1(r)$ & |
904 |
$g(r)-g(r_c)$ & |
905 |
SP $-(r-r_c)h(r_c)$ \\ |
906 |
% |
907 |
% |
908 |
% |
909 |
%Ch-Qu/Di-Di& |
910 |
$v_{21}(r)$ & |
911 |
$-\frac{1}{r^3} $ & |
912 |
$\frac{g_2(r)}{r} $ & |
913 |
$\frac{g(r)}{r}-\frac{g(r_c)}{r_c}$ & |
914 |
SP $-(r-r_c) \left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right)$ \\ |
915 |
% |
916 |
% |
917 |
% |
918 |
$v_{22}(r)$ & |
919 |
$\frac{3}{r^3} $ & |
920 |
$\left(-\frac{g_2(r)}{r} + h_2(r) \right)$ & |
921 |
$\left(-\frac{g(r)}{r}+h(r) \right) -\left(-\frac{g(r_c)}{r_c}+h(r_c) \right)$ |
922 |
& SP $-(r-r_c) \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\ |
923 |
% |
924 |
% |
925 |
% |
926 |
%Di-Qu & |
927 |
$v_{31}(r)$ & |
928 |
$\frac{3}{r^4} $ & |
929 |
$\left(-\frac{g_3(r)}{r^2} + \frac{h_3(r)}{r} \right)$ & |
930 |
$\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r}\right)-\left(-\frac{g(r_c)}{r_c^2}+\frac{h(r_c)}{r_c} \right)$ |
931 |
& SP $-(r-r_c) \left(\frac{2g(r_c)}{r_c^3}-\frac{2h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ \\ |
932 |
% |
933 |
% |
934 |
% |
935 |
$v_{32}(r)$ & |
936 |
$-\frac{15}{r^4} $ & |
937 |
$\left( \frac{3g_3(r)}{r^2} - \frac{3h_3(r)}{r} + s_3(r) \right)$ & |
938 |
$\left( \frac{3g(r)}{r^2} - \frac{3h(r)}{r} + s(r) \right)$& |
939 |
SP $-(r-r_c) \left( \frac{-6g(r_c)}{r_c^3}+\frac{6h(r_c)}{r_c^2}\right.$ \\ |
940 |
&&& $~~~-\left(\frac{3g(r_c)}{r_c^2} - \frac{3h(r_c)}{r_c} + s(r_c)\right)$ & |
941 |
$\phantom{SP-(r-r_c)}\left.-\frac{3s(r_c)}{r_c}+t(r_c) \right)$\\ |
942 |
% |
943 |
% |
944 |
% |
945 |
%Qu-Qu& |
946 |
$v_{41}(r)$ & |
947 |
$\frac{3}{r^5} $ & |
948 |
$\left(-\frac{g_4(r)}{r^3} +\frac{h_4(r)}{r^2} \right) $ & |
949 |
$\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)- \left(-\frac{g(r_c)}{r_c^3} + \frac{h(r_c)}{r_c^2} \right)$ & |
950 |
SP $-(r-r_c) \left( \frac{3g(r_c)}{r_c^4}-\frac{3h(r_c)}{r_c^3}+\frac{s(r_c)}{r_c^2} \right)$ |
951 |
\\ |
952 |
% 2 |
953 |
$v_{42}(r)$ & |
954 |
$- \frac{15}{r^5} $ & |
955 |
$\left( \frac{3g_4(r)}{r^3} - \frac{3h_4(r)}{r^2}+\frac{s_4(r)}{r} \right)$ & |
956 |
$\left( \frac{3g(r)}{r^3} - \frac{3h(r)}{r^2}+\frac{s(r)}{r} \right)$ & |
957 |
SP$-(r-r_c) \left(- \frac{9g(r_c)}{r_c^4}+\frac{9h(r_c)}{r_c^3}\right.$ \\ |
958 |
&&& $~~~-\left( \frac{3g(r_c)}{r_c^3} - \frac{3h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ & |
959 |
$\phantom{SP-(r-r_c)}\left. -\frac{4s(r_c)}{r_c^2} + \frac{t(r_c)}{r_c}\right)$\\ |
960 |
% 3 |
961 |
% |
962 |
% |
963 |
$v_{43}(r)$ & |
964 |
$ \frac{105}{r^5} $ & |
965 |
$\left(-\frac{15g_4(r)}{r^3}+\frac{15h_4(r)}{r^2}-\frac{6s_4(r)}{r} + t_4(r)\right) $ & |
966 |
$ \left(-\frac{15g(r)}{r^3} +\frac{15h(r)}{r^2}-\frac{6s(r)}{r}+t(r)\right) $ & |
967 |
SP $-(r-r_c)\left(\frac{45g(r_c)}{r_c^4}-\frac{45h(r_c)}{r_c^3}\right.$\\ |
968 |
&&& $~~~-\left(-\frac{15g(r_c)}{r_c^3}+\frac{15h(r_c)}{r_c^2}-\frac{6s(r_c)}{r_c}+ t(r_c)\right)$ & |
969 |
$\phantom{SP-(r-r_c)}\left.+\frac{21s(r_c)}{r_c^2}-\frac{6t(r_c)}{r_c}+u(r_c) \right)$\\ |
970 |
\hline |
971 |
\end{tabular} |
972 |
\end{sidewaystable} |
973 |
% |
974 |
% |
975 |
% FORCE TABLE of radial functions ---------------------------------------------------------------------------------------------------------------- |
976 |
% |
977 |
|
978 |
\begin{sidewaystable} |
979 |
\caption{\label{tab:tableFORCE}Radial functions used in the force |
980 |
equations. Gradient shifted (GSF) functions are constructed using the shifted |
981 |
potential (SP) functions. Some of these functions are simple |
982 |
modifications of the functions found in table \ref{tab:tableenergy}} |
983 |
\begin{tabular}{|c|c|l|l|l|} \hline |
984 |
Function&Definition&Taylor-Shifted (TSF)& Shifted Potential (SP) |
985 |
&Gradient-Shifted (GSF) |
986 |
\\ \hline |
987 |
% |
988 |
% |
989 |
% |
990 |
$w_a(r)$& |
991 |
$\frac{d v_{01}}{dr}$& |
992 |
$g_0(r)$& |
993 |
$g(r)$& |
994 |
SP $-g(r_c)$ \\ |
995 |
% |
996 |
% |
997 |
$w_b(r)$ & |
998 |
$\frac{d v_{11}}{dr} - \frac{v_{11}(r)}{r} $& |
999 |
$\left( -\frac{g_1(r)}{r}+h_1(r) \right)$ & |
1000 |
$h(r) - \frac{v_{11}(r)}{r} $ & |
1001 |
SP $- h(r_c)$ \\ |
1002 |
% |
1003 |
$w_c(r)$ & |
1004 |
$\frac{v_{11}(r)}{r}$ & |
1005 |
$\frac{g_1(r)}{r} $ & |
1006 |
$\frac{v_{11}(r)}{r}$& |
1007 |
$\frac{v_{11}(r)}{r}$\\ |
1008 |
% |
1009 |
% |
1010 |
$w_d(r)$& |
1011 |
$\frac{d v_{21}}{dr}$& |
1012 |
$\left( -\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right) $ & |
1013 |
$\left( -\frac{g(r)}{r^2} + \frac{h(r)}{r} \right)$ & |
1014 |
SP $-\left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right) $ \\ |
1015 |
% |
1016 |
$w_e(r)$ & |
1017 |
$\left(-\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right)$ & |
1018 |
$\frac{v_{22}(r)}{r}$ & |
1019 |
$\frac{v_{22}(r)}{r}$ & |
1020 |
$\frac{v_{22}(r)}{r}$ \\ |
1021 |
% |
1022 |
% |
1023 |
$w_f(r)$& |
1024 |
$\frac{d v_{22}}{dr} - \frac{2v_{22}(r)}{r}$& |
1025 |
$\left( \frac{3g_2(r)}{r^2}-\frac{3h_2(r)}{r}+s_2(r) \right)$ & |
1026 |
$ \left( \frac{g(r)}{r^2}-\frac{h(r)}{r}+s(r) \right) -\frac{2v_{22}(r)}{r}$& |
1027 |
SP $- \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\ |
1028 |
% |
1029 |
$w_g(r)$& |
1030 |
$\frac{v_{31}(r)}{r}$& |
1031 |
$ \left( -\frac{g_3(r)}{r^3}+\frac{h_3(r)}{r^2} \right)$& |
1032 |
$\frac{v_{31}(r)}{r}$& |
1033 |
$\frac{v_{31}(r)}{r}$\\ |
1034 |
% |
1035 |
$w_h(r)$ & |
1036 |
$\frac{d v_{31}}{dr} -\frac{v_{31}(r)}{r}$& |
1037 |
$\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ & |
1038 |
$ \left(\frac{2g(r)}{r^3} -\frac{2h(r)}{r^2} +\frac{s(r)}{r} \right) -\frac{v_{31}(r)}{r}$ & |
1039 |
SP $ - \left(\frac{2g(r_c)}{r_c^3} -\frac{2h(r_c)}{r_c^2} +\frac{s(r_c)}{r_c} \right) $ \\ |
1040 |
% 2 |
1041 |
$w_i(r)$ & |
1042 |
$\frac{v_{32}(r)}{r}$ & |
1043 |
$\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ & |
1044 |
$\frac{v_{32}(r)}{r}$& |
1045 |
$\frac{v_{32}(r)}{r}$\\ |
1046 |
% |
1047 |
$w_j(r)$ & |
1048 |
$\frac{d v_{32}}{dr} - \frac{3v_{32}}{r}$& |
1049 |
$\left(\frac{-15g_3(r)}{r^3} + \frac{15h_3(r)}{r^2} - \frac{6s_3(r)}{r} + t_3(r) \right) $ & |
1050 |
$\left(\frac{-6g(r)}{r^3} +\frac{6h(r)}{r^2} -\frac{3s(r)}{r} +t(r) \right) -\frac{3v_{32}}{r}$ & |
1051 |
SP $-\left(\frac{-6g(_cr)}{r_c^3} +\frac{6h(r_c)}{r_c^2} |
1052 |
-\frac{3s(r_c)}{r_c} +t(r_c) \right)$ \\ |
1053 |
% |
1054 |
$w_k(r)$ & |
1055 |
$\frac{d v_{41}}{dr} $ & |
1056 |
$\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ & |
1057 |
$\left(\frac{3g(r)}{r^4} -\frac{3h(r)}{r^3} +\frac{s(r)}{r^2} |
1058 |
\right)$ & |
1059 |
SP $-\left(\frac{3g(r_c)}{r_c^4} -\frac{3h(r_c)}{r_c^3} +\frac{s(r_c)}{r_c^2} \right)$ \\ |
1060 |
% |
1061 |
$w_l(r)$ & |
1062 |
$\frac{d v_{42}}{dr} -\frac{2v_{42}(r)}{r}$ & |
1063 |
$\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ & |
1064 |
$\left(-\frac{9g(r)}{r^4} +\frac{9h(r)}{r^3} -\frac{4s(r)}{r^2} |
1065 |
+\frac{t(r)}{r} \right) -\frac{2v_{42}(r)}{r}$& |
1066 |
SP$-\left(-\frac{9g(r_c)}{r_c^4} +\frac{9h(r_c)}{r_c^3} -\frac{4s(r_c)}{r_c^2} +\frac{t(r_c)}{r_c} \right)$\\ |
1067 |
% |
1068 |
$w_m(r)$ & |
1069 |
$\frac{d v_{43}}{dr} -\frac{4v_{43}(r)}{r}$& |
1070 |
$\left(\frac{105g_4(r)}{r^4} - \frac{105h_4(r)}{r^3} \right.$ & |
1071 |
$\left(\frac{45g(r)}{r^4} -\frac{45h(r)}{r^3} +\frac{21s(r)}{r^2}\right.$ & |
1072 |
SP $- \left(\frac{45g(r_c)}{r_c^4} -\frac{45h(r_c)}{r_c^3}\right.$ \\ |
1073 |
&& $~~~\left.+ \frac{45s_4(r)}{r^2} - \frac{10t_4(r)}{r} +u_4(r) \right)$ |
1074 |
& $~~~\left. -\frac{6t(r)}{r} +u(r) \right) -\frac{4v_{43}(r)}{r}$ & |
1075 |
$\phantom{SP-} \left.+\frac{21s(r_c)}{r_c^2} -\frac{6t(r_c)}{r_c} +u(r_c) \right) $\\ |
1076 |
% |
1077 |
$w_n(r)$ & |
1078 |
$\frac{v_{42}(r)}{r}$ & |
1079 |
$\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ & |
1080 |
$\frac{v_{42}(r)}{r}$& |
1081 |
$\frac{v_{42}(r)}{r}$\\ |
1082 |
% |
1083 |
$w_o(r)$ & |
1084 |
$\frac{v_{43}(r)}{r}$& |
1085 |
$\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ & |
1086 |
$\frac{v_{43}(r)}{r}$& |
1087 |
$\frac{v_{43}(r)}{r}$ \\ \hline |
1088 |
% |
1089 |
|
1090 |
\end{tabular} |
1091 |
\end{sidewaystable} |
1092 |
% |
1093 |
% |
1094 |
% |
1095 |
|
1096 |
\subsection{Forces} |
1097 |
The force on object $a$, $\mathbf{F}_a$, due to object |
1098 |
$b$ is the negative of the force on $b$ due to $a$. For |
1099 |
a simple charge-charge interaction, these forces will point along the |
1100 |
$\pm \hat{\mathbf{r}}$ directions, where $\mathbf{r}=\mathbf{r}_b - |
1101 |
\mathbf{r}_a $. Thus |
1102 |
% |
1103 |
\begin{equation} |
1104 |
F_{a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_a C_b}}{\partial r} |
1105 |
\quad \text{and} \quad F_{b \alpha} |
1106 |
= - \hat{r}_\alpha \frac{\partial U_{C_a C_b}} {\partial r} . |
1107 |
\end{equation} |
1108 |
% |
1109 |
We list below the force equations written in terms of lab-frame |
1110 |
coordinates. The radial functions used in the three methods are listed |
1111 |
in Table \ref{tab:tableFORCE} |
1112 |
% |
1113 |
%SPACE COORDINATES FORCE EQUATIONS |
1114 |
% |
1115 |
% ************************************************************************** |
1116 |
% f ca cb |
1117 |
% |
1118 |
\begin{align} |
1119 |
\mathbf{F}_{a {C_a} {C_b}} =& |
1120 |
C_a C_b w_a(r) \hat{\mathbf{r}} \\ |
1121 |
% |
1122 |
% |
1123 |
% |
1124 |
\mathbf{F}_{a {C_a} {\mathbf{D}_b} } =& |
1125 |
C_a \Bigl[ |
1126 |
\left( \hat{\mathbf{r}} \cdot \mathbf{D}_b \right) |
1127 |
w_b(r) \hat{\mathbf{r}} |
1128 |
+ \mathbf{D}_b w_c(r) \Bigr] \\ |
1129 |
% |
1130 |
% |
1131 |
% |
1132 |
\mathbf{F}_{a {C_a} {\mathsf{Q}_b}} =& |
1133 |
C_a \Bigr[ |
1134 |
\text{Tr}\mathsf{Q}_b w_d(r) \hat{\mathbf{r}} |
1135 |
+ 2 \mathsf{Q}_b \cdot \hat{\mathbf{r}} w_e(r) |
1136 |
+ \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}} |
1137 |
\right) w_f(r) \hat{\mathbf{r}} \Bigr] \\ |
1138 |
% |
1139 |
% |
1140 |
% |
1141 |
% \begin{equation} |
1142 |
% \mathbf{F}_{{\bf a}D_{\bf a}C_{\bf b}} = |
1143 |
% -C_{\bf{b}} \Bigl[ |
1144 |
% \left( \hat{r} \cdot \mathbf{D}_{\mathbf{a}} \right) w_b(r) \hat{r} |
1145 |
% + \mathbf{D}_{\mathbf{a}} w_c(r) \Bigr] |
1146 |
% \end{equation} |
1147 |
% |
1148 |
% |
1149 |
% |
1150 |
\begin{split} |
1151 |
\mathbf{F}_{a \mathbf{D}_a \mathbf{D}_b} =& |
1152 |
- \mathbf{D}_a \cdot \mathbf{D}_b w_d(r) \hat{\mathbf{r}} |
1153 |
+ \left( \mathbf{D}_a |
1154 |
\left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right) |
1155 |
+ \mathbf{D}_b \left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right) \right) w_e(r)\\ |
1156 |
% 2 |
1157 |
& - \left( \hat{\mathbf{r}} \cdot \mathbf{D}_a \right) |
1158 |
\left( \hat{\mathbf{r}} \cdot \mathbf{D}_b \right) w_f(r) \hat{\mathbf{r}} |
1159 |
\end{split}\\ |
1160 |
% |
1161 |
% |
1162 |
% |
1163 |
\begin{split} |
1164 |
\mathbf{F}_{a \mathbf{D}_a \mathsf{Q}_b} =& - \Bigl[ |
1165 |
\text{Tr}\mathsf{Q}_b \mathbf{ D}_a |
1166 |
+2 \mathbf{D}_a \cdot |
1167 |
\mathsf{Q}_b \Bigr] w_g(r) |
1168 |
- \Bigl[ |
1169 |
\text{Tr}\mathsf{Q}_b |
1170 |
\left( \hat{\mathbf{r}} \cdot \mathbf{D}_a \right) |
1171 |
+2 ( \mathbf{D}_a \cdot |
1172 |
\mathsf{Q}_b \cdot \hat{\mathbf{r}}) \Bigr] w_h(r) \hat{\mathbf{r}} \\ |
1173 |
% 3 |
1174 |
& - \Bigl[\mathbf{ D}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1175 |
+2 (\hat{\mathbf{r}} \cdot \mathbf{D}_a ) (\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \Bigr] |
1176 |
w_i(r) |
1177 |
% 4 |
1178 |
- |
1179 |
(\hat{\mathbf{r}} \cdot \mathbf{D}_a ) |
1180 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) w_j(r) \hat{\mathbf{r}} \end{split} \\ |
1181 |
% |
1182 |
% |
1183 |
% \begin{equation} |
1184 |
% \mathbf{F}_{{\bf a}Q_{\bf a}C_{\bf b}} = |
1185 |
% \frac{C_{\bf b }}{4\pi \epsilon_0} \Bigr[ |
1186 |
% \text{Tr}\mathbf{Q}_{\bf a} w_d(r) \hat{r} |
1187 |
% + 2 \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} w_e(r) |
1188 |
% + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) w_f(r) \hat{r} \Bigr] |
1189 |
% \end{equation} |
1190 |
% % |
1191 |
% \begin{equation} |
1192 |
% \begin{split} |
1193 |
% \mathbf{F}_{{\bf a}Q_{\bf a}D_{\bf b}} = |
1194 |
% &\frac{1}{4\pi \epsilon_0} \Bigl[ |
1195 |
% \text{Tr}\mathbf{Q}_{\mathbf{a}} \mathbf{D}_{\mathbf{b}} |
1196 |
% +2 \mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} \Bigr] w_g(r) |
1197 |
% % 2 |
1198 |
% + \frac{1}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}} |
1199 |
% (\hat{r} \cdot \mathbf{D}_{\mathbf{b}}) |
1200 |
% +2 (\mathbf{D}_{\mathbf{b}} \cdot |
1201 |
% \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] w_h(r) \hat{r} \\ |
1202 |
% % 3 |
1203 |
% &+ \frac{1}{4\pi \epsilon_0} \Bigl[ \mathbf{D}_{\mathbf{b}} |
1204 |
% (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) |
1205 |
% +2 (\hat{r} \cdot \mathbf{D}_{\mathbf{b}}) |
1206 |
% (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \Bigr] w_i(r) |
1207 |
% % 4 |
1208 |
% +\frac{1}{4\pi \epsilon_0} |
1209 |
% (\hat{r} \cdot \mathbf{D}_{\mathbf{b}}) |
1210 |
% (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) w_j(r) \hat{r} |
1211 |
% \end{split} |
1212 |
% \end{equation} |
1213 |
% |
1214 |
% |
1215 |
% |
1216 |
\begin{split} |
1217 |
\mathbf{F}_{a \mathsf{Q}_a \mathsf{Q}_b} =& |
1218 |
\Bigl[ |
1219 |
\text{Tr}\mathsf{Q}_a \text{Tr}\mathsf{Q}_b |
1220 |
+ 2 \mathsf{Q}_a : \mathsf{Q}_b \Bigr] w_k(r) \hat{\mathbf{r}} \\ |
1221 |
% 2 |
1222 |
&+ \Bigl[ |
1223 |
2\text{Tr}\mathsf{Q}_b (\hat{\mathbf{r}} \cdot \mathsf{Q}_a ) |
1224 |
+ 2\text{Tr}\mathsf{Q}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) |
1225 |
% 3 |
1226 |
+4 (\mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1227 |
+ 4(\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \mathsf{Q}_b) \Bigr] w_n(r) \\ |
1228 |
% 4 |
1229 |
&+ \Bigl[ |
1230 |
\text{Tr}\mathsf{Q}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1231 |
+ \text{Tr}\mathsf{Q}_b |
1232 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}}) |
1233 |
% 5 |
1234 |
+4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot |
1235 |
\mathsf{Q}_b \cdot \hat{\mathbf{r}}) \Bigr] w_l(r) \hat{\mathbf{r}} \\ |
1236 |
% |
1237 |
&+ \Bigl[ |
1238 |
+ 2 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a ) |
1239 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1240 |
%6 |
1241 |
+2 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}}) |
1242 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \Bigr] w_o(r) \\ |
1243 |
% 7 |
1244 |
&+ |
1245 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}}) |
1246 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) w_m(r) \hat{\mathbf{r}} \end{split} |
1247 |
\end{align} |
1248 |
Note that the forces for higher multipoles on site $a$ |
1249 |
interacting with those of lower order on site $b$ can be |
1250 |
obtained by swapping indices in the expressions above. |
1251 |
|
1252 |
% |
1253 |
% Torques SECTION ----------------------------------------------------------------------------------------- |
1254 |
% |
1255 |
\subsection{Torques} |
1256 |
|
1257 |
% |
1258 |
The torques for the three methods are given in space-frame |
1259 |
coordinates: |
1260 |
% |
1261 |
% |
1262 |
\begin{align} |
1263 |
\mathbf{\tau}_{b C_a \mathbf{D}_b} =& |
1264 |
C_a (\hat{\mathbf{r}} \times \mathbf{D}_b) v_{11}(r) \\ |
1265 |
% |
1266 |
% |
1267 |
% |
1268 |
\mathbf{\tau}_{b C_a \mathsf{Q}_b} =& |
1269 |
2C_a |
1270 |
\hat{\mathbf{r}} \times ( \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{22}(r) \\ |
1271 |
% |
1272 |
% |
1273 |
% |
1274 |
% \begin{equation} |
1275 |
% \mathbf{\tau}_{{\bf a}D_{\bf a}C_{\bf b}} = |
1276 |
% -\frac{C_{\bf b}}{4\pi \epsilon_0} |
1277 |
% (\hat{r} \times \mathbf{D}_{\mathbf{a}}) v_{11}(r) |
1278 |
% \end{equation} |
1279 |
% |
1280 |
% |
1281 |
% |
1282 |
\mathbf{\tau}_{a \mathbf{D}_a \mathbf{D}_b} =& |
1283 |
\mathbf{D}_a \times \mathbf{D}_b v_{21}(r) |
1284 |
% 2 |
1285 |
- |
1286 |
(\hat{\mathbf{r}} \times \mathbf{D}_a ) |
1287 |
(\hat{\mathbf{r}} \cdot \mathbf{D}_b ) v_{22}(r)\\ |
1288 |
% |
1289 |
% |
1290 |
% |
1291 |
% \begin{equation} |
1292 |
% \mathbf{\tau}_{{\bf b}D_{\bf a}D_{\bf b}} = |
1293 |
% -\frac{1}{4\pi \epsilon_0} \mathbf{D}_{\mathbf {a}} \times \mathbf{D}_{\mathbf{b}} v_{21}(r) |
1294 |
% % 2 |
1295 |
% +\frac{1}{4\pi \epsilon_0} |
1296 |
% (\hat{r} \cdot \mathbf{D}_{\mathbf {a}} ) |
1297 |
% (\hat{r} \times \mathbf{D}_{\mathbf {b}} ) v_{22}(r) |
1298 |
% \end{equation} |
1299 |
% |
1300 |
% |
1301 |
% |
1302 |
\mathbf{\tau}_{a \mathbf{D}_a \mathsf{Q}_b} =& |
1303 |
\Bigl[ |
1304 |
-\text{Tr}\mathsf{Q}_b |
1305 |
(\hat{\mathbf{r}} \times \mathbf{D}_a ) |
1306 |
+2 \mathbf{D}_a \times |
1307 |
(\mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1308 |
\Bigr] v_{31}(r) |
1309 |
% 3 |
1310 |
- (\hat{\mathbf{r}} \times \mathbf{D}_a ) |
1311 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{32}(r)\\ |
1312 |
% |
1313 |
% |
1314 |
% |
1315 |
\mathbf{\tau}_{b \mathbf{D}_a \mathsf{Q}_b} =& |
1316 |
\Bigl[ |
1317 |
+2 ( \mathbf{D}_a \cdot \mathsf{Q}_b ) \times |
1318 |
\hat{\mathbf{r}} |
1319 |
-2 \mathbf{D}_a \times |
1320 |
(\mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1321 |
\Bigr] v_{31}(r) |
1322 |
% 2 |
1323 |
+ |
1324 |
(\hat{\mathbf{r}} \cdot \mathbf{D}_a) |
1325 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b) \times \hat{\mathbf{r}} v_{32}(r)\\ |
1326 |
% |
1327 |
% |
1328 |
% |
1329 |
% \begin{equation} |
1330 |
% \mathbf{\tau}_{{\bf a}Q_{\bf a}D_{\bf b}} = |
1331 |
% \frac{1}{4\pi \epsilon_0} \Bigl[ |
1332 |
% -2 (\mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} ) \times \hat{r} |
1333 |
% +2 \mathbf{D}_{\mathbf{b}} \times |
1334 |
% (\mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) |
1335 |
% \Bigr] v_{31}(r) |
1336 |
% % 3 |
1337 |
% - \frac{2}{4\pi \epsilon_0} |
1338 |
% (\hat{r} \cdot \mathbf{D}_{\mathbf{b}} ) |
1339 |
% (\hat{r} \cdot \mathbf |
1340 |
% {Q}_{{\mathbf a}}) \times \hat{r} v_{32}(r) |
1341 |
% \end{equation} |
1342 |
% |
1343 |
% |
1344 |
% |
1345 |
% \begin{equation} |
1346 |
% \mathbf{\tau}_{{\bf b}Q_{\bf a}D_{\bf b}} = |
1347 |
% \frac{1}{4\pi \epsilon_0} \Bigl[ |
1348 |
% \text{Tr}\mathbf{Q}_{\mathbf{a}} |
1349 |
% (\hat{r} \times \mathbf{D}_{\mathbf{b}} ) |
1350 |
% +2 \mathbf{D}_{\mathbf{b}} \times |
1351 |
% ( \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r) |
1352 |
% % 2 |
1353 |
% +\frac{1}{4\pi \epsilon_0} |
1354 |
% (\hat{r} \times \mathbf{D}_{\mathbf{b}} ) |
1355 |
% (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) v_{32}(r) |
1356 |
% \end{equation} |
1357 |
% |
1358 |
% |
1359 |
% |
1360 |
\begin{split} |
1361 |
\mathbf{\tau}_{a \mathsf{Q}_a \mathsf{Q}_b} =& |
1362 |
-4 |
1363 |
\mathsf{Q}_a \times \mathsf{Q}_b |
1364 |
v_{41}(r) \\ |
1365 |
% 2 |
1366 |
&+ |
1367 |
\Bigl[-2\text{Tr}\mathsf{Q}_b |
1368 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_a ) \times \hat{\mathbf{r}} |
1369 |
+4 \hat{\mathbf{r}} \times |
1370 |
( \mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1371 |
% 3 |
1372 |
-4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a )\times |
1373 |
( \mathsf{Q}_b \cdot \hat{\mathbf{r}} ) \Bigr] v_{42}(r) \\ |
1374 |
% 4 |
1375 |
&+ 2 |
1376 |
\hat{\mathbf{r}} \times ( \mathsf{Q}_a \cdot \hat{\mathbf{r}}) |
1377 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{43}(r) \end{split}\\ |
1378 |
% |
1379 |
% |
1380 |
% |
1381 |
\begin{split} |
1382 |
\mathbf{\tau}_{b \mathsf{Q}_a \mathsf{Q}_b} = |
1383 |
&4 |
1384 |
\mathsf{Q}_a \times \mathsf{Q}_b v_{41}(r) \\ |
1385 |
% 2 |
1386 |
&+ \Bigl[- 2\text{Tr}\mathsf{Q}_a |
1387 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \times \hat{\mathbf{r}} |
1388 |
-4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot |
1389 |
\mathsf{Q}_b ) \times |
1390 |
\hat{\mathbf{r}} |
1391 |
+4 ( \hat{\mathbf{r}} \cdot \mathsf{Q}_a ) \times |
1392 |
( \mathsf{Q}_b \cdot \hat{\mathbf{r}}) |
1393 |
\Bigr] v_{42}(r) \\ |
1394 |
% 4 |
1395 |
&+2 |
1396 |
(\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}}) |
1397 |
\hat{\mathbf{r}} \times ( \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{43}(r)\end{split} |
1398 |
\end{align} |
1399 |
% |
1400 |
Here, we have defined the matrix cross product in an identical form |
1401 |
as in Ref. \onlinecite{Smith98}: |
1402 |
\begin{equation} |
1403 |
\left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta |
1404 |
\left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta} |
1405 |
-\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta} |
1406 |
\right] |
1407 |
\label{eq:matrixCross} |
1408 |
\end{equation} |
1409 |
where $\alpha+1$ and $\alpha+2$ are regarded as cyclic |
1410 |
permuations of the matrix indices. |
1411 |
|
1412 |
All of the radial functions required for torques are identical with |
1413 |
the radial functions previously computed for the interaction energies. |
1414 |
These are tabulated for all three methods in table |
1415 |
\ref{tab:tableenergy}. The torques for higher multipoles on site |
1416 |
$a$ interacting with those of lower order on site |
1417 |
$b$ can be obtained by swapping indices in the expressions |
1418 |
above. |
1419 |
|
1420 |
\section{Comparison to known multipolar energies} |
1421 |
|
1422 |
To understand how these new real-space multipole methods behave in |
1423 |
computer simulations, it is vital to test against established methods |
1424 |
for computing electrostatic interactions in periodic systems, and to |
1425 |
evaluate the size and sources of any errors that arise from the |
1426 |
real-space cutoffs. In this paper we test SP, TSF, and GSF |
1427 |
electrostatics against analytical methods for computing the energies |
1428 |
of ordered multipolar arrays. In the following paper, we test the new |
1429 |
methods against the multipolar Ewald sum for computing the energies, |
1430 |
forces and torques for a wide range of typical condensed-phase |
1431 |
(disordered) systems. |
1432 |
|
1433 |
Because long-range electrostatic effects can be significant in |
1434 |
crystalline materials, ordered multipolar arrays present one of the |
1435 |
biggest challenges for real-space cutoff methods. The dipolar |
1436 |
analogues to the Madelung constants were first worked out by Sauer, |
1437 |
who computed the energies of ordered dipole arrays of zero |
1438 |
magnetization and obtained a number of these constants.\cite{Sauer} |
1439 |
This theory was developed more completely by Luttinger and |
1440 |
Tisza\cite{LT,LT2} who tabulated energy constants for the Sauer arrays |
1441 |
and other periodic structures. |
1442 |
|
1443 |
To test the new electrostatic methods, we have constructed very large, |
1444 |
$N=$ 16,000~(bcc) arrays of dipoles in the orientations described in |
1445 |
Ref. \onlinecite{LT}. These structures include ``A'' lattices with |
1446 |
nearest neighbor chains of antiparallel dipoles, as well as ``B'' |
1447 |
lattices with nearest neighbor strings of antiparallel dipoles if the |
1448 |
dipoles are contained in a plane perpendicular to the dipole direction |
1449 |
that passes through the dipole. We have also studied the minimum |
1450 |
energy structure for the BCC lattice that was found by Luttinger \& |
1451 |
Tisza. The total electrostatic energy density for any of the arrays |
1452 |
is given by: |
1453 |
\begin{equation} |
1454 |
E = C N^2 \mu^2 |
1455 |
\end{equation} |
1456 |
where $C$ is the energy constant (equivalent to the Madelung |
1457 |
constant), $N$ is the number of dipoles per unit volume, and $\mu$ is |
1458 |
the strength of the dipole. Energy constants (converged to 1 part in |
1459 |
$10^9$) are given in the supplemental information. |
1460 |
|
1461 |
\begin{figure} |
1462 |
\includegraphics[width=\linewidth]{Dipoles_rcut_threeAlpha.eps} |
1463 |
\caption{Convergence of the lattice energy constants as a function of |
1464 |
cutoff radius (normalized by the lattice constant, $a$) for the new |
1465 |
real-space methods. Three dipolar crystal structures were sampled, |
1466 |
and the analytic energy constants for the three lattices are |
1467 |
indicated with grey dashed lines. The left panel shows results for |
1468 |
the undamped kernel ($1/r$), while the damped kernel, $B_0(r)$ was |
1469 |
used in the center and right panels.} |
1470 |
\label{fig:Dipoles_rCut} |
1471 |
\end{figure} |
1472 |
|
1473 |
For the purposes of testing the energy expressions and the |
1474 |
self-neutralization schemes, the primary quantity of interest is the |
1475 |
analytic energy constant for the perfect arrays. Convergence to these |
1476 |
constants are shown as a function of the cutoff radius, $r_c$, for |
1477 |
three different values of the damping coefficient, $\alpha$ in |
1478 |
Fig.\ref{fig:Dipoles_rCut}. We have simultaneously tested a hard |
1479 |
cutoff (where the kernel is simply truncated at the cutoff radius) in |
1480 |
addition to the three new methods. |
1481 |
|
1482 |
The hard cutoff exhibits oscillations around the analytic energy |
1483 |
constants, and converges to incorrect energies when the complementary |
1484 |
error function damping kernel is used. The shifted potential (SP) |
1485 |
converges to the correct energy smoothly by $r_c = 4.5 a$ even for the |
1486 |
undamped case. This indicates that the shifting and the correction |
1487 |
provided by the self term are required for obtaining accurate |
1488 |
energies. The Taylor-shifted force (TSF) approximation appears to |
1489 |
perturb the potential too much inside the cutoff region to provide |
1490 |
accurate measures of the energy constants. GSF is a compromise, |
1491 |
converging to the correct energies within $r_c = 6 a$. |
1492 |
|
1493 |
{\it Quadrupolar} analogues to the Madelung constants were first |
1494 |
worked out by Nagai and Nakamura who computed the energies of selected |
1495 |
quadrupole arrays based on extensions to the Luttinger and Tisza |
1496 |
approach.\cite{Nagai01081960,Nagai01091963} |
1497 |
|
1498 |
In analogy to the dipolar arrays, the total electrostatic energy for |
1499 |
the quadrupolar arrays is: |
1500 |
\begin{equation} |
1501 |
E = C N \frac{3\bar{Q}^2}{4a^5} |
1502 |
\end{equation} |
1503 |
where $a$ is the lattice parameter, and $\bar{Q}$ is the effective |
1504 |
quadrupole moment, |
1505 |
\begin{equation} |
1506 |
\bar{Q}^2 = 2 \left(3 \mathsf{Q} : \mathsf{Q} - (\text{Tr} \mathsf{Q})^2 \right) |
1507 |
\end{equation} |
1508 |
for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}. |
1509 |
(For the traceless quadrupole tensor, $\mathsf{\Theta} = 3 \mathsf{Q} - \text{Tr} \mathsf{Q}$, |
1510 |
the effective moment, $\bar{Q}^2 = \frac{2}{3} \mathsf{\Theta} : \mathsf{\Theta}$.) |
1511 |
|
1512 |
To test the new electrostatic methods for quadrupoles, we have |
1513 |
constructed very large, $N=$ 8,000~(sc), 16,000~(bcc), and |
1514 |
32,000~(fcc) arrays of linear quadrupoles in the orientations |
1515 |
described in Ref. \onlinecite{Nagai01081960}. We have compared the |
1516 |
energy constants for these low-energy configurations for linear |
1517 |
quadrupoles. Convergence to these constants are shown as a function of |
1518 |
the cutoff radius, $r_c$, for three different values of the damping |
1519 |
parameter, $\alpha$ in Fig.~\ref{fig:Quadrupoles_rCut}. |
1520 |
|
1521 |
\begin{figure} |
1522 |
\includegraphics[width=\linewidth]{Quadrupoles_rcut_threeAlpha.eps} |
1523 |
\caption{Convergence of the lattice energy constants as a function of |
1524 |
cutoff radius (normalized by the lattice constant, $a$) for the new |
1525 |
real-space methods. Three quadrupolar crystal structures were |
1526 |
sampled, and the analytic energy constants for the three lattices |
1527 |
are indicated with grey dashed lines. The left panel shows results |
1528 |
for the undamped kernel ($1/r$), while the damped kernel, $B_0(r)$ |
1529 |
was used in the center and right panels. Note that for quadrupoles, |
1530 |
$\alpha^* = 0.4$ overdamps contributions from repulsive orientations |
1531 |
in the perfect crystal.} |
1532 |
\label{fig:Quadrupoles_rCut} |
1533 |
\end{figure} |
1534 |
|
1535 |
Again, we find that the hard cutoff exhibits oscillations around the |
1536 |
analytic energy constants. The shifted potential (SP) approximation |
1537 |
converges to the correct energy smoothly by $r_c = 3 a$ even for the |
1538 |
undamped case. The Taylor-shifted force (TSF) approximation again |
1539 |
appears to perturb the potential too much inside the cutoff region to |
1540 |
provide accurate measures of the energy constants. GSF again provides |
1541 |
a compromise between the two methods -- energies are converged by $r_c |
1542 |
= 4.5 a$, and the approximation is not as perturbative at short range |
1543 |
as TSF. |
1544 |
|
1545 |
It is also useful to understand the behavior of the lattice energy |
1546 |
constants for different values of the reduced damping parameter |
1547 |
($\alpha^* = \alpha a$) for the real-space methods. All of the methods |
1548 |
(except for TSF) have excellent behavior for the undamped or |
1549 |
weakly-damped cases. For the quadrupoles in particular, overdamping |
1550 |
can cause problems in perfect crystals ({\it cf.} the right panel in |
1551 |
Fig. \ref{fig:Quadrupoles_rCut}). In the perfect crystals, only a few |
1552 |
orientations are sampled, and the damping alters the radial function |
1553 |
for the direct quadrupolar contraction ($v_{41}(r)$) differently than |
1554 |
the radial functions for the terms involving the product of the |
1555 |
separation vector with the quadrupoles ($v_{42}(r)$ and $v_{43}(r)$). |
1556 |
Because these terms are altered by different amounts by the |
1557 |
complementary error function damping, the balance between attractive |
1558 |
and repulsive interactions in the crystal can be altered significantly |
1559 |
in overdamped situations. |
1560 |
|
1561 |
In the second paper in the series, we discuss how large values of |
1562 |
$\alpha$ can perturb the force and torque vectors, but weakly-damped |
1563 |
electrostatics appear to generate reasonable values for the total |
1564 |
electrostatic energies under both the SP and GSF approximations. We |
1565 |
also discuss the effects that $\alpha$ can have on convergence to the |
1566 |
average electrostatic energies in liquids (which sample a much wider |
1567 |
range of local orientations). |
1568 |
|
1569 |
\section{Conclusion} |
1570 |
We have presented three efficient real-space methods for computing the |
1571 |
interactions between point multipoles. One of these (SP) is a |
1572 |
multipolar generalization of Wolf's method that smoothly shifts |
1573 |
electrostatic energies to zero at the cutoff radius. Two of these |
1574 |
methods (GSF and TSF) also smoothly truncate the forces and torques |
1575 |
(in addition to the energies) at the cutoff radius, making them |
1576 |
attractive for both molecular dynamics and Monte Carlo simulations. We |
1577 |
find that the Gradient-Shifted Force (GSF) and the Shifted-Potential |
1578 |
(SP) methods converge rapidly to the correct lattice energies for |
1579 |
ordered dipolar and quadrupolar arrays, while the Taylor-Shifted Force |
1580 |
(TSF) is too severe an approximation to provide accurate convergence |
1581 |
to lattice energies. |
1582 |
|
1583 |
Although the TSF method appears to be |
1584 |
|
1585 |
|
1586 |
In most cases, GSF can obtain nearly quantitative agreement with the |
1587 |
lattice energy constants with reasonably small cutoff radii. The only |
1588 |
exception we have observed is for crystals which exhibit a bulk |
1589 |
macroscopic dipole moment (e.g. Luttinger \& Tisza's $Z_1$ lattice). |
1590 |
In this particular case, the multipole neutralization scheme can |
1591 |
interfere with the correct computation of the energies. We note that |
1592 |
the energies for these arrangements are typically much larger than for |
1593 |
crystals with net-zero moments, so this is not expected to be an issue |
1594 |
in most simulations. |
1595 |
|
1596 |
The techniques used here to derive the force, torque and energy |
1597 |
expressions can be extended to higher order multipoles, although some |
1598 |
of the objects (e.g. the matrix cross product in |
1599 |
Eq. \ref{eq:matrixCross}) will need to be generalized for higher-rank |
1600 |
tensors. We also note that the definitions of the multipoles used |
1601 |
here are in a primitive form, and these need some care when comparing |
1602 |
with experiment or other computational techniques. |
1603 |
|
1604 |
In large systems, these new methods can be made to scale approximately |
1605 |
linearly with system size, and detailed comparisons with the Ewald sum |
1606 |
for a wide range of chemical environments follows in the second paper. |
1607 |
|
1608 |
\begin{acknowledgments} |
1609 |
JDG acknowledges helpful discussions with Christopher |
1610 |
Fennell. Support for this project was provided by the National |
1611 |
Science Foundation under grant CHE-1362211. Computational time was |
1612 |
provided by the Center for Research Computing (CRC) at the |
1613 |
University of Notre Dame. |
1614 |
\end{acknowledgments} |
1615 |
|
1616 |
\newpage |
1617 |
\appendix |
1618 |
|
1619 |
\section{Smith's $B_l(r)$ functions for damped-charge distributions} |
1620 |
\label{SmithFunc} |
1621 |
The following summarizes Smith's $B_l(r)$ functions and includes |
1622 |
formulas given in his appendix.\cite{Smith98} The first function |
1623 |
$B_0(r)$ is defined by |
1624 |
% |
1625 |
\begin{equation} |
1626 |
B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r} |
1627 |
\int_{\alpha r}^{\infty} \text{e}^{-s^2} ds . |
1628 |
\end{equation} |
1629 |
% |
1630 |
The first derivative of this function is |
1631 |
% |
1632 |
\begin{equation} |
1633 |
\frac{dB_0(r)}{dr}=-\frac{1}{r^2}\text{erfc}(\alpha r) |
1634 |
-\frac{2\alpha}{r\sqrt{\pi}}\text{e}^{-{\alpha}^2r^2} |
1635 |
\end{equation} |
1636 |
% |
1637 |
which can be used to define a function $B_1(r)$: |
1638 |
% |
1639 |
\begin{equation} |
1640 |
B_1(r)=-\frac{1}{r}\frac{dB_0(r)}{dr} |
1641 |
\end{equation} |
1642 |
% |
1643 |
In general, the recurrence relation, |
1644 |
\begin{equation} |
1645 |
B_l(r)=-\frac{1}{r}\frac{dB_{l-1}(r)}{dr} |
1646 |
= \frac{1}{r^2} \left[ (2l-1)B_{l-1}(r) + \frac {(2\alpha^2)^l}{\alpha \sqrt{\pi}} |
1647 |
\text{e}^{-{\alpha}^2r^2} |
1648 |
\right] , |
1649 |
\end{equation} |
1650 |
is very useful for building up higher derivatives. As noted by Smith, |
1651 |
it is possible to approximate the $B_l(r)$ functions, |
1652 |
% |
1653 |
\begin{equation} |
1654 |
B_l(r)=\frac{(2l)!}{l!2^lr^{2l+1}} - \frac {(2\alpha^2)^{l+1}}{(2l+1)\alpha \sqrt{\pi}} |
1655 |
+\text{O}(r) . |
1656 |
\end{equation} |
1657 |
\newpage |
1658 |
\section{The $r$-dependent factors for TSF electrostatics} |
1659 |
\label{radialTSF} |
1660 |
|
1661 |
Using the shifted damped functions $f_n(r)$ defined by: |
1662 |
% |
1663 |
\begin{equation} |
1664 |
f_n(r)= B_0(r) -\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} B_0^{(m)}(r_c) , |
1665 |
\end{equation} |
1666 |
% |
1667 |
where the superscript $(m)$ denotes the $m^\mathrm{th}$ derivative. In |
1668 |
this Appendix, we provide formulas for successive derivatives of this |
1669 |
function. (If there is no damping, then $B_0(r)$ is replaced by |
1670 |
$1/r$.) First, we find: |
1671 |
% |
1672 |
\begin{equation} |
1673 |
\frac{\partial f_n}{\partial r_\alpha}=\hat{r}_\alpha \frac{d f_n}{d r} . |
1674 |
\end{equation} |
1675 |
% |
1676 |
This formula clearly brings in derivatives of Smith's $B_0(r)$ |
1677 |
function, and we define higher-order derivatives as follows: |
1678 |
% |
1679 |
\begin{align} |
1680 |
g_n(r)=& \frac{d f_n}{d r} = |
1681 |
B_0^{(1)}(r) -\sum_{m=0}^{n} \frac {(r-r_c)^m}{m!} B_0^{(m+1)}(r_c) \\ |
1682 |
h_n(r)=& \frac{d^2f_n}{d r^2} = |
1683 |
B_0^{(2)}(r) -\sum_{m=0}^{n-1} \frac {(r-r_c)^m}{m!} B_0^{(m+2)}(r_c) \\ |
1684 |
s_n(r)=& \frac{d^3f_n}{d r^3} = |
1685 |
B_0^{(3)}(r) -\sum_{m=0}^{n-2} \frac {(r-r_c)^m}{m!} B_0^{(m+3)}(r_c) \\ |
1686 |
t_n(r)=& \frac{d^4f_n}{d r^4} = |
1687 |
B_0^{(4)}(r) -\sum_{m=0}^{n-3} \frac {(r-r_c)^m}{m!} B_0^{(m+4)}(r_c) \\ |
1688 |
u_n(r)=& \frac{d^5f_n}{d r^5} = |
1689 |
B_0^{(5)}(r) -\sum_{m=0}^{n-4} \frac {(r-r_c)^m}{m!} B_0^{(m+5)}(r_c) . |
1690 |
\end{align} |
1691 |
% |
1692 |
We note that the last function needed (for quadrupole-quadrupole interactions) is |
1693 |
% |
1694 |
\begin{equation} |
1695 |
u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) . |
1696 |
\end{equation} |
1697 |
% The functions |
1698 |
% needed are listed schematically below: |
1699 |
% % |
1700 |
% \begin{eqnarray} |
1701 |
% f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\ |
1702 |
% g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\ |
1703 |
% h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\ |
1704 |
% s_2 \quad s_3 \quad &s_4 \nonumber \\ |
1705 |
% t_3 \quad &t_4 \nonumber \\ |
1706 |
% &u_4 \nonumber . |
1707 |
% \end{eqnarray} |
1708 |
The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and |
1709 |
stored on a grid for values of $r$ from $0$ to $r_c$. Using these |
1710 |
functions, we find |
1711 |
% |
1712 |
\begin{align} |
1713 |
\frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\ |
1714 |
\frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r} |
1715 |
+r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\ |
1716 |
\frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =& |
1717 |
\left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta + |
1718 |
\delta_{ \beta \gamma} r_\alpha \right) |
1719 |
\left( -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\ |
1720 |
& + r_\alpha r_\beta r_\gamma |
1721 |
\left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\ |
1722 |
\frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial |
1723 |
r_\gamma \partial r_\delta} =& |
1724 |
\left( \delta_{\alpha \beta} \delta_{\gamma \delta} |
1725 |
+ \delta_{\alpha \gamma} \delta_{\beta \delta} |
1726 |
+\delta_{ \beta \gamma} \delta_{\alpha \delta} \right) |
1727 |
\left( - \frac{g_n}{r^3} + \frac{h_n}{r^2} \right) \nonumber \\ |
1728 |
&+ \left( \delta_{\alpha \beta} r_\gamma r_\delta |
1729 |
+ \text{5 permutations} |
1730 |
\right) \left( \frac{3 g_n}{r^5} - \frac{3h_n}{r^4} + \frac{s_n}{r^3} |
1731 |
\right) \nonumber \\ |
1732 |
&+ r_\alpha r_\beta r_\gamma r_\delta |
1733 |
\left( -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5} |
1734 |
+ \frac{t_n}{r^4} \right)\\ |
1735 |
\frac{\partial^5 f_n} |
1736 |
{\partial r_\alpha \partial r_\beta \partial r_\gamma \partial |
1737 |
r_\delta \partial r_\epsilon} =& |
1738 |
\left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon |
1739 |
+ \text{14 permutations} \right) |
1740 |
\left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\ |
1741 |
&+ \left( \delta_{\alpha \beta} r_\gamma r_\delta r_\epsilon |
1742 |
+ \text{9 permutations} |
1743 |
\right) \left(- \frac{15g_n}{r^7}+\frac{15h_n}{r^7} -\frac{6s_n}{r^5} +\frac{t_n}{r^4} |
1744 |
\right) \nonumber \\ |
1745 |
&+ r_\alpha r_\beta r_\gamma r_\delta r_\epsilon |
1746 |
\left( \frac{105g_n}{r^9} - \frac{105h_n}{r^8} + \frac{45s_n}{r^7} |
1747 |
- \frac{10t_n}{r^6} +\frac{u_n}{r^5} \right) \label{eq:b13} |
1748 |
\end{align} |
1749 |
% |
1750 |
% |
1751 |
% |
1752 |
\newpage |
1753 |
\section{The $r$-dependent factors for GSF electrostatics} |
1754 |
\label{radialGSF} |
1755 |
|
1756 |
In Gradient-shifted force electrostatics, the kernel is not expanded, |
1757 |
and the expansion is carried out on the individual terms in the |
1758 |
multipole interaction energies. For damped charges, this still brings |
1759 |
multiple derivatives of the Smith's $B_0(r)$ function into the |
1760 |
algebra. To denote these terms, we generalize the notation of the |
1761 |
previous appendix. For either $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$ |
1762 |
(damped), |
1763 |
% |
1764 |
\begin{align} |
1765 |
g(r) &= \frac{df}{d r} && &&=-\frac{1}{r^2} |
1766 |
&&\mathrm{or~~~} -rB_1(r) \\ |
1767 |
h(r) &= \frac{dg}{d r} &&= \frac{d^2f}{d r^2} &&= \frac{2}{r^3} &&\mathrm{or~~~}-B_1(r) + r^2 B_2(r) \\ |
1768 |
s(r) &= \frac{dh}{d r} &&= \frac{d^3f}{d r^3} &&=-\frac{6}{r^4}&&\mathrm{or~~~}3rB_2(r) - r^3 B_3(r)\\ |
1769 |
t(r) &= \frac{ds}{d r} &&= \frac{d^4f}{d r^4} &&= \frac{24}{r^5} &&\mathrm{or~~~} 3 |
1770 |
B_2(r) - 6r^2 B_3(r) + r^4 B_4(r) \\ |
1771 |
u(r) &= \frac{dt}{d r} &&= \frac{d^5f}{d r^5} &&=-\frac{120}{r^6} &&\mathrm{or~~~} -15 |
1772 |
r B_3(r) + 10 r^3B_4(r) -r^5B_5(r). |
1773 |
\end{align} |
1774 |
% |
1775 |
For undamped charges, Table I lists these derivatives under the Bare |
1776 |
Coulomb column. Equations \ref{eq:b9} to \ref{eq:b13} are still |
1777 |
correct for GSF electrostatics if the subscript $n$ is eliminated. |
1778 |
|
1779 |
\newpage |
1780 |
|
1781 |
\bibliography{multipole} |
1782 |
|
1783 |
\end{document} |
1784 |
% |
1785 |
% ****** End of file multipole.tex ****** |