ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
Revision: 4213
Committed: Mon Aug 18 15:15:15 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 73113 byte(s)
Log Message:
Spelling errors

File Contents

# Content
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, Zahn {\it et al.} extended the Wolf
102 approach by adding a force-shifting term to the enhanced damped
103 Coulomb potential.\cite{Zahn02,Kast03} Fennell and Gezelter showed
104 that the simple damped shifted force (DSF) potential can give nearly
105 quantitative agreement with smooth particle mesh Ewald
106 (SPME)\cite{Essmann95} configurational energy differences as well as
107 atomic force and molecular torque vectors.\cite{Fennell:2006zl}
108
109 The computational efficiency and the accuracy of the DSF method are
110 surprisingly good, particularly for systems with uniform charge
111 density. Additionally, dielectric constants obtained using DSF and
112 similar methods where the force vanishes at $r_{c}$ are
113 essentially quantitative.\cite{Izvekov:2008wo} The DSF and other
114 related methods have now been widely investigated,\cite{Hansen:2012uq}
115 and DSF is now used routinely in a diverse set of chemical
116 environments.\cite{doi:10.1021/la400226g,McCann:2013fk,kannam:094701,Forrest:2012ly,English:2008kx,Louden:2013ve,Tokumasu:2013zr}
117 DSF electrostatics provides a compromise between the computational
118 speed of real-space cutoffs and the accuracy of fully-periodic Ewald
119 treatments.
120
121 One common feature of many coarse-graining approaches, which treat
122 entire molecular subsystems as a single rigid body, is simplification
123 of the electrostatic interactions between these bodies so that fewer
124 site-site interactions are required to compute configurational
125 energies. To do this, the interactions between coarse-grained sites
126 are typically taken to be point
127 multipoles.\cite{Golubkov06,ISI:000276097500009,ISI:000298664400012}
128
129 Water, in particular, has been modeled recently with point multipoles
130 up to octupolar
131 order.\cite{Chowdhuri:2006lr,Te:2010rt,Te:2010ys,Te:2010vn} For
132 maximum efficiency, these models require the use of an approximate
133 multipole expansion as the exact multipole expansion can become quite
134 expensive (particularly when handled via the Ewald
135 sum).\cite{Ichiye:2006qy} Point multipoles and multipole
136 polarizability have also been utilized in the AMOEBA water model and
137 related force fields.\cite{Ponder:2010fk,schnieders:124114,Ren:2011uq}
138
139 Higher-order multipoles present a peculiar issue for molecular
140 dynamics. Multipolar interactions are inherently short-ranged, and
141 should not need the relatively expensive Ewald treatment. However,
142 real-space cutoff methods are normally applied in an orientation-blind
143 fashion so multipoles which leave and then re-enter a cutoff sphere in
144 a different orientation can cause energy discontinuities.
145
146 This paper outlines an extension of the original DSF electrostatic
147 kernel to point multipoles. We describe three distinct real-space
148 interaction models for higher-order multipoles based on truncated
149 Taylor expansions that are carried out at the cutoff radius. We are
150 calling these models {\bf Taylor-shifted} (TSF), {\bf
151 gradient-shifted} (GSF) and {\bf shifted potential} (SP)
152 electrostatics. Because of differences in the initial assumptions,
153 the two methods yield related, but distinct expressions for energies,
154 forces, and torques.
155
156 In this paper we outline the new methodology and give functional forms
157 for the energies, forces, and torques up to quadrupole-quadrupole
158 order. We also compare the new methods to analytic energy constants
159 for periodic arrays of point multipoles. In the following paper, we
160 provide numerical comparisons to Ewald-based electrostatics in common
161 simulation environments.
162
163 \section{Methodology}
164 An efficient real-space electrostatic method involves the use of a
165 pair-wise functional form,
166 \begin{equation}
167 U = \sum_i \sum_{j>i} U_\mathrm{pair}(\mathbf{r}_{ij}, \mathsf{A}_i, \mathsf{B}_j) +
168 \sum_i U_i^\mathrm{self}
169 \end{equation}
170 that is short-ranged and easily truncated at a cutoff radius,
171 \begin{equation}
172 U_\mathrm{pair}(\mathbf{r}_{ij},\mathsf{A}_i, \mathsf{B}_j) = \left\{
173 \begin{array}{ll}
174 U_\mathrm{approx} (\mathbf{r}_{ij}, \mathsf{A}_i, \mathsf{B}_j) & \quad \left| \mathbf{r}_{ij} \right| \le r_c \\
175 0 & \quad \left| \mathbf{r}_{ij} \right| > r_c ,
176 \end{array}
177 \right.
178 \end{equation}
179 along with an easily computed self-interaction term ($\sum_i
180 U_i^\mathrm{self}$) which scales linearly with the number of
181 particles. Here $\mathsf{A}_i$ and $\mathsf{B}_j$ represent orientational
182 coordinates of the two sites (expressed as rotation matrices), and $\mathbf{r}_{ij}$ is the vector
183 between the two sites. The computational efficiency, energy
184 conservation, and even some physical properties of a simulation can
185 depend dramatically on how the $U_\mathrm{approx}$ function behaves at
186 the cutoff radius. The goal of any approximation method should be to
187 mimic the real behavior of the electrostatic interactions as closely
188 as possible without sacrificing the near-linear scaling of a cutoff
189 method.
190
191 \subsection{Self-neutralization, damping, and force-shifting}
192 The DSF and Wolf methods operate by neutralizing the total charge
193 contained within the cutoff sphere surrounding each particle. This is
194 accomplished by shifting the potential functions to generate image
195 charges on the surface of the cutoff sphere for each pair interaction
196 computed within $r_c$. Damping using a complementary error function is
197 applied to the potential to accelerate convergence. The interaction
198 for a pair of charges ($C_i$ and $C_j$) in the DSF method,
199 \begin{equation*}
200 U_\mathrm{DSF}(r_{ij}) = C_i C_j \Biggr{[}
201 \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
202 - \frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c} + \left(\frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c^2}
203 + \frac{2\alpha}{\pi^{1/2}}
204 \frac{\exp\left(-\alpha^2r_c^2\right)}{r_c}
205 \right)\left(r_{ij}-r_c\right)\ \Biggr{]}
206 \label{eq:DSFPot}
207 \end{equation*}
208 where $\alpha$ is the adjustable damping parameter. Note that in this
209 potential and in all electrostatic quantities that follow, the
210 standard $1/4 \pi \epsilon_{0}$ has been omitted for clarity.
211
212 To insure net charge neutrality within each cutoff sphere, an
213 additional ``self'' term is added to the potential. This term is
214 constant (as long as the charges and cutoff radius do not change), and
215 exists outside the normal pair-loop for molecular simulations. It can
216 be thought of as a contribution from a charge opposite in sign, but
217 equal in magnitude, to the central charge, which has been spread out
218 over the surface of the cutoff sphere. A portion of the self term is
219 identical to the self term in the Ewald summation, and comes from the
220 utilization of the complimentary error function for electrostatic
221 damping.\cite{deLeeuw80,Wolf99} There have also been recent efforts to
222 extend the Wolf self-neutralization method to zero out the dipole and
223 higher order multipoles contained within the cutoff
224 sphere.\cite{Fukuda:2011jk,Fukuda:2012yu,Fukuda:2013qv}
225
226 In this work, we extend the idea of self-neutralization for the point
227 multipoles by insuring net charge-neutrality and net-zero moments
228 within each cutoff sphere. In Figure \ref{fig:shiftedMultipoles}, point dipole
229 $\mathbf{D}_i$ at the center site $i$ is interacting with point dipole
230 $\mathbf{D}_j$ and point quadrupole $\mathsf{Q}_k$. The
231 self-neutralization scheme for point multipoles involves projecting
232 opposing multipoles for sites $j$ and $k$ on the surface of the cutoff
233 sphere. There are also significant modifications made to make the
234 forces and torques go smoothly to zero at the cutoff distance.
235
236 \begin{figure}
237 \includegraphics[width=3in]{SM.eps}
238 \caption{Reversed multipoles are projected onto the surface of the
239 cutoff sphere. The forces, torques, and potential are then smoothly
240 shifted to zero as the sites leave the cutoff region.}
241 \label{fig:shiftedMultipoles}
242 \end{figure}
243
244 As in the point-charge approach, there is an additional contribution
245 from self-neutralization of site $i$. The self term for multipoles is
246 described in section \ref{sec:selfTerm}.
247
248 \subsection{The multipole expansion}
249
250 Consider two discrete rigid collections of point charges, denoted as objects
251 $a$ and $b$. In the following, we assume that the two objects
252 interact via electrostatics only and describe those interactions in
253 terms of a standard multipole expansion. Putting the origin of the
254 coordinate system at the center of mass of $a$, we use vectors
255 $\mathbf{r}_k$ to denote the positions of all charges $q_k$ in
256 $a$. Then the electrostatic potential of object $a$ at
257 $\mathbf{r}$ is given by
258 \begin{equation}
259 \phi_a(\mathbf r) =
260 \sum_{k \, \text{in } a} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}.
261 \end{equation}
262 The Taylor expansion in $r$ can be written using an implied summation
263 notation. Here Greek indices are used to indicate space coordinates
264 ($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for
265 labeling specific sites for charges in $a$ and $b$ respectively. The
266 Taylor expansion,
267 \begin{equation}
268 \frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} =
269 \left( 1
270 - r_{k\alpha} \frac{\partial}{\partial r_{\alpha}}
271 + \frac{1}{2} r_{k\alpha} r_{k\beta} \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} +\dots
272 \right)
273 \frac{1}{r} ,
274 \end{equation}
275 can then be used to express the electrostatic potential on $a$ in
276 terms of multipole operators,
277 \begin{equation}
278 \phi_a(\mathbf{r}) =M_a \frac{1}{r}
279 \end{equation}
280 where
281 \begin{equation}
282 M_a = C_a - D_{a\alpha} \frac{\partial}{\partial r_{\alpha}}
283 + Q_{a\alpha\beta}
284 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
285 \end{equation}
286 Here, the point charge, dipole, and quadrupole for object $a$ are
287 given by $C_a$, $\mathbf{D}_a$, and $\mathsf{Q}_a$, respectively. These are the primitive multipoles
288 which can be expressed as a distribution of charges,
289 \begin{align}
290 C_a =&\sum_{k \, \text{in }a} q_k , \label{eq:charge} \\
291 D_{a\alpha} =&\sum_{k \, \text{in }a} q_k r_{k\alpha}, \label{eq:dipole}\\
292 Q_{a\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in } a} q_k
293 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
294 \end{align}
295 Note that the definition of the primitive quadrupole here differs from
296 the standard traceless form, and contains an additional Taylor-series
297 based factor of $1/2$. We are essentially treating the mass
298 distribution with higher priority; the moment of inertia tensor,
299 $\mathsf I$, is diagonalized to obtain body-fixed
300 axes, and the charge distribution may result in a quadrupole tensor
301 that is not necessarily diagonal in the body frame. Additional
302 reasons for utilizing the primitive quadrupole are discussed in
303 section \ref{sec:damped}.
304
305 It is convenient to locate charges $q_j$ relative to the center of mass of $b$. Then with $\bf{r}$ pointing from
306 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
307 \begin{equation}
308 U_{ab}(r)
309 = M_a \sum_{j \, \text{in } b} \frac {q_j}{\vert \mathbf{r}+\mathbf{r}_j \vert} .
310 \end{equation}
311 This can also be expanded as a Taylor series in $r$. Using a notation
312 similar to before to define the multipoles in object $b$,
313 \begin{equation}
314 M_b = C_b + D_{b\alpha} \frac{\partial}{\partial r_{\alpha}}
315 + Q_{b\alpha\beta}
316 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
317 \end{equation}
318 we arrive at the multipole expression for the total interaction energy.
319 \begin{equation}
320 U_{ab}(r)=M_a M_b \frac{1}{r} \label{kernel}.
321 \end{equation}
322 This form has the benefit of separating out the energies of
323 interaction into contributions from the charge, dipole, and quadrupole
324 of $a$ interacting with the same types of multipoles in $b$.
325
326 \subsection{Damped Coulomb interactions}
327 \label{sec:damped}
328 In the standard multipole expansion, one typically uses the bare
329 Coulomb potential, with radial dependence $1/r$, as shown in
330 Eq.~(\ref{kernel}). It is also quite common to use a damped Coulomb
331 interaction, which results from replacing point charges with Gaussian
332 distributions of charge with width $\alpha$. In damped multipole
333 electrostatics, the kernel ($1/r$) of the expansion is replaced with
334 the function:
335 \begin{equation}
336 B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}
337 \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
338 \end{equation}
339 We develop equations below using the function $f(r)$ to represent
340 either $1/r$ or $B_0(r)$, and all of the techniques can be applied to
341 bare or damped Coulomb kernels (or any other function) as long as
342 derivatives of these functions are known. Smith's convenient
343 functions $B_l(r)$, which are used for derivatives of the damped
344 kernel, are summarized in Appendix A. (N.B. there is one important
345 distinction between the two kernels, which is the behavior of
346 $\nabla^2 \frac{1}{r}$ compared with $\nabla^2 B_0(r)$. The former is
347 zero everywhere except for a delta function evaluated at the origin.
348 The latter also has delta function behavior, but is non-zero for $r
349 \neq 0$. Thus the standard justification for using a traceless
350 quadrupole tensor fails for the damped case.)
351
352 The main goal of this work is to smoothly cut off the interaction
353 energy as well as forces and torques as $r\rightarrow r_c$. To
354 describe how this goal may be met, we use two examples, charge-charge
355 and charge-dipole, using the bare Coulomb kernel, $f(r)=1/r$, to
356 explain the idea.
357
358 \subsection{Shifted-force methods}
359 In the shifted-force approximation, the interaction energy for two
360 charges $C_a$ and $C_b$ separated by a distance $r$ is
361 written:
362 \begin{equation}
363 U_{C_aC_b}(r)= C_a C_b
364 \left({ \frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} }
365 \right) .
366 \end{equation}
367 Two shifting terms appear in this equations, one from the
368 neutralization procedure ($-1/r_c$), and one that causes the first
369 derivative to vanish at the cutoff radius.
370
371 Since one derivative of the interaction energy is needed for the
372 force, the minimal perturbation is a term linear in $(r-r_c)$ in the
373 interaction energy, that is:
374 \begin{equation}
375 \frac{d\,}{dr}
376 \left( {\frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} }
377 \right) = \left(- \frac{1}{r^2} + \frac{1}{r_c^2}
378 \right) .
379 \end{equation}
380 which clearly vanishes as the $r$ approaches the cutoff radius. There
381 are a number of ways to generalize this derivative shift for
382 higher-order multipoles. Below, we present two methods, one based on
383 higher-order Taylor series for $r$ near $r_c$, and the other based on
384 linear shift of the kernel gradients at the cutoff itself.
385
386 \subsection{Taylor-shifted force (TSF) electrostatics}
387 In the Taylor-shifted force (TSF) method, the procedure that we follow
388 is based on a Taylor expansion containing the same number of
389 derivatives required for each force term to vanish at the cutoff. For
390 example, the quadrupole-quadrupole interaction energy requires four
391 derivatives of the kernel, and the force requires one additional
392 derivative. For quadrupole-quadrupole interactions, we therefore
393 require shifted energy expressions that include up to $(r-r_c)^5$ so
394 that all energies, forces, and torques are zero as $r \rightarrow
395 r_c$. In each case, we subtract off a function $f_n^{\text{shift}}(r)$
396 from the kernel $f(r)=1/r$. The subscript $n$ indicates the number of
397 derivatives to be taken when deriving a given multipole energy. We
398 choose a function with guaranteed smooth derivatives -- a truncated
399 Taylor series of the function $f(r)$, e.g.,
400 %
401 \begin{equation}
402 f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)}(r_c) .
403 \end{equation}
404 %
405 The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
406 Thus, for $f(r)=1/r$, we find
407 %
408 \begin{equation}
409 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} .
410 \end{equation}
411 %
412 Continuing with the example of a charge $a$ interacting with a
413 dipole $b$, we write
414 %
415 \begin{equation}
416 U_{C_a\mathbf{D}_b}(r)=
417 C_a D_{b\alpha} \frac {\partial f_1(r) }{\partial r_\alpha}
418 = C_a D_{b\alpha}
419 \frac {r_\alpha}{r} \frac {\partial f_1(r)}{\partial r} .
420 \end{equation}
421 %
422 The force that dipole $ b$ exerts on charge $a$ is
423 %
424 \begin{equation}
425 F_{C_a \mathbf{D}_b \beta} = C_a D_{b \alpha}
426 \left[ \frac{\delta_{\alpha\beta}}{r} \frac {\partial}{\partial r} +
427 \frac{r_\alpha r_\beta}{r^2}
428 \left( -\frac{1}{r} \frac {\partial} {\partial r}
429 + \frac {\partial ^2} {\partial r^2} \right) \right] f_1(r) .
430 \end{equation}
431 %
432 For undamped coulombic interactions, $f(r)=1/r$, we find
433 %
434 \begin{equation}
435 F_{C_a \mathbf{D}_b \beta} =
436 \frac{C_a D_{b\beta}}{r}
437 \left[ -\frac{1}{r^2}+\frac{1}{r_c^2}-\frac{2(r-r_c)}{r_c^3} \right]
438 +C_a D_{b \alpha}r_\alpha r_\beta
439 \left[ \frac{3}{r^5}-\frac{3}{r^3r_c^2} \right] .
440 \end{equation}
441 %
442 This expansion shows the expected $1/r^3$ dependence of the force.
443
444 In general, we can write
445 %
446 \begin{equation}
447 U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
448 \label{generic}
449 \end{equation}
450 %
451 with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
452 charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
453 $n=4$ for quadrupole-quadrupole. For example, in
454 quadrupole-quadrupole interactions for which the $\text{prefactor}$ is
455 $Q_{a \alpha\beta}Q_{b \gamma\delta}$, the derivatives are
456 $\partial^4/\partial r_\alpha \partial r_\beta \partial
457 r_\gamma \partial r_\delta$, with implied summation combining the
458 space indices. Appendix \ref{radialTSF} contains details on the
459 radial functions.
460
461 In the formulas presented in the tables below, the placeholder
462 function $f(r)$ is used to represent the electrostatic kernel (either
463 damped or undamped). The main functions that go into the force and
464 torque terms, $g_n(r), h_n(r), s_n(r), \mathrm{~and~} t_n(r)$ are
465 successive derivatives of the shifted electrostatic kernel, $f_n(r)$
466 of the same index $n$. The algebra required to evaluate energies,
467 forces and torques is somewhat tedious, so only the final forms are
468 presented in tables \ref{tab:tableenergy} and \ref{tab:tableFORCE}.
469 One of the principal findings of our work is that the individual
470 orientational contributions to the various multipole-multipole
471 interactions must be treated with distinct radial functions, but each
472 of these contributions is independently force shifted at the cutoff
473 radius.
474
475 \subsection{Gradient-shifted force (GSF) electrostatics}
476 The second, and conceptually simpler approach to force-shifting
477 maintains only the linear $(r-r_c)$ term in the truncated Taylor
478 expansion, and has a similar interaction energy for all multipole
479 orders:
480 \begin{equation}
481 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
482 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
483 \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
484 \label{generic2}
485 \end{equation}
486 where $\hat{\mathbf{r}}$ is the unit vector pointing between the two
487 multipoles, and the sum describes a separate force-shifting that is
488 applied to each orientational contribution to the energy. Both the
489 potential and the gradient for force shifting are evaluated for an
490 image multipole projected onto the surface of the cutoff sphere (see
491 fig \ref{fig:shiftedMultipoles}). The image multipole retains the
492 orientation (rotation matrix $\mathsf{B}$) of the interacting multipole. No
493 higher order terms $(r-r_c)^n$ appear. The primary difference between
494 the TSF and GSF methods is the stage at which the Taylor Series is
495 applied; in the Taylor-shifted approach, it is applied to the kernel
496 itself. In the Gradient-shifted approach, it is applied to individual
497 radial interaction terms in the multipole expansion. Energies from
498 this method thus have the general form:
499 \begin{equation}
500 U= \sum (\text{angular factor}) (\text{radial factor}).
501 \label{generic3}
502 \end{equation}
503
504 Functional forms for both methods (TSF and GSF) can both be summarized
505 using the form of Eq.~\ref{generic3}). The basic forms for the
506 energy, force, and torque expressions are tabulated for both shifting
507 approaches below -- for each separate orientational contribution, only
508 the radial factors differ between the two methods.
509
510 \subsection{Generalization of the Wolf shifted potential (SP)}
511 It is also possible to formulate an extension of the Wolf approach for
512 multipoles by simply projecting the image multipole onto the surface
513 of the cutoff sphere, and including the interactions with the central
514 multipole and the image. This effectively shifts the pair potential
515 to zero at the cutoff radius,
516 \begin{equation}
517 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
518 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
519 \label{eq:SP}
520 \end{equation}
521 independent of the orientations of the two multipoles. The sum again
522 describes separate potential shifting that is applied to each
523 orientational contribution to the energy.
524
525 The shifted potential (SP) method is a simple truncation of the GSF
526 method for each orientational contribution, leaving out the $(r-r_c)$
527 terms that multiply the gradient. Functional forms for the
528 shifted-potential (SP) method can also be summarized using the form of
529 Eq.~\ref{generic3}. The energy, force, and torque expressions are
530 tabulated below for all three methods. As in the GSF and TSF methods,
531 for each separate orientational contribution, only the radial factors
532 differ between the SP, GSF, and TSF methods.
533
534
535 \subsection{\label{sec:level2}Body and space axes}
536 Although objects $a$ and $b$ rotate during a molecular
537 dynamics (MD) simulation, their multipole tensors remain fixed in
538 body-frame coordinates. While deriving force and torque expressions,
539 it is therefore convenient to write the energies, forces, and torques
540 in intermediate forms involving the vectors of the rotation matrices.
541 We denote body axes for objects $a$ and $b$ using unit vectors
542 $\hat{\mathbf{A}}_m$ and $\hat{\mathbf{B}}_m$, respectively, with the index $m=(123)$.
543 In a typical simulation, the initial axes are obtained by
544 diagonalizing the moment of inertia tensors for the objects. (N.B.,
545 the body axes are generally {\it not} the same as those for which the
546 quadrupole moment is diagonal.) The rotation matrices are then
547 propagated during the simulation.
548
549 The rotation matrices $\mathsf {A}$ and $\mathsf {B}$ can be
550 expressed using these unit vectors:
551 \begin{eqnarray}
552 \mathsf {A} =
553 \begin{pmatrix}
554 \hat{\mathbf{A}}_1 \\
555 \hat{\mathbf{A}}_2 \\
556 \hat{\mathbf{A}}_3
557 \end{pmatrix}, \qquad
558 \mathsf {B} =
559 \begin{pmatrix}
560 \hat{\mathbf{B}}_1 \\
561 \hat{\mathbf{B}}_2 \\
562 \hat{\mathbf{B}}_3
563 \end{pmatrix}
564 \end{eqnarray}
565 %
566 These matrices convert from space-fixed $(xyz)$ to body-fixed $(123)$
567 coordinates.
568
569 Allen and Germano,\cite{Allen:2006fk} following earlier work by Price
570 {\em et al.},\cite{Price:1984fk} showed that if the interaction
571 energies are written explicitly in terms of $\hat{\mathbf{r}}$ and the body
572 axes ($\hat{\mathbf{A}}_m$, $\hat{\mathbf{B}}_n$) :
573 %
574 \begin{equation}
575 U(r, \{\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}} \},
576 \{\hat{\mathbf{B}}_n\cdot \hat{\mathbf{r}} \},
577 \{\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n \}) .
578 \label{ugeneral}
579 \end{equation}
580 %
581 the forces come out relatively cleanly,
582 %
583 \begin{equation}
584 \mathbf{F}_a=-\mathbf{F}_b = \nabla U =
585 \frac{\partial U}{\partial \mathbf{r}}
586 = \frac{\partial U}{\partial r} \hat{\mathbf{r}}
587 + \sum_m \left[
588 \frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}
589 \frac { \partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}}
590 + \frac{\partial U}{\partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}
591 \frac { \partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}}
592 \right] \label{forceequation}.
593 \end{equation}
594
595 The torques can also be found in a relatively similar
596 manner,
597 %
598 \begin{eqnarray}
599 \mathbf{\tau}_a =
600 \sum_m
601 \frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}
602 ( \hat{\mathbf{r}} \times \hat{\mathbf{A}}_m )
603 -\sum_{mn}
604 \frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n)}
605 (\hat{\mathbf{A}}_m \times \hat{\mathbf{B}}_n) \\
606 %
607 \mathbf{\tau}_b =
608 \sum_m
609 \frac{\partial U}{\partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}
610 ( \hat{\mathbf{r}} \times \hat{\mathbf{B}}_m)
611 +\sum_{mn}
612 \frac{\partial U}{\partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n)}
613 (\hat{\mathbf{A}}_m \times \hat{\mathbf{B}}_n) .
614 \end{eqnarray}
615
616 Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $
617 is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk}
618 We also made use of the identities,
619 %
620 \begin{align}
621 \frac { \partial (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}}
622 =& \frac{1}{r} \left( \hat{\mathbf{A}}_m - (\hat{\mathbf{A}}_m \cdot \hat{\mathbf{r}})\hat{\mathbf{r}}
623 \right) \\
624 \frac { \partial (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})}{\partial \mathbf{r}}
625 =& \frac{1}{r} \left( \hat{\mathbf{B}}_m - (\hat{\mathbf{B}}_m \cdot \hat{\mathbf{r}})\hat{\mathbf{r}}
626 \right).
627 \end{align}
628
629 Many of the multipole contractions required can be written in one of
630 three equivalent forms using the unit vectors $\hat{\mathbf{r}}$, $\hat{\mathbf{A}}_m$,
631 and $\hat{\mathbf{B}}_n$. In the torque expressions, it is useful to have the
632 angular-dependent terms available in all three fashions, e.g. for the
633 dipole-dipole contraction:
634 %
635 \begin{equation}
636 \mathbf{D}_a \cdot \mathbf{D}_b
637 = D_{a \alpha} D_{b \alpha} =
638 \sum_{mn} D_{am} \hat{\mathbf{A}}_m \cdot \hat{\mathbf{B}}_n D_{bn}.
639 \end{equation}
640 %
641 The first two forms are written using space coordinates. The first
642 form is standard in the chemistry literature, while the second is
643 expressed using implied summation notation. The third form shows
644 explicit sums over body indices and the dot products now indicate
645 contractions using space indices.
646
647 In computing our force and torque expressions, we carried out most of
648 the work in body coordinates, and have transformed the expressions
649 back to space-frame coordinates, which are reported below. Interested
650 readers may consult the supplemental information for this paper for
651 the intermediate body-frame expressions.\cite{supplemental}
652
653 \subsection{The Self-Interaction \label{sec:selfTerm}}
654
655 In addition to cutoff-sphere neutralization, the Wolf
656 summation~\cite{Wolf99} and the damped shifted force (DSF)
657 extension~\cite{Fennell:2006zl} also include self-interactions that
658 are handled separately from the pairwise interactions between
659 sites. The self-term is normally calculated via a single loop over all
660 sites in the system, and is relatively cheap to evaluate. The
661 self-interaction has contributions from two sources.
662
663 First, the neutralization procedure within the cutoff radius requires
664 a contribution from a charge opposite in sign, but equal in magnitude,
665 to the central charge, which has been spread out over the surface of
666 the cutoff sphere. For a system of undamped charges, the total
667 self-term is
668 \begin{equation}
669 U_\textrm{self} = - \frac{1}{r_c} \sum_{a=1}^N C_a^{2}.
670 \label{eq:selfTerm}
671 \end{equation}
672
673 Second, charge damping with the complementary error function is a
674 partial analogy to the Ewald procedure which splits the interaction
675 into real and reciprocal space sums. The real space sum is retained
676 in the Wolf and DSF methods. The reciprocal space sum is first
677 minimized by folding the largest contribution (the self-interaction)
678 into the self-interaction from charge neutralization of the damped
679 potential. The remainder of the reciprocal space portion is then
680 discarded (as this contributes the largest computational cost and
681 complexity to the Ewald sum). For a system containing only damped
682 charges, the complete self-interaction can be written as
683 \begin{equation}
684 U_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} +
685 \frac{\alpha}{\sqrt{\pi}} \right)
686 \sum_{a=1}^N
687 C_a^{2}.
688 \label{eq:dampSelfTerm}
689 \end{equation}
690
691 The extension of DSF electrostatics to point multipoles requires
692 treatment of the self-neutralization \textit{and} reciprocal
693 contributions to the self-interaction for higher order multipoles. In
694 this section we give formulae for these interactions up to quadrupolar
695 order.
696
697 The self-neutralization term is computed by taking the {\it
698 non-shifted} kernel for each interaction, placing a multipole of
699 equal magnitude (but opposite in polarization) on the surface of the
700 cutoff sphere, and averaging over the surface of the cutoff sphere.
701 Because the self term is carried out as a single sum over sites, the
702 reciprocal-space portion is identical to half of the self-term
703 obtained by Smith, and also by Aguado and Madden for the application
704 of the Ewald sum to multipoles.\cite{Smith82,Smith98,Aguado03} For a
705 given site which possesses a charge, dipole, and quadrupole, both types
706 of contribution are given in table \ref{tab:tableSelf}.
707
708 \begin{table*}
709 \caption{\label{tab:tableSelf} Self-interaction contributions for
710 site ($a$) that has a charge $(C_a)$, dipole
711 $(\mathbf{D}_a)$, and quadrupole $(\mathsf{Q}_a)$}.
712 \begin{ruledtabular}
713 \begin{tabular}{lccc}
714 Multipole order & Summed Quantity & Self-neutralization & Reciprocal \\ \hline
715 Charge & $C_a^2$ & $-f(r_c)$ & $-\frac{\alpha}{\sqrt{\pi}}$ \\
716 Dipole & $|\mathbf{D}_a|^2$ & $\frac{1}{3} \left( h(r_c) +
717 \frac{2 g(r_c)}{r_c} \right)$ & $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$\\
718 Quadrupole & $2 \mathsf{Q}_a:\mathsf{Q}_a + \text{Tr}(\mathsf{Q}_a)^2$ &
719 $- \frac{1}{15} \left( t(r_c)+ \frac{4 s(r_c)}{r_c} \right)$ &
720 $-\frac{4 \alpha^5}{5 \sqrt{\pi}}$ \\
721 Charge-Quadrupole & $-2 C_a \text{Tr}(\mathsf{Q}_a)$ & $\frac{1}{3} \left(
722 h(r_c) + \frac{2 g(r_c)}{r_c} \right)$& $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$ \\
723 \end{tabular}
724 \end{ruledtabular}
725 \end{table*}
726
727 For sites which simultaneously contain charges and quadrupoles, the
728 self-interaction includes a cross-interaction between these two
729 multipole orders. Symmetry prevents the charge-dipole and
730 dipole-quadrupole interactions from contributing to the
731 self-interaction. The functions that go into the self-neutralization
732 terms, $g(r), h(r), s(r), \mathrm{~and~} t(r)$ are successive
733 derivatives of the electrostatic kernel, $f(r)$ (either the undamped
734 $1/r$ or the damped $B_0(r)=\mathrm{erfc}(\alpha r)/r$ function) that
735 have been evaluated at the cutoff distance. For undamped
736 interactions, $f(r_c) = 1/r_c$, $g(r_c) = -1/r_c^{2}$, and so on. For
737 damped interactions, $f(r_c) = B_0(r_c)$, $g(r_c) = B_0'(r_c)$, and so
738 on. Appendix \ref{SmithFunc} contains recursion relations that allow
739 rapid evaluation of these derivatives.
740
741 \section{Interaction energies, forces, and torques}
742 The main result of this paper is a set of expressions for the
743 energies, forces and torques (up to quadrupole-quadrupole order) that
744 work for the Taylor-shifted, gradient-shifted, and shifted potential
745 approximations. These expressions were derived using a set of generic
746 radial functions. Without using the shifting approximations mentioned
747 above, some of these radial functions would be identical, and the
748 expressions coalesce into the familiar forms for unmodified
749 multipole-multipole interactions. Table \ref{tab:tableenergy} maps
750 between the generic functions and the radial functions derived for the
751 three methods. The energy equations are written in terms of lab-frame
752 representations of the dipoles, quadrupoles, and the unit vector
753 connecting the two objects,
754
755 % Energy in space coordinate form ----------------------------------------------------------------------------------------------
756 %
757 %
758 % u ca cb
759 %
760 \begin{align}
761 U_{C_a C_b}(r)=&
762 C_a C_b v_{01}(r) \label{uchch}
763 \\
764 %
765 % u ca db
766 %
767 U_{C_a \mathbf{D}_b}(r)=&
768 C_a \left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right) v_{11}(r)
769 \label{uchdip}
770 \\
771 %
772 % u ca qb
773 %
774 U_{C_a \mathsf{Q}_b}(r)=& C_a \Bigl[ \text{Tr}\mathsf{Q}_b
775 v_{21}(r) + \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot
776 \hat{\mathbf{r}} \right) v_{22}(r) \Bigr]
777 \label{uchquad}
778 \\
779 %
780 % u da cb
781 %
782 %U_{D_{\bf a}C_{\bf b}}(r)=&
783 %-\frac{C_{\bf b}}{4\pi \epsilon_0}
784 %\left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right) v_{11}(r) \label{udipch}
785 %\\
786 %
787 % u da db
788 %
789 U_{\mathbf{D}_a \mathbf{D}_b}(r)=&
790 -\Bigr[ \left( \mathbf{D}_a \cdot
791 \mathbf{D}_b \right) v_{21}(r)
792 +\left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right)
793 \left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right)
794 v_{22}(r) \Bigr]
795 \label{udipdip}
796 \\
797 %
798 % u da qb
799 %
800 \begin{split}
801 % 1
802 U_{\mathbf{D}_a \mathsf{Q}_b}(r) =&
803 -\Bigl[
804 \text{Tr}\mathsf{Q}_b
805 \left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right)
806 +2 ( \mathbf{D}_a \cdot
807 \mathsf{Q}_b \cdot \hat{\mathbf{r}} ) \Bigr] v_{31}(r) \\
808 % 2
809 &- \left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right)
810 \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}} \right) v_{32}(r)
811 \label{udipquad}
812 \end{split}
813 \\
814 %
815 % u qa cb
816 %
817 %U_{Q_{\bf a}C_{\bf b}}(r)=&
818 %\frac{C_{\bf b }}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\bf a} v_{21}(r)
819 %\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{22}(r) \Bigr]
820 %\label{uquadch}
821 %\\
822 %
823 % u qa db
824 %
825 %\begin{split}
826 %1
827 %U_{Q_{\bf a}D_{\bf b}}(r)=&
828 %\frac{1}{4\pi \epsilon_0} \Bigl[
829 %\text{Tr}\mathbf{Q}_{\mathbf{a}}
830 %\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
831 %+2 ( \mathbf{D}_{\mathbf{b}} \cdot
832 %\mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r)\\
833 % 2
834 %&+\frac{1}{4\pi \epsilon_0}
835 %\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
836 %\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{32}(r)
837 %\label{uquaddip}
838 %\end{split}
839 %\\
840 %
841 % u qa qb
842 %
843 \begin{split}
844 %1
845 U_{\mathsf{Q}_a \mathsf{Q}_b}(r)=&
846 \Bigl[
847 \text{Tr} \mathsf{Q}_a \text{Tr} \mathsf{Q}_b
848 +2
849 \mathsf{Q}_a : \mathsf{Q}_b \Bigr] v_{41}(r)
850 \\
851 % 2
852 &+\Bigl[ \text{Tr}\mathsf{Q}_a
853 \left( \hat{\mathbf{r}} \cdot
854 \mathsf{Q}_b \cdot \hat{\mathbf{r}} \right)
855 +\text{Tr}\mathsf{Q}_b
856 \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_a
857 \cdot \hat{\mathbf{r}} \right) +4 (\hat{\mathbf{r}} \cdot
858 \mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
859 \Bigr] v_{42}(r)
860 \\
861 % 4
862 &+
863 \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}} \right)
864 \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}} \right) v_{43}(r).
865 \label{uquadquad}
866 \end{split}
867 \end{align}
868 %
869 Note that the energies of multipoles on site $b$ interacting
870 with those on site $a$ can be obtained by swapping indices
871 along with the sign of the intersite vector, $\hat{\mathbf{r}}$.
872
873 %
874 %
875 % TABLE of radial functions ----------------------------------------------------------------------------------------------------------------
876 %
877
878 \begin{sidewaystable}
879 \caption{\label{tab:tableenergy}Radial functions used in the energy
880 and torque equations. The $f, g, h, s, t, \mathrm{and~} u$
881 functions used in this table are defined in Appendices
882 \ref{radialTSF} and \ref{radialGSF}. The gradient shifted (GSF)
883 functions include the shifted potential (SP)
884 contributions (\textit{cf.} Eqs. \ref{generic2} and
885 \ref{eq:SP}).}
886 \begin{tabular}{|c|c|l|l|l|} \hline
887 Generic&Bare Coulomb&Taylor-Shifted (TSF)&Shifted Potential (SP)&Gradient-Shifted (GSF)
888 \\ \hline
889 %
890 %
891 %
892 %Ch-Ch&
893 $v_{01}(r)$ &
894 $\frac{1}{r}$ &
895 $f_0(r)$ &
896 $f(r)-f(r_c)$ &
897 SP $-(r-r_c)g(r_c)$
898 \\
899 %
900 %
901 %
902 %Ch-Di&
903 $v_{11}(r)$ &
904 $-\frac{1}{r^2}$ &
905 $g_1(r)$ &
906 $g(r)-g(r_c)$ &
907 SP $-(r-r_c)h(r_c)$ \\
908 %
909 %
910 %
911 %Ch-Qu/Di-Di&
912 $v_{21}(r)$ &
913 $-\frac{1}{r^3} $ &
914 $\frac{g_2(r)}{r} $ &
915 $\frac{g(r)}{r}-\frac{g(r_c)}{r_c}$ &
916 SP $-(r-r_c) \left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right)$ \\
917 %
918 %
919 %
920 $v_{22}(r)$ &
921 $\frac{3}{r^3} $ &
922 $\left(-\frac{g_2(r)}{r} + h_2(r) \right)$ &
923 $\left(-\frac{g(r)}{r}+h(r) \right) -\left(-\frac{g(r_c)}{r_c}+h(r_c) \right)$
924 & SP $-(r-r_c) \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\
925 %
926 %
927 %
928 %Di-Qu &
929 $v_{31}(r)$ &
930 $\frac{3}{r^4} $ &
931 $\left(-\frac{g_3(r)}{r^2} + \frac{h_3(r)}{r} \right)$ &
932 $\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r}\right)-\left(-\frac{g(r_c)}{r_c^2}+\frac{h(r_c)}{r_c} \right)$
933 & SP $-(r-r_c) \left(\frac{2g(r_c)}{r_c^3}-\frac{2h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ \\
934 %
935 %
936 %
937 $v_{32}(r)$ &
938 $-\frac{15}{r^4} $ &
939 $\left( \frac{3g_3(r)}{r^2} - \frac{3h_3(r)}{r} + s_3(r) \right)$ &
940 $\left( \frac{3g(r)}{r^2} - \frac{3h(r)}{r} + s(r) \right)$&
941 SP $-(r-r_c) \left( \frac{-6g(r_c)}{r_c^3}+\frac{6h(r_c)}{r_c^2}\right.$ \\
942 &&& $~~~-\left(\frac{3g(r_c)}{r_c^2} - \frac{3h(r_c)}{r_c} + s(r_c)\right)$ &
943 $\phantom{SP-(r-r_c)}\left.-\frac{3s(r_c)}{r_c}+t(r_c) \right)$\\
944 %
945 %
946 %
947 %Qu-Qu&
948 $v_{41}(r)$ &
949 $\frac{3}{r^5} $ &
950 $\left(-\frac{g_4(r)}{r^3} +\frac{h_4(r)}{r^2} \right) $ &
951 $\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)- \left(-\frac{g(r_c)}{r_c^3} + \frac{h(r_c)}{r_c^2} \right)$ &
952 SP $-(r-r_c) \left( \frac{3g(r_c)}{r_c^4}-\frac{3h(r_c)}{r_c^3}+\frac{s(r_c)}{r_c^2} \right)$
953 \\
954 % 2
955 $v_{42}(r)$ &
956 $- \frac{15}{r^5} $ &
957 $\left( \frac{3g_4(r)}{r^3} - \frac{3h_4(r)}{r^2}+\frac{s_4(r)}{r} \right)$ &
958 $\left( \frac{3g(r)}{r^3} - \frac{3h(r)}{r^2}+\frac{s(r)}{r} \right)$ &
959 SP$-(r-r_c) \left(- \frac{9g(r_c)}{r_c^4}+\frac{9h(r_c)}{r_c^3}\right.$ \\
960 &&& $~~~-\left( \frac{3g(r_c)}{r_c^3} - \frac{3h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ &
961 $\phantom{SP-(r-r_c)}\left. -\frac{4s(r_c)}{r_c^2} + \frac{t(r_c)}{r_c}\right)$\\
962 % 3
963 %
964 %
965 $v_{43}(r)$ &
966 $ \frac{105}{r^5} $ &
967 $\left(-\frac{15g_4(r)}{r^3}+\frac{15h_4(r)}{r^2}-\frac{6s_4(r)}{r} + t_4(r)\right) $ &
968 $ \left(-\frac{15g(r)}{r^3} +\frac{15h(r)}{r^2}-\frac{6s(r)}{r}+t(r)\right) $ &
969 SP $-(r-r_c)\left(\frac{45g(r_c)}{r_c^4}-\frac{45h(r_c)}{r_c^3}\right.$\\
970 &&& $~~~-\left(-\frac{15g(r_c)}{r_c^3}+\frac{15h(r_c)}{r_c^2}-\frac{6s(r_c)}{r_c}+ t(r_c)\right)$ &
971 $\phantom{SP-(r-r_c)}\left.+\frac{21s(r_c)}{r_c^2}-\frac{6t(r_c)}{r_c}+u(r_c) \right)$\\
972 \hline
973 \end{tabular}
974 \end{sidewaystable}
975 %
976 %
977 % FORCE TABLE of radial functions ----------------------------------------------------------------------------------------------------------------
978 %
979
980 \begin{sidewaystable}
981 \caption{\label{tab:tableFORCE}Radial functions used in the force
982 equations. Gradient shifted (GSF) functions are constructed using the shifted
983 potential (SP) functions. Some of these functions are simple
984 modifications of the functions found in table \ref{tab:tableenergy}}
985 \begin{tabular}{|c|c|l|l|l|} \hline
986 Function&Definition&Taylor-Shifted (TSF)& Shifted Potential (SP)
987 &Gradient-Shifted (GSF)
988 \\ \hline
989 %
990 %
991 %
992 $w_a(r)$&
993 $\frac{d v_{01}}{dr}$&
994 $g_0(r)$&
995 $g(r)$&
996 SP $-g(r_c)$ \\
997 %
998 %
999 $w_b(r)$ &
1000 $\frac{d v_{11}}{dr} - \frac{v_{11}(r)}{r} $&
1001 $\left( -\frac{g_1(r)}{r}+h_1(r) \right)$ &
1002 $h(r) - \frac{v_{11}(r)}{r} $ &
1003 SP $- h(r_c)$ \\
1004 %
1005 $w_c(r)$ &
1006 $\frac{v_{11}(r)}{r}$ &
1007 $\frac{g_1(r)}{r} $ &
1008 $\frac{v_{11}(r)}{r}$&
1009 $\frac{v_{11}(r)}{r}$\\
1010 %
1011 %
1012 $w_d(r)$&
1013 $\frac{d v_{21}}{dr}$&
1014 $\left( -\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right) $ &
1015 $\left( -\frac{g(r)}{r^2} + \frac{h(r)}{r} \right)$ &
1016 SP $-\left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right) $ \\
1017 %
1018 $w_e(r)$ &
1019 $\left(-\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right)$ &
1020 $\frac{v_{22}(r)}{r}$ &
1021 $\frac{v_{22}(r)}{r}$ &
1022 $\frac{v_{22}(r)}{r}$ \\
1023 %
1024 %
1025 $w_f(r)$&
1026 $\frac{d v_{22}}{dr} - \frac{2v_{22}(r)}{r}$&
1027 $\left( \frac{3g_2(r)}{r^2}-\frac{3h_2(r)}{r}+s_2(r) \right)$ &
1028 $ \left( \frac{g(r)}{r^2}-\frac{h(r)}{r}+s(r) \right) -\frac{2v_{22}(r)}{r}$&
1029 SP $- \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\
1030 %
1031 $w_g(r)$&
1032 $\frac{v_{31}(r)}{r}$&
1033 $ \left( -\frac{g_3(r)}{r^3}+\frac{h_3(r)}{r^2} \right)$&
1034 $\frac{v_{31}(r)}{r}$&
1035 $\frac{v_{31}(r)}{r}$\\
1036 %
1037 $w_h(r)$ &
1038 $\frac{d v_{31}}{dr} -\frac{v_{31}(r)}{r}$&
1039 $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ &
1040 $ \left(\frac{2g(r)}{r^3} -\frac{2h(r)}{r^2} +\frac{s(r)}{r} \right) -\frac{v_{31}(r)}{r}$ &
1041 SP $ - \left(\frac{2g(r_c)}{r_c^3} -\frac{2h(r_c)}{r_c^2} +\frac{s(r_c)}{r_c} \right) $ \\
1042 % 2
1043 $w_i(r)$ &
1044 $\frac{v_{32}(r)}{r}$ &
1045 $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ &
1046 $\frac{v_{32}(r)}{r}$&
1047 $\frac{v_{32}(r)}{r}$\\
1048 %
1049 $w_j(r)$ &
1050 $\frac{d v_{32}}{dr} - \frac{3v_{32}}{r}$&
1051 $\left(\frac{-15g_3(r)}{r^3} + \frac{15h_3(r)}{r^2} - \frac{6s_3(r)}{r} + t_3(r) \right) $ &
1052 $\left(\frac{-6g(r)}{r^3} +\frac{6h(r)}{r^2} -\frac{3s(r)}{r} +t(r) \right) -\frac{3v_{32}}{r}$ &
1053 SP $-\left(\frac{-6g(_cr)}{r_c^3} +\frac{6h(r_c)}{r_c^2}
1054 -\frac{3s(r_c)}{r_c} +t(r_c) \right)$ \\
1055 %
1056 $w_k(r)$ &
1057 $\frac{d v_{41}}{dr} $ &
1058 $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ &
1059 $\left(\frac{3g(r)}{r^4} -\frac{3h(r)}{r^3} +\frac{s(r)}{r^2}
1060 \right)$ &
1061 SP $-\left(\frac{3g(r_c)}{r_c^4} -\frac{3h(r_c)}{r_c^3} +\frac{s(r_c)}{r_c^2} \right)$ \\
1062 %
1063 $w_l(r)$ &
1064 $\frac{d v_{42}}{dr} -\frac{2v_{42}(r)}{r}$ &
1065 $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1066 $\left(-\frac{9g(r)}{r^4} +\frac{9h(r)}{r^3} -\frac{4s(r)}{r^2}
1067 +\frac{t(r)}{r} \right) -\frac{2v_{42}(r)}{r}$&
1068 SP$-\left(-\frac{9g(r_c)}{r_c^4} +\frac{9h(r_c)}{r_c^3} -\frac{4s(r_c)}{r_c^2} +\frac{t(r_c)}{r_c} \right)$\\
1069 %
1070 $w_m(r)$ &
1071 $\frac{d v_{43}}{dr} -\frac{4v_{43}(r)}{r}$&
1072 $\left(\frac{105g_4(r)}{r^4} - \frac{105h_4(r)}{r^3} \right.$ &
1073 $\left(\frac{45g(r)}{r^4} -\frac{45h(r)}{r^3} +\frac{21s(r)}{r^2}\right.$ &
1074 SP $- \left(\frac{45g(r_c)}{r_c^4} -\frac{45h(r_c)}{r_c^3}\right.$ \\
1075 && $~~~\left.+ \frac{45s_4(r)}{r^2} - \frac{10t_4(r)}{r} +u_4(r) \right)$
1076 & $~~~\left. -\frac{6t(r)}{r} +u(r) \right) -\frac{4v_{43}(r)}{r}$ &
1077 $\phantom{SP-} \left.+\frac{21s(r_c)}{r_c^2} -\frac{6t(r_c)}{r_c} +u(r_c) \right) $\\
1078 %
1079 $w_n(r)$ &
1080 $\frac{v_{42}(r)}{r}$ &
1081 $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ &
1082 $\frac{v_{42}(r)}{r}$&
1083 $\frac{v_{42}(r)}{r}$\\
1084 %
1085 $w_o(r)$ &
1086 $\frac{v_{43}(r)}{r}$&
1087 $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1088 $\frac{v_{43}(r)}{r}$&
1089 $\frac{v_{43}(r)}{r}$ \\ \hline
1090 %
1091
1092 \end{tabular}
1093 \end{sidewaystable}
1094 %
1095 %
1096 %
1097
1098 \subsection{Forces}
1099 The force on object $a$, $\mathbf{F}_a$, due to object
1100 $b$ is the negative of the force on $b$ due to $a$. For
1101 a simple charge-charge interaction, these forces will point along the
1102 $\pm \hat{\mathbf{r}}$ directions, where $\mathbf{r}=\mathbf{r}_b -
1103 \mathbf{r}_a $. Thus
1104 %
1105 \begin{equation}
1106 F_{a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_a C_b}}{\partial r}
1107 \quad \text{and} \quad F_{b \alpha}
1108 = - \hat{r}_\alpha \frac{\partial U_{C_a C_b}} {\partial r} .
1109 \end{equation}
1110 %
1111 We list below the force equations written in terms of lab-frame
1112 coordinates. The radial functions used in the three methods are listed
1113 in Table \ref{tab:tableFORCE}
1114 %
1115 %SPACE COORDINATES FORCE EQUATIONS
1116 %
1117 % **************************************************************************
1118 % f ca cb
1119 %
1120 \begin{align}
1121 \mathbf{F}_{a {C_a} {C_b}} =&
1122 C_a C_b w_a(r) \hat{\mathbf{r}} \\
1123 %
1124 %
1125 %
1126 \mathbf{F}_{a {C_a} {\mathbf{D}_b} } =&
1127 C_a \Bigl[
1128 \left( \hat{\mathbf{r}} \cdot \mathbf{D}_b \right)
1129 w_b(r) \hat{\mathbf{r}}
1130 + \mathbf{D}_b w_c(r) \Bigr] \\
1131 %
1132 %
1133 %
1134 \mathbf{F}_{a {C_a} {\mathsf{Q}_b}} =&
1135 C_a \Bigr[
1136 \text{Tr}\mathsf{Q}_b w_d(r) \hat{\mathbf{r}}
1137 + 2 \mathsf{Q}_b \cdot \hat{\mathbf{r}} w_e(r)
1138 + \left( \hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}
1139 \right) w_f(r) \hat{\mathbf{r}} \Bigr] \\
1140 %
1141 %
1142 %
1143 % \begin{equation}
1144 % \mathbf{F}_{{\bf a}D_{\bf a}C_{\bf b}} =
1145 % -C_{\bf{b}} \Bigl[
1146 % \left( \hat{r} \cdot \mathbf{D}_{\mathbf{a}} \right) w_b(r) \hat{r}
1147 % + \mathbf{D}_{\mathbf{a}} w_c(r) \Bigr]
1148 % \end{equation}
1149 %
1150 %
1151 %
1152 \begin{split}
1153 \mathbf{F}_{a \mathbf{D}_a \mathbf{D}_b} =&
1154 - \mathbf{D}_a \cdot \mathbf{D}_b w_d(r) \hat{\mathbf{r}}
1155 + \left( \mathbf{D}_a
1156 \left( \mathbf{D}_b \cdot \hat{\mathbf{r}} \right)
1157 + \mathbf{D}_b \left( \mathbf{D}_a \cdot \hat{\mathbf{r}} \right) \right) w_e(r)\\
1158 % 2
1159 & - \left( \hat{\mathbf{r}} \cdot \mathbf{D}_a \right)
1160 \left( \hat{\mathbf{r}} \cdot \mathbf{D}_b \right) w_f(r) \hat{\mathbf{r}}
1161 \end{split}\\
1162 %
1163 %
1164 %
1165 \begin{split}
1166 \mathbf{F}_{a \mathbf{D}_a \mathsf{Q}_b} =& - \Bigl[
1167 \text{Tr}\mathsf{Q}_b \mathbf{ D}_a
1168 +2 \mathbf{D}_a \cdot
1169 \mathsf{Q}_b \Bigr] w_g(r)
1170 - \Bigl[
1171 \text{Tr}\mathsf{Q}_b
1172 \left( \hat{\mathbf{r}} \cdot \mathbf{D}_a \right)
1173 +2 ( \mathbf{D}_a \cdot
1174 \mathsf{Q}_b \cdot \hat{\mathbf{r}}) \Bigr] w_h(r) \hat{\mathbf{r}} \\
1175 % 3
1176 & - \Bigl[\mathbf{ D}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1177 +2 (\hat{\mathbf{r}} \cdot \mathbf{D}_a ) (\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \Bigr]
1178 w_i(r)
1179 % 4
1180 -
1181 (\hat{\mathbf{r}} \cdot \mathbf{D}_a )
1182 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) w_j(r) \hat{\mathbf{r}} \end{split} \\
1183 %
1184 %
1185 % \begin{equation}
1186 % \mathbf{F}_{{\bf a}Q_{\bf a}C_{\bf b}} =
1187 % \frac{C_{\bf b }}{4\pi \epsilon_0} \Bigr[
1188 % \text{Tr}\mathbf{Q}_{\bf a} w_d(r) \hat{r}
1189 % + 2 \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} w_e(r)
1190 % + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) w_f(r) \hat{r} \Bigr]
1191 % \end{equation}
1192 % %
1193 % \begin{equation}
1194 % \begin{split}
1195 % \mathbf{F}_{{\bf a}Q_{\bf a}D_{\bf b}} =
1196 % &\frac{1}{4\pi \epsilon_0} \Bigl[
1197 % \text{Tr}\mathbf{Q}_{\mathbf{a}} \mathbf{D}_{\mathbf{b}}
1198 % +2 \mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} \Bigr] w_g(r)
1199 % % 2
1200 % + \frac{1}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}}
1201 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1202 % +2 (\mathbf{D}_{\mathbf{b}} \cdot
1203 % \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] w_h(r) \hat{r} \\
1204 % % 3
1205 % &+ \frac{1}{4\pi \epsilon_0} \Bigl[ \mathbf{D}_{\mathbf{b}}
1206 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1207 % +2 (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1208 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \Bigr] w_i(r)
1209 % % 4
1210 % +\frac{1}{4\pi \epsilon_0}
1211 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1212 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) w_j(r) \hat{r}
1213 % \end{split}
1214 % \end{equation}
1215 %
1216 %
1217 %
1218 \begin{split}
1219 \mathbf{F}_{a \mathsf{Q}_a \mathsf{Q}_b} =&
1220 \Bigl[
1221 \text{Tr}\mathsf{Q}_a \text{Tr}\mathsf{Q}_b
1222 + 2 \mathsf{Q}_a : \mathsf{Q}_b \Bigr] w_k(r) \hat{\mathbf{r}} \\
1223 % 2
1224 &+ \Bigl[
1225 2\text{Tr}\mathsf{Q}_b (\hat{\mathbf{r}} \cdot \mathsf{Q}_a )
1226 + 2\text{Tr}\mathsf{Q}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b )
1227 % 3
1228 +4 (\mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1229 + 4(\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \mathsf{Q}_b) \Bigr] w_n(r) \\
1230 % 4
1231 &+ \Bigl[
1232 \text{Tr}\mathsf{Q}_a (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1233 + \text{Tr}\mathsf{Q}_b
1234 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}})
1235 % 5
1236 +4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot
1237 \mathsf{Q}_b \cdot \hat{\mathbf{r}}) \Bigr] w_l(r) \hat{\mathbf{r}} \\
1238 %
1239 &+ \Bigl[
1240 + 2 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a )
1241 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1242 %6
1243 +2 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}})
1244 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \Bigr] w_o(r) \\
1245 % 7
1246 &+
1247 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}})
1248 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) w_m(r) \hat{\mathbf{r}} \end{split}
1249 \end{align}
1250 Note that the forces for higher multipoles on site $a$
1251 interacting with those of lower order on site $b$ can be
1252 obtained by swapping indices in the expressions above.
1253
1254 %
1255 % Torques SECTION -----------------------------------------------------------------------------------------
1256 %
1257 \subsection{Torques}
1258
1259 %
1260 The torques for the three methods are given in space-frame
1261 coordinates:
1262 %
1263 %
1264 \begin{align}
1265 \mathbf{\tau}_{b C_a \mathbf{D}_b} =&
1266 C_a (\hat{\mathbf{r}} \times \mathbf{D}_b) v_{11}(r) \\
1267 %
1268 %
1269 %
1270 \mathbf{\tau}_{b C_a \mathsf{Q}_b} =&
1271 2C_a
1272 \hat{\mathbf{r}} \times ( \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{22}(r) \\
1273 %
1274 %
1275 %
1276 % \begin{equation}
1277 % \mathbf{\tau}_{{\bf a}D_{\bf a}C_{\bf b}} =
1278 % -\frac{C_{\bf b}}{4\pi \epsilon_0}
1279 % (\hat{r} \times \mathbf{D}_{\mathbf{a}}) v_{11}(r)
1280 % \end{equation}
1281 %
1282 %
1283 %
1284 \mathbf{\tau}_{a \mathbf{D}_a \mathbf{D}_b} =&
1285 \mathbf{D}_a \times \mathbf{D}_b v_{21}(r)
1286 % 2
1287 -
1288 (\hat{\mathbf{r}} \times \mathbf{D}_a )
1289 (\hat{\mathbf{r}} \cdot \mathbf{D}_b ) v_{22}(r)\\
1290 %
1291 %
1292 %
1293 % \begin{equation}
1294 % \mathbf{\tau}_{{\bf b}D_{\bf a}D_{\bf b}} =
1295 % -\frac{1}{4\pi \epsilon_0} \mathbf{D}_{\mathbf {a}} \times \mathbf{D}_{\mathbf{b}} v_{21}(r)
1296 % % 2
1297 % +\frac{1}{4\pi \epsilon_0}
1298 % (\hat{r} \cdot \mathbf{D}_{\mathbf {a}} )
1299 % (\hat{r} \times \mathbf{D}_{\mathbf {b}} ) v_{22}(r)
1300 % \end{equation}
1301 %
1302 %
1303 %
1304 \mathbf{\tau}_{a \mathbf{D}_a \mathsf{Q}_b} =&
1305 \Bigl[
1306 -\text{Tr}\mathsf{Q}_b
1307 (\hat{\mathbf{r}} \times \mathbf{D}_a )
1308 +2 \mathbf{D}_a \times
1309 (\mathsf{Q}_b \cdot \hat{\mathbf{r}})
1310 \Bigr] v_{31}(r)
1311 % 3
1312 - (\hat{\mathbf{r}} \times \mathbf{D}_a )
1313 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{32}(r)\\
1314 %
1315 %
1316 %
1317 \mathbf{\tau}_{b \mathbf{D}_a \mathsf{Q}_b} =&
1318 \Bigl[
1319 +2 ( \mathbf{D}_a \cdot \mathsf{Q}_b ) \times
1320 \hat{\mathbf{r}}
1321 -2 \mathbf{D}_a \times
1322 (\mathsf{Q}_b \cdot \hat{\mathbf{r}})
1323 \Bigr] v_{31}(r)
1324 % 2
1325 +
1326 (\hat{\mathbf{r}} \cdot \mathbf{D}_a)
1327 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b) \times \hat{\mathbf{r}} v_{32}(r)\\
1328 %
1329 %
1330 %
1331 % \begin{equation}
1332 % \mathbf{\tau}_{{\bf a}Q_{\bf a}D_{\bf b}} =
1333 % \frac{1}{4\pi \epsilon_0} \Bigl[
1334 % -2 (\mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} ) \times \hat{r}
1335 % +2 \mathbf{D}_{\mathbf{b}} \times
1336 % (\mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1337 % \Bigr] v_{31}(r)
1338 % % 3
1339 % - \frac{2}{4\pi \epsilon_0}
1340 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}} )
1341 % (\hat{r} \cdot \mathbf
1342 % {Q}_{{\mathbf a}}) \times \hat{r} v_{32}(r)
1343 % \end{equation}
1344 %
1345 %
1346 %
1347 % \begin{equation}
1348 % \mathbf{\tau}_{{\bf b}Q_{\bf a}D_{\bf b}} =
1349 % \frac{1}{4\pi \epsilon_0} \Bigl[
1350 % \text{Tr}\mathbf{Q}_{\mathbf{a}}
1351 % (\hat{r} \times \mathbf{D}_{\mathbf{b}} )
1352 % +2 \mathbf{D}_{\mathbf{b}} \times
1353 % ( \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r)
1354 % % 2
1355 % +\frac{1}{4\pi \epsilon_0}
1356 % (\hat{r} \times \mathbf{D}_{\mathbf{b}} )
1357 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) v_{32}(r)
1358 % \end{equation}
1359 %
1360 %
1361 %
1362 \begin{split}
1363 \mathbf{\tau}_{a \mathsf{Q}_a \mathsf{Q}_b} =&
1364 -4
1365 \mathsf{Q}_a \times \mathsf{Q}_b
1366 v_{41}(r) \\
1367 % 2
1368 &+
1369 \Bigl[-2\text{Tr}\mathsf{Q}_b
1370 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a ) \times \hat{\mathbf{r}}
1371 +4 \hat{\mathbf{r}} \times
1372 ( \mathsf{Q}_a \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1373 % 3
1374 -4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a )\times
1375 ( \mathsf{Q}_b \cdot \hat{\mathbf{r}} ) \Bigr] v_{42}(r) \\
1376 % 4
1377 &+ 2
1378 \hat{\mathbf{r}} \times ( \mathsf{Q}_a \cdot \hat{\mathbf{r}})
1379 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{43}(r) \end{split}\\
1380 %
1381 %
1382 %
1383 \begin{split}
1384 \mathbf{\tau}_{b \mathsf{Q}_a \mathsf{Q}_b} =
1385 &4
1386 \mathsf{Q}_a \times \mathsf{Q}_b v_{41}(r) \\
1387 % 2
1388 &+ \Bigl[- 2\text{Tr}\mathsf{Q}_a
1389 (\hat{\mathbf{r}} \cdot \mathsf{Q}_b ) \times \hat{\mathbf{r}}
1390 -4 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot
1391 \mathsf{Q}_b ) \times
1392 \hat{\mathbf{r}}
1393 +4 ( \hat{\mathbf{r}} \cdot \mathsf{Q}_a ) \times
1394 ( \mathsf{Q}_b \cdot \hat{\mathbf{r}})
1395 \Bigr] v_{42}(r) \\
1396 % 4
1397 &+2
1398 (\hat{\mathbf{r}} \cdot \mathsf{Q}_a \cdot \hat{\mathbf{r}})
1399 \hat{\mathbf{r}} \times ( \mathsf{Q}_b \cdot \hat{\mathbf{r}}) v_{43}(r)\end{split}
1400 \end{align}
1401 %
1402 Here, we have defined the matrix cross product in an identical form
1403 as in Ref. \onlinecite{Smith98}:
1404 \begin{equation}
1405 \left[\mathsf{A} \times \mathsf{B}\right]_\alpha = \sum_\beta
1406 \left[\mathsf{A}_{\alpha+1,\beta} \mathsf{B}_{\alpha+2,\beta}
1407 -\mathsf{A}_{\alpha+2,\beta} \mathsf{B}_{\alpha+1,\beta}
1408 \right]
1409 \label{eq:matrixCross}
1410 \end{equation}
1411 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic
1412 permuations of the matrix indices.
1413
1414 All of the radial functions required for torques are identical with
1415 the radial functions previously computed for the interaction energies.
1416 These are tabulated for all three methods in table
1417 \ref{tab:tableenergy}. The torques for higher multipoles on site
1418 $a$ interacting with those of lower order on site
1419 $b$ can be obtained by swapping indices in the expressions
1420 above.
1421
1422 \section{Comparison to known multipolar energies}
1423
1424 To understand how these new real-space multipole methods behave in
1425 computer simulations, it is vital to test against established methods
1426 for computing electrostatic interactions in periodic systems, and to
1427 evaluate the size and sources of any errors that arise from the
1428 real-space cutoffs. In this paper we test SP, TSF, and GSF
1429 electrostatics against analytical methods for computing the energies
1430 of ordered multipolar arrays. In the following paper, we test the new
1431 methods against the multipolar Ewald sum for computing the energies,
1432 forces and torques for a wide range of typical condensed-phase
1433 (disordered) systems.
1434
1435 Because long-range electrostatic effects can be significant in
1436 crystalline materials, ordered multipolar arrays present one of the
1437 biggest challenges for real-space cutoff methods. The dipolar
1438 analogues to the Madelung constants were first worked out by Sauer,
1439 who computed the energies of ordered dipole arrays of zero
1440 magnetization and obtained a number of these constants.\cite{Sauer}
1441 This theory was developed more completely by Luttinger and
1442 Tisza\cite{LT,LT2} who tabulated energy constants for the Sauer arrays
1443 and other periodic structures.
1444
1445 To test the new electrostatic methods, we have constructed very large,
1446 $N=$ 16,000~(bcc) arrays of dipoles in the orientations described in
1447 Ref. \onlinecite{LT}. These structures include ``A'' lattices with
1448 nearest neighbor chains of antiparallel dipoles, as well as ``B''
1449 lattices with nearest neighbor strings of antiparallel dipoles if the
1450 dipoles are contained in a plane perpendicular to the dipole direction
1451 that passes through the dipole. We have also studied the minimum
1452 energy structure for the BCC lattice that was found by Luttinger \&
1453 Tisza. The total electrostatic energy density for any of the arrays
1454 is given by:
1455 \begin{equation}
1456 E = C N^2 \mu^2
1457 \end{equation}
1458 where $C$ is the energy constant (equivalent to the Madelung
1459 constant), $N$ is the number of dipoles per unit volume, and $\mu$ is
1460 the strength of the dipole. Energy constants (converged to 1 part in
1461 $10^9$) are given in the supplemental information.\cite{supplemental}
1462
1463 \begin{figure}
1464 \includegraphics[width=\linewidth]{Dipoles_rcut_threeAlpha.eps}
1465 \caption{Convergence of the lattice energy constants as a function of
1466 cutoff radius (normalized by the lattice constant, $a$) for the new
1467 real-space methods. Three dipolar crystal structures were sampled,
1468 and the analytic energy constants for the three lattices are
1469 indicated with grey dashed lines. The left panel shows results for
1470 the undamped kernel ($1/r$), while the damped kernel, $B_0(r)$ was
1471 used in the center and right panels.}
1472 \label{fig:Dipoles_rCut}
1473 \end{figure}
1474
1475 For the purposes of testing the energy expressions and the
1476 self-neutralization schemes, the primary quantity of interest is the
1477 analytic energy constant for the perfect arrays. Convergence to these
1478 constants are shown as a function of the cutoff radius, $r_c$, for
1479 three different values of the damping coefficient, $\alpha$ in
1480 Fig.\ref{fig:Dipoles_rCut}. We have simultaneously tested a hard
1481 cutoff (where the kernel is simply truncated at the cutoff radius) in
1482 addition to the three new methods.
1483
1484 The hard cutoff exhibits oscillations around the analytic energy
1485 constants, and converges to incorrect energies when the complementary
1486 error function damping kernel is used. The shifted potential (SP)
1487 converges to the correct energy smoothly by $r_c = 4.5 a$ even for the
1488 undamped case. This indicates that the shifting and the correction
1489 provided by the self term are required for obtaining accurate
1490 energies. The Taylor-shifted force (TSF) approximation appears to
1491 perturb the potential too much inside the cutoff region to provide
1492 accurate measures of the energy constants. GSF is a compromise,
1493 converging to the correct energies within $r_c = 6 a$.
1494
1495 {\it Quadrupolar} analogues to the Madelung constants were first
1496 worked out by Nagai and Nakamura who computed the energies of selected
1497 quadrupole arrays based on extensions to the Luttinger and Tisza
1498 approach.\cite{Nagai01081960,Nagai01091963}
1499
1500 In analogy to the dipolar arrays, the total electrostatic energy for
1501 the quadrupolar arrays is:
1502 \begin{equation}
1503 E = C N \frac{3\bar{Q}^2}{4a^5}
1504 \end{equation}
1505 where $a$ is the lattice parameter, and $\bar{Q}$ is the effective
1506 quadrupole moment,
1507 \begin{equation}
1508 \bar{Q}^2 = 2 \left(3 \mathsf{Q} : \mathsf{Q} - (\text{Tr} \mathsf{Q})^2 \right)
1509 \end{equation}
1510 for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}.
1511 (For the traceless quadrupole tensor, $\mathsf{\Theta} = 3 \mathsf{Q} - \text{Tr} \mathsf{Q}$,
1512 the effective moment, $\bar{Q}^2 = \frac{2}{3} \mathsf{\Theta} : \mathsf{\Theta}$.)
1513
1514 To test the new electrostatic methods for quadrupoles, we have
1515 constructed very large, $N=$ 8,000~(sc), 16,000~(bcc), and
1516 32,000~(fcc) arrays of linear quadrupoles in the orientations
1517 described in Ref. \onlinecite{Nagai01081960}. We have compared the
1518 energy constants for these low-energy configurations for linear
1519 quadrupoles. Convergence to these constants are shown as a function of
1520 the cutoff radius, $r_c$, for three different values of the damping
1521 parameter, $\alpha$ in Fig.~\ref{fig:Quadrupoles_rCut}.
1522
1523 \begin{figure}
1524 \includegraphics[width=\linewidth]{Quadrupoles_rcut_threeAlpha.eps}
1525 \caption{Convergence of the lattice energy constants as a function of
1526 cutoff radius (normalized by the lattice constant, $a$) for the new
1527 real-space methods. Three quadrupolar crystal structures were
1528 sampled, and the analytic energy constants for the three lattices
1529 are indicated with grey dashed lines. The left panel shows results
1530 for the undamped kernel ($1/r$), while the damped kernel, $B_0(r)$
1531 was used in the center and right panels. Note that for quadrupoles,
1532 $\alpha^* = 0.4$ overdamps contributions from repulsive orientations
1533 in the perfect crystal.}
1534 \label{fig:Quadrupoles_rCut}
1535 \end{figure}
1536
1537 Again, we find that the hard cutoff exhibits oscillations around the
1538 analytic energy constants. The shifted potential (SP) approximation
1539 converges to the correct energy smoothly by $r_c = 3 a$ even for the
1540 undamped case. The Taylor-shifted force (TSF) approximation again
1541 appears to perturb the potential too much inside the cutoff region to
1542 provide accurate measures of the energy constants. GSF again provides
1543 a compromise between the two methods -- energies are converged by $r_c
1544 = 4.5 a$, and the approximation is not as perturbative at short range
1545 as TSF.
1546
1547 It is also useful to understand the behavior of the lattice energy
1548 constants for different values of the reduced damping parameter
1549 ($\alpha^* = \alpha a$) for the real-space methods. All of the methods
1550 (except for TSF) have excellent behavior for the undamped or
1551 weakly-damped cases. Overdamping can cause problems in perfect
1552 crystals for the quadrupoles in particular ({\it cf.} the right panel
1553 in Fig. \ref{fig:Quadrupoles_rCut}). In the perfect crystals, only a
1554 few orientations are being sampled. E.g. in the simple cubic (SC)
1555 lattice of linear quadrupoles aligned in the 111 direction, the
1556 nearest-neighbor quadrupoles only sample 3 distinct orientations
1557 relative to the vector between the sites. The damping alters the
1558 radial function for the direct quadrupolar contraction, $v_{41}(r)$,
1559 differently than the radial functions for the terms involving the
1560 product of the separation vector with the quadrupoles, $v_{42}(r)$ and
1561 $v_{43}(r)$. Because these terms are altered by different amounts by
1562 the complementary error function damping, the effect of damping is
1563 non-spherical for multipoles, and the balance between attractive and
1564 repulsive interactions in the crystal is therefore altered
1565 significantly in overdamped situations.
1566
1567 In the second paper in the series, we discuss how large values of
1568 $\alpha$ can perturb the force and torque vectors, but weakly-damped
1569 electrostatics appear to generate reasonable values for the total
1570 electrostatic energies under both the SP and GSF approximations. We
1571 also discuss the effects that $\alpha$ can have on convergence to the
1572 average electrostatic energies in liquids (which sample a much wider
1573 range of local orientations).
1574
1575 \section{Conclusion}
1576 We have presented three efficient real-space methods for computing the
1577 interactions between point multipoles. One of these (SP) is a
1578 multipolar generalization of Wolf's method that smoothly shifts
1579 electrostatic energies to zero at the cutoff radius. Two of these
1580 methods (GSF and TSF) also smoothly truncate the forces and torques
1581 (in addition to the energies) at the cutoff radius, making them
1582 attractive for both molecular dynamics and Monte Carlo simulations. We
1583 find that the Gradient-Shifted Force (GSF) and the Shifted-Potential
1584 (SP) methods converge rapidly to the correct lattice energies for
1585 ordered dipolar and quadrupolar arrays, while the Taylor-Shifted Force
1586 (TSF) is too severe an approximation to provide convergence to lattice
1587 energies within reasonable cutoff radii.
1588
1589 Although the TSF method appears to perform poorly for the analytical
1590 energy constants, the structure of the radial functions used in the
1591 force and torque expressions in the other two methods would not have
1592 been revealed without first developing the TSF approach. TSF also
1593 generates a set of electrostatic kernels that have multiple
1594 derivatives that vanish at the cutoff radius, a property that is
1595 valuable in estimating dielectric constants using the conducting
1596 boundary fluctuation formula.\cite{Izvekov:2008wo}
1597
1598 In most cases, GSF can obtain nearly quantitative agreement with the
1599 lattice energy constants with reasonably small cutoff radii. The only
1600 exception we have observed is for crystals which exhibit a bulk
1601 macroscopic dipole moment (e.g. Luttinger \& Tisza's $Z_1$ lattice).
1602 In this particular case, the multipole neutralization scheme can
1603 interfere with the correct computation of the energies. We note that
1604 the energies for these arrangements are typically much larger than for
1605 crystals with net-zero moments, so this is not expected to be an issue
1606 in most simulations.
1607
1608 Relatively weak damping is sufficient to converge to the analytical
1609 energy constants within moderately short cutoff distances. Because
1610 overdamping can present additional issues with higher order
1611 multipoles, our results indicate that the damping coefficient should
1612 be taken as small as possible, and that the {\it undamped} GSF and SP
1613 methods may be the best choice in crystalline systems.
1614
1615 The techniques used here to derive the force, torque and energy
1616 expressions can be extended to higher order multipoles, although some
1617 of the objects (e.g. the matrix cross product in
1618 Eq. \ref{eq:matrixCross}) will need to be generalized for higher-rank
1619 tensors. We also note that the definitions of the multipoles used
1620 here are in a primitive form, and these need some care when comparing
1621 with experiment or other computational techniques.
1622
1623 In large systems, these new methods can be made to scale approximately
1624 linearly with system size, and detailed comparisons with the Ewald sum
1625 for a wide range of chemical environments follows in the second paper.
1626
1627 \begin{acknowledgments}
1628 JDG acknowledges helpful discussions with Christopher
1629 Fennell. Support for this project was provided by the National
1630 Science Foundation under grant CHE-1362211. Computational time was
1631 provided by the Center for Research Computing (CRC) at the
1632 University of Notre Dame.
1633 \end{acknowledgments}
1634
1635 \newpage
1636 \appendix
1637
1638 \section{Smith's $B_l(r)$ functions for damped-charge distributions}
1639 \label{SmithFunc}
1640 The following summarizes Smith's $B_l(r)$ functions and includes
1641 formulas given in his appendix.\cite{Smith98} The first function
1642 $B_0(r)$ is defined by
1643 %
1644 \begin{equation}
1645 B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}
1646 \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
1647 \end{equation}
1648 %
1649 The first derivative of this function is
1650 %
1651 \begin{equation}
1652 \frac{dB_0(r)}{dr}=-\frac{1}{r^2}\text{erfc}(\alpha r)
1653 -\frac{2\alpha}{r\sqrt{\pi}}\text{e}^{-{\alpha}^2r^2}
1654 \end{equation}
1655 %
1656 which can be used to define a function $B_1(r)$:
1657 %
1658 \begin{equation}
1659 B_1(r)=-\frac{1}{r}\frac{dB_0(r)}{dr}
1660 \end{equation}
1661 %
1662 In general, the recurrence relation,
1663 \begin{equation}
1664 B_l(r)=-\frac{1}{r}\frac{dB_{l-1}(r)}{dr}
1665 = \frac{1}{r^2} \left[ (2l-1)B_{l-1}(r) + \frac {(2\alpha^2)^l}{\alpha \sqrt{\pi}}
1666 \text{e}^{-{\alpha}^2r^2}
1667 \right] ,
1668 \end{equation}
1669 is very useful for building up higher derivatives. As noted by Smith,
1670 it is possible to approximate the $B_l(r)$ functions,
1671 %
1672 \begin{equation}
1673 B_l(r)=\frac{(2l)!}{l!2^lr^{2l+1}} - \frac {(2\alpha^2)^{l+1}}{(2l+1)\alpha \sqrt{\pi}}
1674 +\text{O}(r) .
1675 \end{equation}
1676 \newpage
1677 \section{The $r$-dependent factors for TSF electrostatics}
1678 \label{radialTSF}
1679
1680 Using the shifted damped functions $f_n(r)$ defined by:
1681 %
1682 \begin{equation}
1683 f_n(r)= B_0(r) -\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} B_0^{(m)}(r_c) ,
1684 \end{equation}
1685 %
1686 where the superscript $(m)$ denotes the $m^\mathrm{th}$ derivative. In
1687 this Appendix, we provide formulas for successive derivatives of this
1688 function. (If there is no damping, then $B_0(r)$ is replaced by
1689 $1/r$.) First, we find:
1690 %
1691 \begin{equation}
1692 \frac{\partial f_n}{\partial r_\alpha}=\hat{r}_\alpha \frac{d f_n}{d r} .
1693 \end{equation}
1694 %
1695 This formula clearly brings in derivatives of Smith's $B_0(r)$
1696 function, and we define higher-order derivatives as follows:
1697 %
1698 \begin{align}
1699 g_n(r)=& \frac{d f_n}{d r} =
1700 B_0^{(1)}(r) -\sum_{m=0}^{n} \frac {(r-r_c)^m}{m!} B_0^{(m+1)}(r_c) \\
1701 h_n(r)=& \frac{d^2f_n}{d r^2} =
1702 B_0^{(2)}(r) -\sum_{m=0}^{n-1} \frac {(r-r_c)^m}{m!} B_0^{(m+2)}(r_c) \\
1703 s_n(r)=& \frac{d^3f_n}{d r^3} =
1704 B_0^{(3)}(r) -\sum_{m=0}^{n-2} \frac {(r-r_c)^m}{m!} B_0^{(m+3)}(r_c) \\
1705 t_n(r)=& \frac{d^4f_n}{d r^4} =
1706 B_0^{(4)}(r) -\sum_{m=0}^{n-3} \frac {(r-r_c)^m}{m!} B_0^{(m+4)}(r_c) \\
1707 u_n(r)=& \frac{d^5f_n}{d r^5} =
1708 B_0^{(5)}(r) -\sum_{m=0}^{n-4} \frac {(r-r_c)^m}{m!} B_0^{(m+5)}(r_c) .
1709 \end{align}
1710 %
1711 We note that the last function needed (for quadrupole-quadrupole interactions) is
1712 %
1713 \begin{equation}
1714 u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1715 \end{equation}
1716 % The functions
1717 % needed are listed schematically below:
1718 % %
1719 % \begin{eqnarray}
1720 % f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1721 % g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1722 % h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1723 % s_2 \quad s_3 \quad &s_4 \nonumber \\
1724 % t_3 \quad &t_4 \nonumber \\
1725 % &u_4 \nonumber .
1726 % \end{eqnarray}
1727 The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and
1728 stored on a grid for values of $r$ from $0$ to $r_c$. Using these
1729 functions, we find
1730 %
1731 \begin{align}
1732 \frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\
1733 \frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r}
1734 +r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\
1735 \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =&
1736 \left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta +
1737 \delta_{ \beta \gamma} r_\alpha \right)
1738 \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\
1739 & + r_\alpha r_\beta r_\gamma
1740 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\
1741 \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial
1742 r_\gamma \partial r_\delta} =&
1743 \left( \delta_{\alpha \beta} \delta_{\gamma \delta}
1744 + \delta_{\alpha \gamma} \delta_{\beta \delta}
1745 +\delta_{ \beta \gamma} \delta_{\alpha \delta} \right)
1746 \left( - \frac{g_n}{r^3} + \frac{h_n}{r^2} \right) \nonumber \\
1747 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta
1748 + \text{5 permutations}
1749 \right) \left( \frac{3 g_n}{r^5} - \frac{3h_n}{r^4} + \frac{s_n}{r^3}
1750 \right) \nonumber \\
1751 &+ r_\alpha r_\beta r_\gamma r_\delta
1752 \left( -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5}
1753 + \frac{t_n}{r^4} \right)\\
1754 \frac{\partial^5 f_n}
1755 {\partial r_\alpha \partial r_\beta \partial r_\gamma \partial
1756 r_\delta \partial r_\epsilon} =&
1757 \left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon
1758 + \text{14 permutations} \right)
1759 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\
1760 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta r_\epsilon
1761 + \text{9 permutations}
1762 \right) \left(- \frac{15g_n}{r^7}+\frac{15h_n}{r^7} -\frac{6s_n}{r^5} +\frac{t_n}{r^4}
1763 \right) \nonumber \\
1764 &+ r_\alpha r_\beta r_\gamma r_\delta r_\epsilon
1765 \left( \frac{105g_n}{r^9} - \frac{105h_n}{r^8} + \frac{45s_n}{r^7}
1766 - \frac{10t_n}{r^6} +\frac{u_n}{r^5} \right) \label{eq:b13}
1767 \end{align}
1768 %
1769 %
1770 %
1771 \newpage
1772 \section{The $r$-dependent factors for GSF electrostatics}
1773 \label{radialGSF}
1774
1775 In Gradient-shifted force electrostatics, the kernel is not expanded,
1776 and the expansion is carried out on the individual terms in the
1777 multipole interaction energies. For damped charges, this still brings
1778 multiple derivatives of the Smith's $B_0(r)$ function into the
1779 algebra. To denote these terms, we generalize the notation of the
1780 previous appendix. For either $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$
1781 (damped),
1782 %
1783 \begin{align}
1784 g(r) &= \frac{df}{d r} && &&=-\frac{1}{r^2}
1785 &&\mathrm{or~~~} -rB_1(r) \\
1786 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) \\
1787 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)\\
1788 t(r) &= \frac{ds}{d r} &&= \frac{d^4f}{d r^4} &&= \frac{24}{r^5} &&\mathrm{or~~~} 3
1789 B_2(r) - 6r^2 B_3(r) + r^4 B_4(r) \\
1790 u(r) &= \frac{dt}{d r} &&= \frac{d^5f}{d r^5} &&=-\frac{120}{r^6} &&\mathrm{or~~~} -15
1791 r B_3(r) + 10 r^3B_4(r) -r^5B_5(r).
1792 \end{align}
1793 %
1794 For undamped charges, Table I lists these derivatives under the Bare
1795 Coulomb column. Equations \ref{eq:b9} to \ref{eq:b13} are still
1796 correct for GSF electrostatics if the subscript $n$ is eliminated.
1797
1798 \newpage
1799
1800 \bibliography{multipole}
1801
1802 \end{document}
1803 %
1804 % ****** End of file multipole.tex ******