ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
Revision: 4195
Committed: Thu Jul 17 18:24:41 2014 UTC (9 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 72630 byte(s)
Log Message:
Most recent version including dielectric stuff

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 alternatives to the Ewald
47 Sum. I. Shifted electrostatics for multipoles}
48
49 \author{Madan Lamichhane}
50 \affiliation{Department of Physics, University
51 of Notre Dame, Notre Dame, IN 46556}
52
53 \author{J. Daniel Gezelter}
54 \email{gezelter@nd.edu.}
55 \affiliation{Department of Chemistry and Biochemistry, University
56 of Notre Dame, Notre Dame, IN 46556}
57
58 \author{Kathie E. Newman}
59 \affiliation{Department of Physics, University
60 of Notre Dame, Notre Dame, IN 46556}
61
62
63 \date{\today}% It is always \today, today,
64 % but any date may be explicitly specified
65
66 \begin{abstract}
67 We have extended the original damped-shifted force (DSF)
68 electrostatic kernel and have been able to derive three new
69 electrostatic potentials for higher-order multipoles that are based
70 on truncated Taylor expansions around the cutoff radius. These
71 include a shifted potential (SP) that generalizes the Wolf method
72 for point multipoles, and Taylor-shifted force (TSF) and
73 gradient-shifted force (GSF) potentials that are both
74 generalizations of DSF electrostatics for multipoles. We find that
75 each of the distinct orientational contributions requires a separate
76 radial function to ensure that pairwise energies, forces and torques
77 all vanish at the cutoff radius. In this paper, we present energy,
78 force, and torque expressions for the new models, and compare these
79 real-space interaction models to exact results for ordered arrays of
80 multipoles. We find that the GSF and SP methods converge rapidly to
81 the correct lattice energies for ordered dipolar and quadrupolar
82 arrays, while the Taylor-Shifted Force (TSF) is too severe an
83 approximation to provide accurate convergence to lattice energies.
84 Because real-space methods can be made to scale linearly with system
85 size, SP and GSF are attractive options for large Monte
86 Carlo and molecular dynamics simulations, respectively.
87 \end{abstract}
88
89 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
90 % Classification Scheme.
91 %\keywords{Suggested keywords}%Use showkeys class option if keyword
92 %display desired
93 \maketitle
94
95 \section{Introduction}
96 There has been increasing interest in real-space methods for
97 calculating electrostatic interactions in computer simulations of
98 condensed molecular
99 systems.\cite{Wolf99,Zahn02,Kast03,BeckD.A.C._bi0486381,Ma05,Fennell:2006zl,Chen:2004du,Chen:2006ii,Rodgers:2006nw,Denesyuk:2008ez,Izvekov:2008wo}
100 The simplest of these techniques was developed by Wolf {\it et al.}
101 in their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
102 For systems of point charges, Fennell and Gezelter showed that a
103 simple damped shifted force (DSF) modification to Wolf's method could
104 give nearly quantitative agreement with smooth particle mesh Ewald
105 (SPME)\cite{Essmann95} configurational energy differences as well as
106 atomic force and molecular torque vectors.\cite{Fennell:2006zl}
107
108 The computational efficiency and the accuracy of the DSF method are
109 surprisingly good, particularly for systems with uniform charge
110 density. Additionally, dielectric constants obtained using DSF and
111 similar methods where the force vanishes at $r_{c}$ are
112 essentially quantitative.\cite{Izvekov:2008wo} The DSF and other
113 related methods have now been widely investigated,\cite{Hansen:2012uq}
114 and DSF is now used routinely in a diverse set of chemical
115 environments.\cite{doi:10.1021/la400226g,McCann:2013fk,kannam:094701,Forrest:2012ly,English:2008kx,Louden:2013ve,Tokumasu:2013zr}
116 DSF electrostatics provides a compromise between the computational
117 speed of real-space cutoffs and the accuracy of fully-periodic Ewald
118 treatments.
119
120 One common feature of many coarse-graining approaches, which treat
121 entire molecular subsystems as a single rigid body, is simplification
122 of the electrostatic interactions between these bodies so that fewer
123 site-site interactions are required to compute configurational
124 energies. To do this, the interactions between coarse-grained sites
125 are typically taken to be point
126 multipoles.\cite{Golubkov06,ISI:000276097500009,ISI:000298664400012}
127
128 Water, in particular, has been modeled recently with point multipoles
129 up to octupolar
130 order.\cite{Chowdhuri:2006lr,Te:2010rt,Te:2010ys,Te:2010vn} For
131 maximum efficiency, these models require the use of an approximate
132 multipole expansion as the exact multipole expansion can become quite
133 expensive (particularly when handled via the Ewald
134 sum).\cite{Ichiye:2006qy} Point multipoles and multipole
135 polarizability have also been utilized in the AMOEBA water model and
136 related force fields.\cite{Ponder:2010fk,schnieders:124114,Ren:2011uq}
137
138 Higher-order multipoles present a peculiar issue for molecular
139 dynamics. Multipolar interactions are inherently short-ranged, and
140 should not need the relatively expensive Ewald treatment. However,
141 real-space cutoff methods are normally applied in an orientation-blind
142 fashion so multipoles which leave and then re-enter a cutoff sphere in
143 a different orientation can cause energy discontinuities.
144
145 This paper outlines an extension of the original DSF electrostatic
146 kernel to point multipoles. We describe three distinct real-space
147 interaction models for higher-order multipoles based on truncated
148 Taylor expansions that are carried out at the cutoff radius. We are
149 calling these models {\bf Taylor-shifted} (TSF), {\bf
150 gradient-shifted} (GSF) and {\bf shifted potential} (SP)
151 electrostatics. Because of differences in the initial assumptions,
152 the three methods yield related, but distinct expressions for energies,
153 forces, and torques.
154
155 In this paper we outline the new methodology and give functional forms
156 for the energies, forces, and torques up to quadrupole-quadrupole
157 order. We also compare the new methods to analytic energy constants
158 for periodic arrays of point multipoles. In the following paper, we
159 provide numerical comparisons to Ewald-based electrostatics in common
160 simulation enviornments.
161
162 \section{Methodology}
163 An efficient real-space electrostatic method involves the use of a
164 pair-wise functional form,
165 \begin{equation}
166 U = \sum_i \sum_{j>i} U_\mathrm{pair}(\mathbf{r}_{ij}, \Omega_i, \Omega_j) +
167 \sum_i U_i^\mathrm{self}
168 \end{equation}
169 that is short-ranged and easily truncated at a cutoff radius,
170 \begin{equation}
171 U_\mathrm{pair}(\mathbf{r}_{ij},\Omega_i, \Omega_j) = \left\{
172 \begin{array}{ll}
173 U_\mathrm{approx} (\mathbf{r}_{ij}, \Omega_i, \Omega_j) & \quad \left| \mathbf{r}_{ij} \right| \le r_c \\
174 0 & \quad \left| \mathbf{r}_{ij} \right| > r_c ,
175 \end{array}
176 \right.
177 \end{equation}
178 along with an easily computed self-interaction term ($\sum_i
179 U_i^\mathrm{self}$) which scales linearly with the number of
180 particles. Here $\Omega_i$ and $\Omega_j$ represent orientational
181 coordinates of the two sites, and $\mathbf{r}_{ij}$ is the vector
182 between the two sites. The computational efficiency, energy
183 conservation, and even some physical properties of a simulation can
184 depend dramatically on how the $U_\mathrm{approx}$ function behaves at
185 the cutoff radius. The goal of any approximation method should be to
186 mimic the real behavior of the electrostatic interactions as closely
187 as possible without sacrificing the near-linear scaling of a cutoff
188 method.
189
190 \subsection{Self-neutralization, damping, and force-shifting}
191 The DSF and Wolf methods operate by neutralizing the total charge
192 contained within the cutoff sphere surrounding each particle. This is
193 accomplished by shifting the potential functions to generate image
194 charges on the surface of the cutoff sphere for each pair interaction
195 computed within $r_c$. Damping using a complementary error function is
196 applied to the potential to accelerate convergence. The interaction
197 for a pair of charges ($C_i$ and $C_j$) in the DSF method,
198 \begin{equation*}
199 U_\mathrm{DSF}(r_{ij}) = C_i C_j \Biggr{[}
200 \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
201 - \frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c} + \left(\frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c^2}
202 + \frac{2\alpha}{\pi^{1/2}}
203 \frac{\exp\left(-\alpha^2r_c^2\right)}{r_c}
204 \right)\left(r_{ij}-r_c\right)\ \Biggr{]}
205 \label{eq:DSFPot}
206 \end{equation*}
207 where $\alpha$ is the adjustable damping parameter. Note that in this
208 potential and in all electrostatic quantities that follow, the
209 standard $1/4 \pi \epsilon_{0}$ has been omitted for clarity.
210
211 To insure net charge neutrality within each cutoff sphere, an
212 additional ``self'' term is added to the potential. This term is
213 constant (as long as the charges and cutoff radius do not change), and
214 exists outside the normal pair-loop for molecular simulations. It can
215 be thought of as a contribution from a charge opposite in sign, but
216 equal in magnitude, to the central charge, which has been spread out
217 over the surface of the cutoff sphere. A portion of the self term is
218 identical to the self term in the Ewald summation, and comes from the
219 utilization of the complimentary error function for electrostatic
220 damping.\cite{deLeeuw80,Wolf99} There have also been recent efforts to
221 extend the Wolf self-neutralization method to zero out the dipole and
222 higher order multipoles contained within the cutoff
223 sphere.\cite{Fukuda:2011jk,Fukuda:2012yu,Fukuda:2013qv}
224
225 In this work, we extend the idea of self-neutralization for the point
226 multipoles by insuring net charge-neutrality and net-zero moments
227 within each cutoff sphere. In Figure \ref{fig:shiftedMultipoles}, the
228 central dipolar site $\mathbf{D}_i$ is interacting with point dipole
229 $\mathbf{D}_j$ and point quadrupole, $\mathbf{Q}_k$. The
230 self-neutralization scheme for point multipoles involves projecting
231 opposing multipoles for sites $j$ and $k$ on the surface of the cutoff
232 sphere. There are also significant modifications made to make the
233 forces and torques go smoothly to zero at the cutoff distance.
234
235 \begin{figure}
236 \includegraphics[width=3in]{SM.eps}
237 \caption{Reversed multipoles are projected onto the surface of the
238 cutoff sphere. The forces, torques, and potential are then smoothly
239 shifted to zero as the sites leave the cutoff region.}
240 \label{fig:shiftedMultipoles}
241 \end{figure}
242
243 As in the point-charge approach, there is an additional contribution
244 from self-neutralization of site $i$. The self term for multipoles is
245 described in section \ref{sec:selfTerm}.
246
247 \subsection{The multipole expansion}
248
249 Consider two discrete rigid collections of point charges, denoted as
250 $\bf a$ and $\bf b$. In the following, we assume that the two objects
251 interact via electrostatics only and describe those interactions in
252 terms of a standard multipole expansion. Putting the origin of the
253 coordinate system at the center of mass of $\bf a$, we use vectors
254 $\mathbf{r}_k$ to denote the positions of all charges $q_k$ in $\bf
255 a$. Then the electrostatic potential of object $\bf a$ at
256 $\mathbf{r}$ is given by
257 \begin{equation}
258 \phi_a(\mathbf r) =
259 \sum_{k \, \text{in \bf a}} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}.
260 \end{equation}
261 The Taylor expansion in $r$ can be written using an implied summation
262 notation. Here Greek indices are used to indicate space coordinates
263 ($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for
264 labeling specific charges in $\bf a$ and $\bf b$ respectively. The
265 Taylor expansion,
266 \begin{equation}
267 \frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} =
268 \left( 1
269 - r_{k\alpha} \frac{\partial}{\partial r_{\alpha}}
270 + \frac{1}{2} r_{k\alpha} r_{k\beta} \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} +\dots
271 \right)
272 \frac{1}{r} ,
273 \end{equation}
274 can then be used to express the electrostatic potential on $\bf a$ in
275 terms of multipole operators,
276 \begin{equation}
277 \phi_{\bf a}(\mathbf{r}) =\hat{M}_{\bf a} \frac{1}{r}
278 \end{equation}
279 where
280 \begin{equation}
281 \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
282 + Q_{{\bf a}\alpha\beta}
283 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
284 \end{equation}
285 Here, the point charge, dipole, and quadrupole for object $\bf a$ are
286 given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
287 a}\alpha\beta}$, respectively. These are the primitive multipoles
288 which can be expressed as a distribution of charges,
289 \begin{align}
290 C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
291 D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
292 Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf 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 $\overleftrightarrow{\mathsf I}$, 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 $\bf b$. Then with $\bf{r}$ pointing from
306 $\bf a$ to $\bf b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $), the interaction energy is given by
307 \begin{equation}
308 U_{\bf{ab}}(r)
309 = \hat{M}_a \sum_{j \, \text{in \bf b}} \frac {q_j}{\vert \bf{r}+\bf{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 on object {\bf b},
313 \begin{equation}
314 \hat{M}_{\bf b} = C_{\bf b} + D_{{\bf b}\alpha} \frac{\partial}{\partial r_{\alpha}}
315 + Q_{{\bf 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_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf 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 $\bf a$ interacting with the same types of multipoles in $\bf b$.
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 $C_{\bf a}$ and $C_{\bf b}$ separated by a distance $r$ is
361 written:
362 \begin{equation}
363 U_{C_{\bf a}C_{\bf b}}(r)= C_{\bf a} C_{\bf 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 $\bf a$ interacting with a
413 dipole $\bf b$, we write
414 %
415 \begin{equation}
416 U_{C_{\bf a}D_{\bf b}}(r)=
417 C_{\bf a} D_{{\bf b}\alpha} \frac {\partial f_1(r) }{\partial r_\alpha}
418 = C_{\bf a} D_{{\bf b}\alpha}
419 \frac {r_\alpha}{r} \frac {\partial f_1(r)}{\partial r} .
420 \end{equation}
421 %
422 The force that dipole $\bf b$ exerts on charge $\bf a$ is
423 %
424 \begin{equation}
425 F_{C_{\bf a}D_{\bf b}\beta} = C_{\bf a} D_{{\bf 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_{\bf a}D_{\bf b}\beta} =
436 \frac{C_{\bf a} D_{{\bf 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_{\bf a} D_{{\bf 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 $Q_{{\bf a}\alpha\beta}Q_{{\bf b}\gamma\delta}$, 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}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
482 U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c)
483 \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{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 ($\hat{\mathbf{b}}$) 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}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
518 U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{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 $\bf a$ and $\bf b$ 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 $\bf a$ and $\bf b$ using unit vectors
542 $\hat{a}_m$ and $\hat{b}_m$, 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 $\hat{\mathbf {a}}$ and $\hat{\mathbf {b}}$ can be
550 expressed using these unit vectors:
551 \begin{eqnarray}
552 \hat{\mathbf {a}} =
553 \begin{pmatrix}
554 \hat{a}_1 \\
555 \hat{a}_2 \\
556 \hat{a}_3
557 \end{pmatrix}, \qquad
558 \hat{\mathbf {b}} =
559 \begin{pmatrix}
560 \hat{b}_1 \\
561 \hat{b}_2 \\
562 \hat{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 $\hat{r}$ and the body
572 axes ($\hat{a}_m$, $\hat{b}_n$) :
573 %
574 \begin{equation}
575 U(r, \{\hat{a}_m \cdot \hat{r} \},
576 \{\hat{b}_n\cdot \hat{r} \},
577 \{\hat{a}_m \cdot \hat{b}_n \}) .
578 \label{ugeneral}
579 \end{equation}
580 %
581 the forces come out relatively cleanly,
582 %
583 \begin{equation}
584 \mathbf{F}_{\bf a}=-\mathbf{F}_{\bf b} = \frac{\partial U}{\partial \mathbf{r}}
585 = \frac{\partial U}{\partial r} \hat{r}
586 + \sum_m \left[
587 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
588 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
589 + \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
590 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
591 \right] \label{forceequation}.
592 \end{equation}
593
594 The torques can also be found in a relatively similar
595 manner,
596 %
597 \begin{eqnarray}
598 \mathbf{\tau}_{\bf a} =
599 \sum_m
600 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
601 ( \hat{r} \times \hat{a}_m )
602 -\sum_{mn}
603 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
604 (\hat{a}_m \times \hat{b}_n) \\
605 %
606 \mathbf{\tau}_{\bf b} =
607 \sum_m
608 \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
609 ( \hat{r} \times \hat{b}_m)
610 +\sum_{mn}
611 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
612 (\hat{a}_m \times \hat{b}_n) .
613 \end{eqnarray}
614
615 Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $
616 is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk}
617 We also made use of the identities,
618 %
619 \begin{align}
620 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
621 =& \frac{1}{r} \left( \hat{a}_m - (\hat{a}_m \cdot \hat{r})\hat{r}
622 \right) \\
623 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
624 =& \frac{1}{r} \left( \hat{b}_m - (\hat{b}_m \cdot \hat{r})\hat{r}
625 \right).
626 \end{align}
627
628 Many of the multipole contractions required can be written in one of
629 three equivalent forms using the unit vectors $\hat{r}$, $\hat{a}_m$,
630 and $\hat{b}_n$. In the torque expressions, it is useful to have the
631 angular-dependent terms available in all three fashions, e.g. for the
632 dipole-dipole contraction:
633 %
634 \begin{equation}
635 \mathbf{D}_{\mathbf {a}} \cdot \mathbf{D}_{\mathbf{b}}
636 = D_{\bf {a}\alpha} D_{\bf {b}\alpha} =
637 \sum_{mn} {D_{\mathbf{a}m} \hat{a}_m \cdot \hat{b}_n D_{\mathbf{b}n}}.
638 \end{equation}
639 %
640 The first two forms are written using space coordinates. The first
641 form is standard in the chemistry literature, while the second is
642 expressed using implied summation notation. The third form shows
643 explicit sums over body indices and the dot products now indicate
644 contractions using space indices.
645
646 In computing our force and torque expressions, we carried out most of
647 the work in body coordinates, and have transformed the expressions
648 back to space-frame coordinates, which are reported below. Interested
649 readers may consult the supplemental information for this paper for
650 the intermediate body-frame expressions.
651
652 \subsection{The Self-Interaction \label{sec:selfTerm}}
653
654 In addition to cutoff-sphere neutralization, the Wolf
655 summation~\cite{Wolf99} and the damped shifted force (DSF)
656 extension~\cite{Fennell:2006zl} also include self-interactions that
657 are handled separately from the pairwise interactions between
658 sites. The self-term is normally calculated via a single loop over all
659 sites in the system, and is relatively cheap to evaluate. The
660 self-interaction has contributions from two sources.
661
662 First, the neutralization procedure within the cutoff radius requires
663 a contribution from a charge opposite in sign, but equal in magnitude,
664 to the central charge, which has been spread out over the surface of
665 the cutoff sphere. For a system of undamped charges, the total
666 self-term is
667 \begin{equation}
668 U_\textrm{self} = - \frac{1}{r_c} \sum_{{\bf a}=1}^N C_{\bf a}^{2}.
669 \label{eq:selfTerm}
670 \end{equation}
671
672 Second, charge damping with the complementary error function is a
673 partial analogy to the Ewald procedure which splits the interaction
674 into real and reciprocal space sums. The real space sum is retained
675 in the Wolf and DSF methods. The reciprocal space sum is first
676 minimized by folding the largest contribution (the self-interaction)
677 into the self-interaction from charge neutralization of the damped
678 potential. The remainder of the reciprocal space portion is then
679 discarded (as this contributes the largest computational cost and
680 complexity to the Ewald sum). For a system containing only damped
681 charges, the complete self-interaction can be written as
682 \begin{equation}
683 U_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} +
684 \frac{\alpha}{\sqrt{\pi}} \right) \sum_{{\bf a}=1}^N
685 C_{\bf 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 ({\bf a}) that has a charge $(C_{\bf a})$, dipole
709 $(\mathbf{D}_{\bf a})$, and quadrupole $(\mathbf{Q}_{\bf a})$}
710 \begin{ruledtabular}
711 \begin{tabular}{lccc}
712 Multipole order & Summed Quantity & Self-neutralization & Reciprocal \\ \hline
713 Charge & $C_{\bf a}^2$ & $-f(r_c)$ & $-\frac{\alpha}{\sqrt{\pi}}$ \\
714 Dipole & $|\mathbf{D}_{\bf a}|^2$ & $\frac{1}{3} \left( h(r_c) +
715 \frac{2 g(r_c)}{r_c} \right)$ & $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$\\
716 Quadrupole & $2 \mathbf{Q}_{\bf a}:\mathbf{Q}_{\bf a} + \text{Tr}(\mathbf{Q}_{\bf a})^2$ &
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 & $-2 C_{\bf a} \text{Tr}(\mathbf{Q}_{\bf a})$ & $\frac{1}{3} \left(
720 h(r_c) + \frac{2 g(r_c)}{r_c} \right)$& $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$ \\
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_{\bf a}C_{\bf b}}(r)=&
760 C_{\bf a} C_{\bf b} v_{01}(r) \label{uchch}
761 \\
762 %
763 % u ca db
764 %
765 U_{C_{\bf a}D_{\bf b}}(r)=&
766 C_{\bf a} \left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right) v_{11}(r)
767 \label{uchdip}
768 \\
769 %
770 % u ca qb
771 %
772 U_{C_{\bf a}Q_{\bf b}}(r)=& C_{\bf a } \Bigl[ \text{Tr}Q_{\bf b}
773 v_{21}(r) + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot
774 \hat{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_{D_{\bf a}D_{\bf b}}(r)=&
788 -\Bigr[ \left( \mathbf{D}_{\mathbf {a}} \cdot
789 \mathbf{D}_{\mathbf{b}} \right) v_{21}(r)
790 +\left( \mathbf{D}_{\mathbf {a}} \cdot \hat{r} \right)
791 \left( \mathbf{D}_{\mathbf {b}} \cdot \hat{r} \right)
792 v_{22}(r) \Bigr]
793 \label{udipdip}
794 \\
795 %
796 % u da qb
797 %
798 \begin{split}
799 % 1
800 U_{D_{\bf a}Q_{\bf b}}(r) =&
801 -\Bigl[
802 \text{Tr}\mathbf{Q}_{\mathbf{b}}
803 \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right)
804 +2 ( \mathbf{D}_{\mathbf{a}} \cdot
805 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r} ) \Bigr] v_{31}(r) \\
806 % 2
807 &- \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right)
808 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{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_{Q_{\bf a}Q_{\bf b}}(r)=&
844 \Bigl[
845 \text{Tr} \mathbf{Q}_{\mathbf{a}} \text{Tr} \mathbf{Q}_{\mathbf{b}}
846 +2
847 \mathbf{Q}_{\mathbf{a}} : \mathbf{Q}_{\mathbf{b}} \Bigr] v_{41}(r)
848 \\
849 % 2
850 &+\Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}}
851 \left( \hat{r} \cdot
852 \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} \right)
853 +\text{Tr}\mathbf{Q}_{\mathbf{b}}
854 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}}
855 \cdot \hat{r} \right) +4 (\hat{r} \cdot
856 \mathbf{Q}_{{\mathbf a}}\cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
857 \Bigr] v_{42}(r)
858 \\
859 % 4
860 &+
861 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right)
862 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} \right) v_{43}(r).
863 \label{uquadquad}
864 \end{split}
865 \end{align}
866 %
867 Note that the energies of multipoles on site $\mathbf{b}$ interacting
868 with those on site $\mathbf{a}$ can be obtained by swapping indices
869 along with the sign of the intersite vector, $\hat{r}$.
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 $\bf{a}$, $\mathbf{F}_{\bf a}$, due to object
1098 $\bf{b}$ is the negative of the force on $\bf{b}$ due to $\bf{a}$. For
1099 a simple charge-charge interaction, these forces will point along the
1100 $\pm \hat{r}$ directions, where $\mathbf{r}=\mathbf{r}_b -
1101 \mathbf{r}_a $. Thus
1102 %
1103 \begin{equation}
1104 F_{\bf a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_{\bf a}C_{\bf b}}}{\partial r}
1105 \quad \text{and} \quad F_{\bf b \alpha}
1106 = - \hat{r}_\alpha \frac{\partial U_{C_{\bf a}C_{\bf 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}_{{\bf a}C_{\bf a}C_{\bf b}} =&
1120 C_{\bf a} C_{\bf b} w_a(r) \hat{r} \\
1121 %
1122 %
1123 %
1124 \mathbf{F}_{{\bf a}C_{\bf a}D_{\bf b}} =&
1125 C_{\bf a} \Bigl[
1126 \left( \hat{r} \cdot \mathbf{D}_{\mathbf{b}} \right)
1127 w_b(r) \hat{r}
1128 + \mathbf{D}_{\mathbf{b}} w_c(r) \Bigr] \\
1129 %
1130 %
1131 %
1132 \mathbf{F}_{{\bf a}C_{\bf a}Q_{\bf b}} =&
1133 C_{\bf a } \Bigr[
1134 \text{Tr}\mathbf{Q}_{\bf b} w_d(r) \hat{r}
1135 + 2 \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} w_e(r)
1136 + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}
1137 \right) w_f(r) \hat{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}_{{\bf a}D_{\bf a}D_{\bf b}} =&
1152 - \mathbf{D}_{\mathbf {a}} \cdot \mathbf{D}_{\mathbf{b}} w_d(r) \hat{r}
1153 + \left( \mathbf{D}_{\mathbf {a}}
1154 \left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
1155 + \mathbf{D}_{\mathbf {b}} \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right) \right) w_e(r)\\
1156 % 2
1157 & - \left( \hat{r} \cdot \mathbf{D}_{\mathbf {a}} \right)
1158 \left( \hat{r} \cdot \mathbf{D}_{\mathbf {b}} \right) w_f(r) \hat{r}
1159 \end{split}\\
1160 %
1161 %
1162 %
1163 \begin{split}
1164 \mathbf{F}_{{\bf a}D_{\bf a}Q_{\bf b}} =& - \Bigl[
1165 \text{Tr}\mathbf{Q}_{\mathbf{b}} \mathbf{ D}_{\mathbf{a}}
1166 +2 \mathbf{D}_{\mathbf{a}} \cdot
1167 \mathbf{Q}_{\mathbf{b}} \Bigr] w_g(r)
1168 - \Bigl[
1169 \text{Tr}\mathbf{Q}_{\mathbf{b}}
1170 \left( \hat{r} \cdot \mathbf{D}_{\mathbf{a}} \right)
1171 +2 ( \mathbf{D}_{\mathbf{a}} \cdot
1172 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) \Bigr] w_h(r) \hat{r} \\
1173 % 3
1174 & - \Bigl[\mathbf{ D}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1175 +2 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}} ) (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} ) \Bigr]
1176 w_i(r)
1177 % 4
1178 -
1179 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}} )
1180 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) w_j(r) \hat{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}_{{\bf a}Q_{\bf a}Q_{\bf b}} =&
1218 \Bigl[
1219 \text{Tr}\mathbf{Q}_{\mathbf{a}} \text{Tr}\mathbf{Q}_{\mathbf{b}}
1220 + 2 \mathbf{Q}_{\mathbf{a}} : \mathbf{Q}_{\mathbf{b}} \Bigr] w_k(r) \hat{r} \\
1221 % 2
1222 &+ \Bigl[
1223 2\text{Tr}\mathbf{Q}_{\mathbf{b}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} )
1224 + 2\text{Tr}\mathbf{Q}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} )
1225 % 3
1226 +4 (\mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1227 + 4(\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}}) \Bigr] w_n(r) \\
1228 % 4
1229 &+ \Bigl[
1230 \text{Tr}\mathbf{Q}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1231 + \text{Tr}\mathbf{Q}_{\mathbf{b}}
1232 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1233 % 5
1234 +4 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot
1235 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) \Bigr] w_l(r) \hat{r} \\
1236 %
1237 &+ \Bigl[
1238 + 2 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} )
1239 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1240 %6
1241 +2 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1242 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} ) \Bigr] w_o(r) \\
1243 % 7
1244 &+
1245 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1246 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) w_m(r) \hat{r} \end{split}
1247 \end{align}
1248 Note that the forces for higher multipoles on site $\mathbf{a}$
1249 interacting with those of lower order on site $\mathbf{b}$ 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}_{{\bf b}C_{\bf a}D_{\bf b}} =&
1264 C_{\bf a} (\hat{r} \times \mathbf{D}_{\mathbf{b}}) v_{11}(r) \\
1265 %
1266 %
1267 %
1268 \mathbf{\tau}_{{\bf b}C_{\bf a}Q_{\bf b}} =&
1269 2C_{\bf a}
1270 \hat{r} \times ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{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}_{{\bf a}D_{\bf a}D_{\bf b}} =&
1283 \mathbf{D}_{\mathbf {a}} \times \mathbf{D}_{\mathbf{b}} v_{21}(r)
1284 % 2
1285 -
1286 (\hat{r} \times \mathbf{D}_{\mathbf {a}} )
1287 (\hat{r} \cdot \mathbf{D}_{\mathbf {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}_{{\bf a}D_{\bf a}Q_{\bf b}} =&
1303 \Bigl[
1304 -\text{Tr}\mathbf{Q}_{\mathbf{b}}
1305 (\hat{r} \times \mathbf{D}_{\mathbf{a}} )
1306 +2 \mathbf{D}_{\mathbf{a}} \times
1307 (\mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1308 \Bigr] v_{31}(r)
1309 % 3
1310 - (\hat{r} \times \mathbf{D}_{\mathbf{a}} )
1311 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{32}(r)\\
1312 %
1313 %
1314 %
1315 \mathbf{\tau}_{{\bf b}D_{\bf a}Q_{\bf b}} =&
1316 \Bigl[
1317 +2 ( \mathbf{D}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} ) \times
1318 \hat{r}
1319 -2 \mathbf{D}_{\mathbf{a}} \times
1320 (\mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1321 \Bigr] v_{31}(r)
1322 % 2
1323 +
1324 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}})
1325 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}}) \times \hat{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}_{{\bf a}Q_{\bf a}Q_{\bf b}} =&
1362 -4
1363 \mathbf{Q}_{{\mathbf a}} \times \mathbf{Q}_{{\mathbf b}}
1364 v_{41}(r) \\
1365 % 2
1366 &+
1367 \Bigl[-2\text{Tr}\mathbf{Q}_{\mathbf{b}}
1368 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \times \hat{r}
1369 +4 \hat{r} \times
1370 ( \mathbf{Q}_{{\mathbf a}} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1371 % 3
1372 -4 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} )\times
1373 ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} ) \Bigr] v_{42}(r) \\
1374 % 4
1375 &+ 2
1376 \hat{r} \times ( \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1377 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{43}(r) \end{split}\\
1378 %
1379 %
1380 %
1381 \begin{split}
1382 \mathbf{\tau}_{{\bf b}Q_{\bf a}Q_{\bf b}} =
1383 &4
1384 \mathbf{Q}_{{\mathbf a}} \times \mathbf{Q}_{{\mathbf b}} v_{41}(r) \\
1385 % 2
1386 &+ \Bigl[- 2\text{Tr}\mathbf{Q}_{\mathbf{a}}
1387 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} ) \times \hat{r}
1388 -4 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot
1389 \mathbf{Q}_{{\mathbf b}} ) \times
1390 \hat{r}
1391 +4 ( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \times
1392 ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1393 \Bigr] v_{42}(r) \\
1394 % 4
1395 &+2
1396 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1397 \hat{r} \times ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{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[\mathbf{A} \times \mathbf{B}\right]_\alpha = \sum_\beta
1404 \left[\mathbf{A}_{\alpha+1,\beta} \mathbf{B}_{\alpha+2,\beta}
1405 -\mathbf{A}_{\alpha+2,\beta} \mathbf{B}_{\alpha+2,\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 $\mathbf{a}$ interacting with those of lower order on site
1417 $\mathbf{b}$ 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_rCutNew.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 error function kernel,
1469 $B_0(r)$ was used in the right panel.}
1470 \label{fig:Dipoles_rCut}
1471 \end{figure}
1472
1473 \begin{figure}
1474 \includegraphics[width=\linewidth]{Dipoles_alphaNew.eps}
1475 \caption{Convergence to the lattice energy constants as a function of
1476 the reduced damping parameter ($\alpha^* = \alpha a$) for the
1477 different real-space methods in the same three dipolar crystals in
1478 Figure \ref{fig:Dipoles_rCut}. The left panel shows results for a
1479 relatively small cutoff radius ($r_c = 4.5 a$) while a larger cutoff
1480 radius ($r_c = 6 a$) was used in the right panel. }
1481 \label{fig:Dipoles_alpha}
1482 \end{figure}
1483
1484 For the purposes of testing the energy expressions and the
1485 self-neutralization schemes, the primary quantity of interest is the
1486 analytic energy constant for the perfect arrays. Convergence to these
1487 constants are shown as a function of both the cutoff radius, $r_c$,
1488 and the damping parameter, $\alpha$ in Figs.\ref{fig:Dipoles_rCut}
1489 and \ref{fig:Dipoles_alpha}. We have simultaneously tested a hard
1490 cutoff (where the kernel is simply truncated at the cutoff radius) in
1491 addition to the three new methods.
1492
1493 The hard cutoff exhibits oscillations around the analytic energy
1494 constants, and converges to incorrect energies when the complementary
1495 error function damping kernel is used. The shifted potential (SP)
1496 converges to the correct energy smoothly by $r_c = 4.5 a$ even for the
1497 undamped case. This indicates that the shifting and the correction
1498 provided by the self term are required for obtaining accurate energies.
1499 The Taylor-shifted force (TSF) approximation appears to perturb the
1500 potential too much inside the cutoff region to provide accurate
1501 measures of the energy constants. GSF is a compromise, converging to
1502 the correct energies within $r_c = 6 a$.
1503
1504 {\it Quadrupolar} analogues to the Madelung constants were first
1505 worked out by Nagai and Nakamura who computed the energies of selected
1506 quadrupole arrays based on extensions to the Luttinger and Tisza
1507 approach.\cite{Nagai01081960,Nagai01091963}
1508
1509 In analogy to the dipolar arrays, the total electrostatic energy for
1510 the quadrupolar arrays is:
1511 \begin{equation}
1512 E = C N \frac{3\bar{Q}^2}{4a^5}
1513 \end{equation}
1514 where $a$ is the lattice parameter, and $\bar{Q}$ is the effective
1515 quadrupole moment,
1516 \begin{equation}
1517 \bar{Q}^2 = 2 \left(3 Q : Q - (\text{Tr} Q)^2 \right)
1518 \end{equation}
1519 for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}.
1520 (For the traceless quadrupole tensor, $\Theta = 3 Q - \text{Tr} Q$,
1521 the effective moment, $\bar{Q}^2 = \frac{2}{3} \Theta : \Theta$.)
1522
1523 To test the new electrostatic methods for quadrupoles, we have
1524 constructed very large, $N=$ 8,000~(sc), 16,000~(bcc), and
1525 32,000~(fcc) arrays of linear quadrupoles in the orientations
1526 described in Ref. \onlinecite{Nagai01081960}. We have compared the
1527 energy constants for these low-energy configurations for linear
1528 quadrupoles. Convergence to these constants are shown as a function of
1529 both the cutoff radius, $r_c$, and the damping parameter, $\alpha$ in
1530 Figs.~\ref{fig:Quadrupoles_rCut} and \ref{fig:Quadrupoles_alpha}.
1531
1532 \begin{figure}
1533 \includegraphics[width=\linewidth]{Quadrupoles_rcutConvergence.eps}
1534 \caption{Convergence of the lattice energy constants as a function of
1535 cutoff radius (normalized by the lattice constant, $a$) for the new
1536 real-space methods. Three quadrupolar crystal structures were
1537 sampled, and the analytic energy constants for the three lattices
1538 are indicated with grey dashed lines. The left panel shows results
1539 for the undamped kernel ($1/r$), while the damped error function
1540 kernel, $B_0(r)$ was used in the right panel.}
1541 \label{fig:Quadrupoles_rCut}
1542 \end{figure}
1543
1544
1545 \begin{figure}
1546 \includegraphics[width=\linewidth]{Quadrupoles_newAlpha.eps}
1547 \caption{Convergence to the lattice energy constants as a function of
1548 the reduced damping parameter ($\alpha^* = \alpha a$) for the
1549 different real-space methods in the same three quadrupolar crystals
1550 in Figure \ref{fig:Quadrupoles_rCut}. The left panel shows
1551 results for a relatively small cutoff radius ($r_c = 4.5 a$) while a
1552 larger cutoff radius ($r_c = 6 a$) was used in the right panel. }
1553 \label{fig:Quadrupoles_alpha}
1554 \end{figure}
1555
1556 Again, we find that the hard cutoff exhibits oscillations around the
1557 analytic energy constants. The shifted potential (SP) approximation
1558 converges to the correct energy smoothly by $r_c = 3 a$ even for the
1559 undamped case. The Taylor-shifted force (TSF) approximation again
1560 appears to perturb the potential too much inside the cutoff region to
1561 provide accurate measures of the energy constants. GSF again provides
1562 a compromise between the two methods -- energies are converged by $r_c
1563 = 4.5 a$, and the approximation is not as perturbative at short range
1564 as TSF.
1565
1566 It is also useful to understand the convergence to the lattice energy
1567 constants as a function of the reduced damping parameter ($\alpha^* =
1568 \alpha a$) for the different real-space methods.
1569 Figures \ref{fig:Dipoles_alpha} and \ref{fig:Quadrupoles_alpha} show
1570 this comparison for the dipolar and quadrupolar lattices,
1571 respectively. All of the methods (except for TSF) have excellent
1572 behavior for the undamped or weakly-damped cases. All of the methods
1573 can be forced to converge by increasing the value of $\alpha$, which
1574 effectively shortens the overall range of the potential by equalizing
1575 the truncation effects on the different orientational contributions.
1576 In the second paper in the series, we discuss how large values of
1577 $\alpha$ can perturb the force and torque vectors, but both
1578 weakly-damped or over-damped electrostatics appear to generate
1579 reasonable values for the total electrostatic energies under both the
1580 SP and GSF approximations.
1581
1582 \section{Conclusion}
1583 We have presented three efficient real-space methods for computing the
1584 interactions between point multipoles. One of these (SP) is a
1585 multipolar generalization of Wolf's method that smoothly shifts
1586 electrostatic energies to zero at the cutoff radius. Two of these
1587 methods (GSF and TSF) also smoothly truncate the forces and torques
1588 (in addition to the energies) at the cutoff radius, making them
1589 attractive for both molecular dynamics and Monte Carlo simulations. We
1590 find that the Gradient-Shifted Force (GSF) and the Shifted-Potential
1591 (SP) methods converge rapidly to the correct lattice energies for
1592 ordered dipolar and quadrupolar arrays, while the Taylor-Shifted Force
1593 (TSF) is too severe an approximation to provide accurate convergence
1594 to lattice energies.
1595
1596 In most cases, GSF can obtain nearly quantitative agreement with the
1597 lattice energy constants with reasonably small cutoff radii. The only
1598 exception we have observed is for crystals which exhibit a bulk
1599 macroscopic dipole moment (e.g. Luttinger \& Tisza's $Z_1$ lattice).
1600 In this particular case, the multipole neutralization scheme can
1601 interfere with the correct computation of the energies. We note that
1602 the energies for these arrangements are typically much larger than for
1603 crystals with net-zero moments, so this is not expected to be an issue
1604 in most simulations.
1605
1606 The techniques used here to derive the force, torque and energy
1607 expressions can be extended to higher order multipoles, although some
1608 of the objects (e.g. the matrix cross product in
1609 Eq. \ref{eq:matrixCross}) will need to be generalized for higher-rank
1610 tensors. We also note that the definitions of the multipoles used
1611 here are in a primitive form, and these need some care when comparing
1612 with experiment or other computational techniques.
1613
1614 In large systems, these new methods can be made to scale approximately
1615 linearly with system size, and detailed comparisons with the Ewald sum
1616 for a wide range of chemical environments follows in the second paper.
1617
1618 \begin{acknowledgments}
1619 JDG acknowledges helpful discussions with Christopher
1620 Fennell. Support for this project was provided by the National
1621 Science Foundation under grant CHE-1362211. Computational time was
1622 provided by the Center for Research Computing (CRC) at the
1623 University of Notre Dame.
1624 \end{acknowledgments}
1625
1626 \newpage
1627 \appendix
1628
1629 \section{Smith's $B_l(r)$ functions for damped-charge distributions}
1630 \label{SmithFunc}
1631 The following summarizes Smith's $B_l(r)$ functions and includes
1632 formulas given in his appendix.\cite{Smith98} The first function
1633 $B_0(r)$ is defined by
1634 %
1635 \begin{equation}
1636 B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}=
1637 \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
1638 \end{equation}
1639 %
1640 The first derivative of this function is
1641 %
1642 \begin{equation}
1643 \frac{dB_0(r)}{dr}=-\frac{1}{r^2}\text{erfc}(\alpha r)
1644 -\frac{2\alpha}{r\sqrt{\pi}}\text{e}^{-{\alpha}^2r^2}
1645 \end{equation}
1646 %
1647 which can be used to define a function $B_1(r)$:
1648 %
1649 \begin{equation}
1650 B_1(r)=-\frac{1}{r}\frac{dB_0(r)}{dr}
1651 \end{equation}
1652 %
1653 In general, the recurrence relation,
1654 \begin{equation}
1655 B_l(r)=-\frac{1}{r}\frac{dB_{l-1}(r)}{dr}
1656 = \frac{1}{r^2} \left[ (2l-1)B_{l-1}(r) + \frac {(2\alpha^2)^l}{\alpha \sqrt{\pi}}
1657 \text{e}^{-{\alpha}^2r^2}
1658 \right] ,
1659 \end{equation}
1660 is very useful for building up higher derivatives. As noted by Smith,
1661 it is possible to approximate the $B_l(r)$ functions,
1662 %
1663 \begin{equation}
1664 B_l(r)=\frac{(2l)!}{l!2^lr^{2l+1}} - \frac {(2\alpha^2)^{l+1}}{(2l+1)\alpha \sqrt{\pi}}
1665 +\text{O}(r) .
1666 \end{equation}
1667 \newpage
1668 \section{The $r$-dependent factors for TSF electrostatics}
1669 \label{radialTSF}
1670
1671 Using the shifted damped functions $f_n(r)$ defined by:
1672 %
1673 \begin{equation}
1674 f_n(r)= B_0(r) -\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} B_0^{(m)}(r_c) ,
1675 \end{equation}
1676 %
1677 where the superscript $(m)$ denotes the $m^\mathrm{th}$ derivative. In
1678 this Appendix, we provide formulas for successive derivatives of this
1679 function. (If there is no damping, then $B_0(r)$ is replaced by
1680 $1/r$.) First, we find:
1681 %
1682 \begin{equation}
1683 \frac{\partial f_n}{\partial r_\alpha}=\hat{r}_\alpha \frac{d f_n}{d r} .
1684 \end{equation}
1685 %
1686 This formula clearly brings in derivatives of Smith's $B_0(r)$
1687 function, and we define higher-order derivatives as follows:
1688 %
1689 \begin{align}
1690 g_n(r)=& \frac{d f_n}{d r} =
1691 B_0^{(1)}(r) -\sum_{m=0}^{n} \frac {(r-r_c)^m}{m!} B_0^{(m+1)}(r_c) \\
1692 h_n(r)=& \frac{d^2f_n}{d r^2} =
1693 B_0^{(2)}(r) -\sum_{m=0}^{n-1} \frac {(r-r_c)^m}{m!} B_0^{(m+2)}(r_c) \\
1694 s_n(r)=& \frac{d^3f_n}{d r^3} =
1695 B_0^{(3)}(r) -\sum_{m=0}^{n-2} \frac {(r-r_c)^m}{m!} B_0^{(m+3)}(r_c) \\
1696 t_n(r)=& \frac{d^4f_n}{d r^4} =
1697 B_0^{(4)}(r) -\sum_{m=0}^{n-3} \frac {(r-r_c)^m}{m!} B_0^{(m+4)}(r_c) \\
1698 u_n(r)=& \frac{d^5f_n}{d r^5} =
1699 B_0^{(5)}(r) -\sum_{m=0}^{n-4} \frac {(r-r_c)^m}{m!} B_0^{(m+5)}(r_c) .
1700 \end{align}
1701 %
1702 We note that the last function needed (for quadrupole-quadrupole interactions) is
1703 %
1704 \begin{equation}
1705 u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1706 \end{equation}
1707 % The functions
1708 % needed are listed schematically below:
1709 % %
1710 % \begin{eqnarray}
1711 % f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1712 % g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1713 % h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1714 % s_2 \quad s_3 \quad &s_4 \nonumber \\
1715 % t_3 \quad &t_4 \nonumber \\
1716 % &u_4 \nonumber .
1717 % \end{eqnarray}
1718 The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and
1719 stored on a grid for values of $r$ from $0$ to $r_c$. Using these
1720 functions, we find
1721 %
1722 \begin{align}
1723 \frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\
1724 \frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r}
1725 +r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\
1726 \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =&
1727 \left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta +
1728 \delta_{ \beta \gamma} r_\alpha \right)
1729 \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\
1730 & + r_\alpha r_\beta r_\gamma
1731 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\
1732 \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial
1733 r_\gamma \partial r_\delta} =&
1734 \left( \delta_{\alpha \beta} \delta_{\gamma \delta}
1735 + \delta_{\alpha \gamma} \delta_{\beta \delta}
1736 +\delta_{ \beta \gamma} \delta_{\alpha \delta} \right)
1737 \left( - \frac{g_n}{r^3} + \frac{h_n}{r^2} \right) \nonumber \\
1738 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta
1739 + \text{5 permutations}
1740 \right) \left( \frac{3 g_n}{r^5} - \frac{3h_n}{r^4} + \frac{s_n}{r^3}
1741 \right) \nonumber \\
1742 &+ r_\alpha r_\beta r_\gamma r_\delta
1743 \left( -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5}
1744 + \frac{t_n}{r^4} \right)\\
1745 \frac{\partial^5 f_n}
1746 {\partial r_\alpha \partial r_\beta \partial r_\gamma \partial
1747 r_\delta \partial r_\epsilon} =&
1748 \left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon
1749 + \text{14 permutations} \right)
1750 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\
1751 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta r_\epsilon
1752 + \text{9 permutations}
1753 \right) \left(- \frac{15g_n}{r^7}+\frac{15h_n}{r^7} -\frac{6s_n}{r^5} +\frac{t_n}{r^4}
1754 \right) \nonumber \\
1755 &+ r_\alpha r_\beta r_\gamma r_\delta r_\epsilon
1756 \left( \frac{105g_n}{r^9} - \frac{105h_n}{r^8} + \frac{45s_n}{r^7}
1757 - \frac{10t_n}{r^6} +\frac{u_n}{r^5} \right) \label{eq:b13}
1758 \end{align}
1759 %
1760 %
1761 %
1762 \newpage
1763 \section{The $r$-dependent factors for GSF electrostatics}
1764 \label{radialGSF}
1765
1766 In Gradient-shifted force electrostatics, the kernel is not expanded,
1767 and the expansion is carried out on the individual terms in the
1768 multipole interaction energies. For damped charges, this still brings
1769 multiple derivatives of the Smith's $B_0(r)$ function into the
1770 algebra. To denote these terms, we generalize the notation of the
1771 previous appendix. For either $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$
1772 (damped),
1773 %
1774 \begin{align}
1775 g(r) &= \frac{df}{d r} && &&=-\frac{1}{r^2}
1776 &&\mathrm{or~~~} -rB_1(r) \\
1777 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) \\
1778 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)\\
1779 t(r) &= \frac{ds}{d r} &&= \frac{d^4f}{d r^4} &&= \frac{24}{r^5} &&\mathrm{or~~~} 3
1780 B_2(r) - 6r^2 B_3(r) + r^4 B_4(r) \\
1781 u(r) &= \frac{dt}{d r} &&= \frac{d^5f}{d r^5} &&=-\frac{120}{r^6} &&\mathrm{or~~~} -15
1782 r B_3(r) + 10 r^3B_4(r) -r^5B_5(r).
1783 \end{align}
1784 %
1785 For undamped charges, Table I lists these derivatives under the Bare
1786 Coulomb column. Equations \ref{eq:b9} to \ref{eq:b13} are still
1787 correct for GSF electrostatics if the subscript $n$ is eliminated.
1788
1789 \newpage
1790
1791 \bibliography{multipole}
1792
1793 \end{document}
1794 %
1795 % ****** End of file multipole.tex ******