ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
Revision: 4181
Committed: Thu Jun 12 14:58:06 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 72907 byte(s)
Log Message:
Edits

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