ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
Revision: 3996
Committed: Wed Jan 15 15:26:27 2014 UTC (10 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 66412 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. Taylor-shifted and Gradient-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 two new
69 electrostatic potentials for higher-order multipoles that are based
70 on truncated Taylor expansions around the cutoff radius. For
71 multipole-multipole interactions, we find that each of the distinct
72 orientational contributions has a separate radial function to ensure
73 that the overall forces and torques vanish at the cutoff radius. In
74 this paper, we present energy, force, and torque expressions for the
75 new models, and compare these real-space interaction models to exact
76 results for ordered arrays of multipoles.
77 \end{abstract}
78
79 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
80 % Classification Scheme.
81 %\keywords{Suggested keywords}%Use showkeys class option if keyword
82 %display desired
83 \maketitle
84
85 \section{Introduction}
86 There has been increasing interest in real-space methods for
87 calculating electrostatic interactions in computer simulations of
88 condensed molecular
89 systems.\cite{Wolf99,Zahn02,Kast03,BeckD.A.C._bi0486381,Ma05,Fennell:2006zl,Chen:2004du,Chen:2006ii,Rodgers:2006nw,Denesyuk:2008ez,Izvekov:2008wo}
90 The simplest of these techniques was developed by Wolf {\it et al.}
91 in their work towards an $\mathcal{O}(N)$ Coulombic sum.\cite{Wolf99}
92 For systems of point charges, Fennell and Gezelter showed that a
93 simple damped shifted force (DSF) modification to Wolf's method could
94 give nearly quantitative agreement with smooth particle mesh Ewald
95 (SPME)\cite{Essmann95} configurational energy differences as well as
96 atomic force and molecular torque vectors.\cite{Fennell:2006zl}
97
98 The computational efficiency and the accuracy of the DSF method are
99 surprisingly good, particularly for systems with uniform charge
100 density. Additionally, dielectric constants obtained using DSF and
101 similar methods where the force vanishes at $r_{c}$ are
102 essentially quantitative.\cite{Izvekov:2008wo} The DSF and other
103 related methods have now been widely investigated,\cite{Hansen:2012uq}
104 and DSF is now used routinely in a diverse set of chemical
105 environments.\cite{doi:10.1021/la400226g,McCann:2013fk,kannam:094701,Forrest:2012ly,English:2008kx,Louden:2013ve,Tokumasu:2013zr}
106 DSF electrostatics provides a compromise between the computational
107 speed of real-space cutoffs and the accuracy of fully-periodic Ewald
108 treatments.
109
110 One common feature of many coarse-graining approaches, which treat
111 entire molecular subsystems as a single rigid body, is simplification
112 of the electrostatic interactions between these bodies so that fewer
113 site-site interactions are required to compute configurational
114 energies. To do this, the interactions between coarse-grained sites
115 are typically taken to be point
116 multipoles.\cite{Golubkov06,ISI:000276097500009,ISI:000298664400012}
117
118 Water, in particular, has been modeled recently with point multipoles
119 up to octupolar
120 order.\cite{Chowdhuri:2006lr,Te:2010rt,Te:2010ys,Te:2010vn} For
121 maximum efficiency, these models require the use of an approximate
122 multipole expansion as the exact multipole expansion can become quite
123 expensive (particularly when handled via the Ewald
124 sum).\cite{Ichiye:2006qy} Point multipoles and multipole
125 polarizability have also been utilized in the AMOEBA water model and
126 related force fields.\cite{Ponder:2010fk,schnieders:124114,Ren:2011uq}
127
128 Higher-order multipoles present a peculiar issue for molecular
129 dynamics. Multipolar interactions are inherently short-ranged, and
130 should not need the relatively expensive Ewald treatment. However,
131 real-space cutoff methods are normally applied in an orientation-blind
132 fashion so multipoles which leave and then re-enter a cutoff sphere in
133 a different orientation can cause energy discontinuities.
134
135 This paper outlines an extension of the original DSF electrostatic
136 kernel to point multipoles. We describe two distinct real-space
137 interaction models for higher-order multipoles based on two truncated
138 Taylor expansions that are carried out at the cutoff radius. We are
139 calling these models {\bf Taylor-shifted} and {\bf Gradient-shifted}
140 electrostatics. Because of differences in the initial assumptions,
141 the two methods yield related, but somewhat different expressions for
142 energies, forces, and torques.
143
144 In this paper we outline the new methodology and give functional forms
145 for the energies, forces, and torques up to quadrupole-quadrupole
146 order. We also compare the new methods to analytic energy constants
147 for periodic arrays of point multipoles. In the following paper, we
148 provide numerical comparisons to Ewald-based electrostatics in common
149 simulation enviornments.
150
151 \section{Methodology}
152 An efficient real-space electrostatic method involves the use of a
153 pair-wise functional form,
154 \begin{equation}
155 V = \sum_i \sum_{j>i} V_\mathrm{pair}(r_{ij}, \Omega_i, \Omega_j) +
156 \sum_i V_i^\mathrm{self}
157 \end{equation}
158 that is short-ranged and easily truncated at a cutoff radius,
159 \begin{equation}
160 V_\mathrm{pair}(r_{ij},\Omega_i, \Omega_j) = \left\{
161 \begin{array}{ll}
162 V_\mathrm{approx} (r_{ij}, \Omega_i, \Omega_j) & \quad r \le r_c \\
163 0 & \quad r > r_c ,
164 \end{array}
165 \right.
166 \end{equation}
167 along with an easily computed self-interaction term ($\sum_i
168 V_i^\mathrm{self}$) which has linear-scaling with the number of
169 particles. Here $\Omega_i$ and $\Omega_j$ represent orientational
170 coordinates of the two sites. The computational efficiency, energy
171 conservation, and even some physical properties of a simulation can
172 depend dramatically on how the $V_\mathrm{approx}$ function behaves at
173 the cutoff radius. The goal of any approximation method should be to
174 mimic the real behavior of the electrostatic interactions as closely
175 as possible without sacrificing the near-linear scaling of a cutoff
176 method.
177
178 \subsection{Self-neutralization, damping, and force-shifting}
179 The DSF and Wolf methods operate by neutralizing the total charge
180 contained within the cutoff sphere surrounding each particle. This is
181 accomplished by shifting the potential functions to generate image
182 charges on the surface of the cutoff sphere for each pair interaction
183 computed within $r_c$. Damping using a complementary error
184 function is applied to the potential to accelerate convergence. The
185 potential for the DSF method, where $\alpha$ is the adjustable damping
186 parameter, is given by
187 \begin{equation*}
188 V_\mathrm{DSF}(r) = C_i C_j \Biggr{[}
189 \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
190 - \frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c} + \left(\frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c^2}
191 + \frac{2\alpha}{\pi^{1/2}}
192 \frac{\exp\left(-\alpha^2r_c^2\right)}{r_c}
193 \right)\left(r_{ij}-r_c\right)\ \Biggr{]}
194 \label{eq:DSFPot}
195 \end{equation*}
196 Note that in this potential and in all electrostatic quantities that
197 follow, the standard $1/4 \pi \epsilon_{0}$ has been omitted for
198 clarity.
199
200 To insure net charge neutrality within each cutoff sphere, an
201 additional ``self'' term is added to the potential. This term is
202 constant (as long as the charges and cutoff radius do not change), and
203 exists outside the normal pair-loop for molecular simulations. It can
204 be thought of as a contribution from a charge opposite in sign, but
205 equal in magnitude, to the central charge, which has been spread out
206 over the surface of the cutoff sphere. A portion of the self term is
207 identical to the self term in the Ewald summation, and comes from the
208 utilization of the complimentary error function for electrostatic
209 damping.\cite{deLeeuw80,Wolf99} There have also been recent efforts to
210 extend the Wolf self-neutralization method to zero out the dipole and
211 higher order multipoles contained within the cutoff
212 sphere.\cite{Fukuda:2011jk,Fukuda:2012yu,Fukuda:2013qv}
213
214 In this work, we extend the idea of self-neutralization for the point
215 multipoles by insuring net charge-neutrality and net-zero moments
216 within each cutoff sphere. In Figure \ref{fig:shiftedMultipoles}, the
217 central dipolar site $\mathbf{D}_i$ is interacting with point dipole
218 $\mathbf{D}_j$ and point quadrupole, $\mathbf{Q}_k$. The
219 self-neutralization scheme for point multipoles involves projecting
220 opposing multipoles for sites $j$ and $k$ on the surface of the cutoff
221 sphere. There are also significant modifications made to make the
222 forces and torques go smoothly to zero at the cutoff distance.
223
224 \begin{figure}
225 \includegraphics[width=3in]{SM}
226 \caption{Reversed multipoles are projected onto the surface of the
227 cutoff sphere. The forces, torques, and potential are then smoothly
228 shifted to zero as the sites leave the cutoff region.}
229 \label{fig:shiftedMultipoles}
230 \end{figure}
231
232 As in the point-charge approach, there is an additional contribution
233 from self-neutralization of site $i$. The self term for multipoles is
234 described in section \ref{sec:selfTerm}.
235
236 \subsection{The multipole expansion}
237
238 Consider two discrete rigid collections of point charges, denoted as
239 $\bf a$ and $\bf b$. In the following, we assume that the two objects
240 interact via electrostatics only and describe those interactions in
241 terms of a standard multipole expansion. Putting the origin of the
242 coordinate system at the center of mass of $\bf a$, we use vectors
243 $\mathbf{r}_k$ to denote the positions of all charges $q_k$ in $\bf
244 a$. Then the electrostatic potential of object $\bf a$ at
245 $\mathbf{r}$ is given by
246 \begin{equation}
247 V_a(\mathbf r) =
248 \sum_{k \, \text{in \bf a}} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}.
249 \end{equation}
250 The Taylor expansion in $r$ can be written using an implied summation
251 notation. Here Greek indices are used to indicate space coordinates
252 ($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for
253 labelling specific charges in $\bf a$ and $\bf b$ respectively. The
254 Taylor expansion,
255 \begin{equation}
256 \frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} =
257 \left( 1
258 - r_{k\alpha} \frac{\partial}{\partial r_{\alpha}}
259 + \frac{1}{2} r_{k\alpha} r_{k\beta} \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} +\dots
260 \right)
261 \frac{1}{r} ,
262 \end{equation}
263 can then be used to express the electrostatic potential on $\bf a$ in
264 terms of multipole operators,
265 \begin{equation}
266 V_{\bf a}(\mathbf{r}) =\hat{M}_{\bf a} \frac{1}{r}
267 \end{equation}
268 where
269 \begin{equation}
270 \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
271 + Q_{{\bf a}\alpha\beta}
272 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
273 \end{equation}
274 Here, the point charge, dipole, and quadrupole for object $\bf a$ are
275 given by $C_{\bf a}$, $D_{{\bf a}\alpha}$, and $Q_{{\bf
276 a}\alpha\beta}$, respectively. These are the primitive multipoles
277 which can be expressed as a distribution of charges,
278 \begin{align}
279 C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
280 D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
281 Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k
282 r_{k\alpha} r_{k\beta} . \label{eq:quadrupole}
283 \end{align}
284 Note that the definition of the primitive quadrupole here differs from
285 the standard traceless form, and contains an additional Taylor-series
286 based factor of $1/2$. We are essentially treating the mass
287 distribution with higher priority; the moment of inertia tensor,
288 $\overleftrightarrow{\mathsf I}$, is diagonalized to obtain body-fixed
289 axes, and the charge distribution may result in a quadrupole tensor
290 that is not necessarily diagonal in the body frame. Additional
291 reasons for utilizing the primitive quadrupole are discussed in
292 section \ref{sec:damped}.
293
294 It is convenient to locate charges $q_j$ relative to the center of mass of $\bf b$. Then with $\bf{r}$ pointing from
295 $\bf a$ to $\bf b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_b $), the interaction energy is given by
296 \begin{equation}
297 U_{\bf{ab}}(r)
298 = \hat{M}_a \sum_{j \, \text{in \bf b}} \frac {q_j}{\vert \bf{r}+\bf{r}_j \vert} .
299 \end{equation}
300 This can also be expanded as a Taylor series in $r$. Using a notation
301 similar to before to define the multipoles on object {\bf b},
302 \begin{equation}
303 \hat{M}_{\bf b} = C_{\bf b} + D_{{\bf b}\alpha} \frac{\partial}{\partial r_{\alpha}}
304 + Q_{{\bf b}\alpha\beta}
305 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
306 \end{equation}
307 we arrive at the multipole expression for the total interaction energy.
308 \begin{equation}
309 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
310 \end{equation}
311 This form has the benefit of separating out the energies of
312 interaction into contributions from the charge, dipole, and quadrupole
313 of $\bf a$ interacting with the same multipoles on $\bf b$.
314
315 \subsection{Damped Coulomb interactions}
316 \label{sec:damped}
317 In the standard multipole expansion, one typically uses the bare
318 Coulomb potential, with radial dependence $1/r$, as shown in
319 Eq.~(\ref{kernel}). It is also quite common to use a damped Coulomb
320 interaction, which results from replacing point charges with Gaussian
321 distributions of charge with width $\alpha$. In damped multipole
322 electrostatics, the kernel ($1/r$) of the expansion is replaced with
323 the function:
324 \begin{equation}
325 B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}
326 \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
327 \end{equation}
328 We develop equations below using the function $f(r)$ to represent
329 either $1/r$ or $B_0(r)$, and all of the techniques can be applied to
330 bare or damped Coulomb kernels (or any other function) as long as
331 derivatives of these functions are known. Smith's convenient
332 functions $B_l(r)$ are summarized in Appendix A. (N.B. there is one
333 important distinction between the two kernels, which is the behavior
334 of $\nabla^2 \frac{1}{r}$ compared with $\nabla^2 B_0(r)$. The former
335 is zero everywhere except for a delta function evaluated at the
336 origin. The latter also has delta function behavior, but is non-zero
337 for $r \neq 0$. Thus the standard justification for using a traceless
338 quadrupole tensor fails for the damped case.)
339
340 The main goal of this work is to smoothly cut off the interaction
341 energy as well as forces and torques as $r\rightarrow r_c$. To
342 describe how this goal may be met, we use two examples, charge-charge
343 and charge-dipole, using the bare Coulomb kernel, $f(r)=1/r$, to
344 explain the idea.
345
346 \subsection{Shifted-force methods}
347 In the shifted-force approximation, the interaction energy for two
348 charges $C_{\bf a}$ and $C_{\bf b}$ separated by a distance $r$ is
349 written:
350 \begin{equation}
351 U_{C_{\bf a}C_{\bf b}}(r)= C_{\bf a} C_{\bf b}
352 \left({ \frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} }
353 \right) .
354 \end{equation}
355 Two shifting terms appear in this equations, one from the
356 neutralization procedure ($-1/r_c$), and one that causes the first
357 derivative to vanish at the cutoff radius.
358
359 Since one derivative of the interaction energy is needed for the
360 force, the minimal perturbation is a term linear in $(r-r_c)$ in the
361 interaction energy, that is:
362 \begin{equation}
363 \frac{d\,}{dr}
364 \left( {\frac{1}{r} - \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} }
365 \right) = \left(- \frac{1}{r^2} + \frac{1}{r_c^2}
366 \right) .
367 \end{equation}
368 which clearly vanishes as the $r$ approaches the cutoff radius. There
369 are a number of ways to generalize this derivative shift for
370 higher-order multipoles. Below, we present two methods, one based on
371 higher-order Taylor series for $r$ near $r_c$, and the other based on
372 linear shift of the kernel gradients at the cutoff itself.
373
374 \subsection{Taylor-shifted force (TSF) electrostatics}
375 In the Taylor-shifted force (TSF) method, the procedure that we follow
376 is based on a Taylor expansion containing the same number of
377 derivatives required for each force term to vanish at the cutoff. For
378 example, the quadrupole-quadrupole interaction energy requires four
379 derivatives of the kernel, and the force requires one additional
380 derivative. For quadrupole-quadrupole interactions, we therefore
381 require shifted energy expressions that include up to $(r-r_c)^5$ so
382 that all energies, forces, and torques are zero as $r \rightarrow
383 r_c$. In each case, we subtract off a function $f_n^{\text{shift}}(r)$
384 from the kernel $f(r)=1/r$. The subscript $n$ indicates the number of
385 derivatives to be taken when deriving a given multipole energy. We
386 choose a function with guaranteed smooth derivatives -- a truncated
387 Taylor series of the function $f(r)$, e.g.,
388 %
389 \begin{equation}
390 f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)}(r_c) .
391 \end{equation}
392 %
393 The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
394 Thus, for $f(r)=1/r$, we find
395 %
396 \begin{equation}
397 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} .
398 \end{equation}
399 %
400 Continuing with the example of a charge $\bf a$ interacting with a
401 dipole $\bf b$, we write
402 %
403 \begin{equation}
404 U_{C_{\bf a}D_{\bf b}}(r)=
405 C_{\bf a} D_{{\bf b}\alpha} \frac {\partial f_1(r) }{\partial r_\alpha}
406 = C_{\bf a} D_{{\bf b}\alpha}
407 \frac {r_\alpha}{r} \frac {\partial f_1(r)}{\partial r} .
408 \end{equation}
409 %
410 The force that dipole $\bf b$ exerts on charge $\bf a$ is
411 %
412 \begin{equation}
413 F_{C_{\bf a}D_{\bf b}\beta} = C_{\bf a} D_{{\bf b}\alpha}
414 \left[ \frac{\delta_{\alpha\beta}}{r} \frac {\partial}{\partial r} +
415 \frac{r_\alpha r_\beta}{r^2}
416 \left( -\frac{1}{r} \frac {\partial} {\partial r}
417 + \frac {\partial ^2} {\partial r^2} \right) \right] f_1(r) .
418 \end{equation}
419 %
420 For undamped coulombic interactions, $f(r)=1/r$, we find
421 %
422 \begin{equation}
423 F_{C_{\bf a}D_{\bf b}\beta} =
424 \frac{C_{\bf a} D_{{\bf b}\beta}}{r}
425 \left[ -\frac{1}{r^2}+\frac{1}{r_c^2}-\frac{2(r-r_c)}{r_c^3} \right]
426 +C_{\bf a} D_{{\bf b}\alpha}r_\alpha r_\beta
427 \left[ \frac{3}{r^5}-\frac{3}{r^3r_c^2} \right] .
428 \end{equation}
429 %
430 This expansion shows the expected $1/r^3$ dependence of the force.
431
432 In general, we can write
433 %
434 \begin{equation}
435 U= (\text{prefactor}) (\text{derivatives}) f_n(r)
436 \label{generic}
437 \end{equation}
438 %
439 with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
440 charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
441 $n=4$ for quadrupole-quadrupole. For example, in
442 quadrupole-quadrupole interactions for which the $\text{prefactor}$ is
443 $Q_{{\bf a}\alpha\beta}Q_{{\bf b}\gamma\delta}$, the derivatives are
444 $\partial^4/\partial r_\alpha \partial r_\beta \partial
445 r_\gamma \partial r_\delta$, with implied summation combining the
446 space indices.
447
448 In the formulas presented in the tables below, the placeholder
449 function $f(r)$ is used to represent the electrostatic kernel (either
450 damped or undamped). The main functions that go into the force and
451 torque terms, $g_n(r), h_n(r), s_n(r), \mathrm{~and~} t_n(r)$ are
452 successive derivatives of the shifted electrostatic kernel, $f_n(r)$
453 of the same index $n$. The algebra required to evaluate energies,
454 forces and torques is somewhat tedious, so only the final forms are
455 presented in tables \ref{tab:tableenergy} and \ref{tab:tableFORCE}.
456 One of the principal findings of our work is that the individual
457 orientational contributions to the various multipole-multipole
458 interactions must be treated with distinct radial functions, but each
459 of these contributions is independently force shifted at the cutoff
460 radius.
461
462 \subsection{Gradient-shifted force (GSF) electrostatics}
463 The second, and conceptually simpler approach to force-shifting
464 maintains only the linear $(r-r_c)$ term in the truncated Taylor
465 expansion, and has a similar interaction energy for all multipole
466 orders:
467 \begin{equation}
468 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
469 U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
470 \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert _{r_c} \right]
471 \label{generic2}
472 \end{equation}
473 where the sum describes a separate force-shifting that is applied to
474 each orientational contribution to the energy. Both the potential and
475 the gradient for force shifting are evaluated for an image multipole
476 projected onto the surface of the cutoff sphere (see fig
477 \ref{fig:shiftedMultipoles}). The image multipole retains the
478 orientation ($\hat{\mathbf{b}}$) of the interacting multipole. No
479 higher order terms $(r-r_c)^n$ appear. The primary difference between
480 the TSF and GSF methods is the stage at which the Taylor Series is
481 applied; in the Taylor-shifted approach, it is applied to the kernel
482 itself. In the Gradient-shifted approach, it is applied to individual
483 radial interactions terms in the multipole expansion. Energies from
484 this method thus have the general form:
485 \begin{equation}
486 U= \sum (\text{angular factor}) (\text{radial factor}).
487 \label{generic3}
488 \end{equation}
489
490 Functional forms for both methods (TSF and GSF) can both be summarized
491 using the form of Eq.~(\ref{generic3}). The basic forms for the
492 energy, force, and torque expressions are tabulated for both shifting
493 approaches below -- for each separate orientational contribution, only
494 the radial factors differ between the two methods.
495
496 \subsection{\label{sec:level2}Body and space axes}
497 Although objects $\bf a$ and $\bf b$ rotate during a molecular
498 dynamics (MD) simulation, their multipole tensors remain fixed in
499 body-frame coordinates. While deriving force and torque expressions,
500 it is therefore convenient to write the energies, forces, and torques
501 in intermediate forms involving the vectors of the rotation matrices.
502 We denote body axes for objects $\bf a$ and $\bf b$ using unit vectors
503 $\hat{a}_m$ and $\hat{b}_m$, respectively, with the index $m=(123)$.
504 In a typical simulation , the initial axes are obtained by
505 diagonalizing the moment of inertia tensors for the objects. (N.B.,
506 the body axes are generally {\it not} the same as those for which the
507 quadrupole moment is diagonal.) The rotation matrices are then
508 propagated during the simulation.
509
510 The rotation matrices $\hat{\mathbf {a}}$ and $\hat{\mathbf {b}}$ can be
511 expressed using these unit vectors:
512 \begin{eqnarray}
513 \hat{\mathbf {a}} =
514 \begin{pmatrix}
515 \hat{a}_1 \\
516 \hat{a}_2 \\
517 \hat{a}_3
518 \end{pmatrix}, \qquad
519 \hat{\mathbf {b}} =
520 \begin{pmatrix}
521 \hat{b}_1 \\
522 \hat{b}_2 \\
523 \hat{b}_3
524 \end{pmatrix}
525 \end{eqnarray}
526 %
527 These matrices convert from space-fixed $(xyz)$ to body-fixed $(123)$
528 coordinates.
529
530 Allen and Germano,\cite{Allen:2006fk} following earlier work by Price
531 {\em et al.},\cite{Price:1984fk} showed that if the interaction
532 energies are written explicitly in terms of $\hat{r}$ and the body
533 axes ($\hat{a}_m$, $\hat{b}_n$) :
534 %
535 \begin{equation}
536 U(r, \{\hat{a}_m \cdot \hat{r} \},
537 \{\hat{b}_n\cdot \hat{r} \},
538 \{\hat{a}_m \cdot \hat{b}_n \}) .
539 \label{ugeneral}
540 \end{equation}
541 %
542 the forces come out relatively cleanly,
543 %
544 \begin{equation}
545 \mathbf{F}_{\bf a}=-\mathbf{F}_{\bf b} = \frac{\partial U}{\partial \mathbf{r}}
546 = \frac{\partial U}{\partial r} \hat{r}
547 + \sum_m \left[
548 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
549 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
550 + \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
551 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
552 \right] \label{forceequation}.
553 \end{equation}
554
555 The torques can also be found in a relatively similar
556 manner,
557 %
558 \begin{eqnarray}
559 \mathbf{\tau}_{\bf a} =
560 \sum_m
561 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
562 ( \hat{r} \times \hat{a}_m )
563 -\sum_{mn}
564 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
565 (\hat{a}_m \times \hat{b}_n) \\
566 %
567 \mathbf{\tau}_{\bf b} =
568 \sum_m
569 \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
570 ( \hat{r} \times \hat{b}_m)
571 +\sum_{mn}
572 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
573 (\hat{a}_m \times \hat{b}_n) .
574 \end{eqnarray}
575
576 Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_b $
577 is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk}
578 We also made use of the identities,
579 %
580 \begin{align}
581 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
582 =& \frac{1}{r} \left( \hat{a}_m - (\hat{a}_m \cdot \hat{r})\hat{r}
583 \right) \\
584 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
585 =& \frac{1}{r} \left( \hat{b}_m - (\hat{b}_m \cdot \hat{r})\hat{r}
586 \right) .
587 \end{align}
588
589 Many of the multipole contractions required can be written in one of
590 three equivalent forms using the unit vectors $\hat{r}$, $\hat{a}_m$,
591 and $\hat{b}_n$. In the torque expressions, it is useful to have the
592 angular-dependent terms available in all three fashions, e.g. for the
593 dipole-dipole contraction:
594 %
595 \begin{equation}
596 \mathbf{D}_{\mathbf {a}} \cdot \mathbf{D}_{\mathbf{b}}
597 = D_{\bf {a}\alpha} D_{\bf {b}\alpha} =
598 \sum_{mn} {D_{\mathbf{a}m} \hat{a}_m \cdot \hat{b}_n D_{\mathbf{b}n}}
599 \end{equation}
600 %
601 The first two forms are written using space coordinates. The first
602 form is standard in the chemistry literature, while the second is
603 expressed using implied summation notation. The third form shows
604 explicit sums over body indices and the dot products now indicate
605 contractions using space indices.
606
607 In computing our force and torque expressions, we carried out most of
608 the work in body coordinates, and have transformed the expressions
609 back to space-frame coordinates, which are reported below. Interested
610 readers may consult the supplemental information for this paper for
611 the intermediate body-frame expressions.
612
613 \subsection{The Self-Interaction \label{sec:selfTerm}}
614
615 In addition to cutoff-sphere neutralization, the Wolf
616 summation~\cite{Wolf99} and the damped shifted force (DSF)
617 extension~\cite{Fennell:2006zl} also included self-interactions that
618 are handled separately from the pairwise interactions between
619 sites. The self-term is normally calculated via a single loop over all
620 sites in the system, and is relatively cheap to evaluate. The
621 self-interaction has contributions from two sources.
622
623 First, the neutralization procedure within the cutoff radius requires
624 a contribution from a charge opposite in sign, but equal in magnitude,
625 to the central charge, which has been spread out over the surface of
626 the cutoff sphere. For a system of undamped charges, the total
627 self-term is
628 \begin{equation}
629 V_\textrm{self} = - \frac{1}{r_c} \sum_{{\bf a}=1}^N C_{\bf a}^{2}
630 \label{eq:selfTerm}
631 \end{equation}
632
633 Second, charge damping with the complementary error function is a
634 partial analogy to the Ewald procedure which splits the interaction
635 into real and reciprocal space sums. The real space sum is retained
636 in the Wolf and DSF methods. The reciprocal space sum is first
637 minimized by folding the largest contribution (the self-interaction)
638 into the self-interaction from charge neutralization of the damped
639 potential. The remainder of the reciprocal space portion is then
640 discarded (as this contributes the largest computational cost and
641 complexity to the Ewald sum). For a system containing only damped
642 charges, the complete self-interaction can be written as
643 \begin{equation}
644 V_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} +
645 \frac{\alpha}{\sqrt{\pi}} \right) \sum_{{\bf a}=1}^N
646 C_{\bf a}^{2}.
647 \label{eq:dampSelfTerm}
648 \end{equation}
649
650 The extension of DSF electrostatics to point multipoles requires
651 treatment of {\it both} the self-neutralization and reciprocal
652 contributions to the self-interaction for higher order multipoles. In
653 this section we give formulae for these interactions up to quadrupolar
654 order.
655
656 The self-neutralization term is computed by taking the {\it
657 non-shifted} kernel for each interaction, placing a multipole of
658 equal magnitude (but opposite in polarization) on the surface of the
659 cutoff sphere, and averaging over the surface of the cutoff sphere.
660 Because the self term is carried out as a single sum over sites, the
661 reciprocal-space portion is identical to half of the self-term
662 obtained by Smith and Aguado and Madden for the application of the
663 Ewald sum to multipoles.\cite{Smith82,Smith98,Aguado03} For a given
664 site which posesses a charge, dipole, and multipole, both types of
665 contribution are given in table \ref{tab:tableSelf}.
666
667 \begin{table*}
668 \caption{\label{tab:tableSelf} Self-interaction contributions for
669 site ({\bf a}) that has a charge $(C_{\bf a})$, dipole
670 $(\mathbf{D}_{\bf a})$, and quadrupole $(\mathbf{Q}_{\bf a})$}
671 \begin{ruledtabular}
672 \begin{tabular}{lccc}
673 Multipole order & Summed Quantity & Self-neutralization & Reciprocal \\ \hline
674 Charge & $C_{\bf a}^2$ & $-f(r_c)$ & $-\frac{\alpha}{\sqrt{\pi}}$ \\
675 Dipole & $|\mathbf{D}_{\bf a}|^2$ & $\frac{1}{3} \left( h(r_c) +
676 \frac{2 g(r_c)}{r_c} \right)$ & $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$\\
677 Quadrupole & $2 \mathbf{Q}_{\bf a}:\mathbf{Q}_{\bf a} + \text{Tr}(\mathbf{Q}_{\bf a})^2$ &
678 $- \frac{1}{15} \left( t(r_c)+ \frac{4 s(r_c)}{r_c} \right)$ &
679 $-\frac{4 \alpha^5}{5 \sqrt{\pi}}$ \\
680 Charge-Quadrupole & $-2 C_{\bf a} \text{Tr}(\mathbf{Q}_{\bf a})$ & $\frac{1}{3} \left(
681 h(r_c) + \frac{2 g(r_c)}{r_c} \right)$& $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$ \\
682 \end{tabular}
683 \end{ruledtabular}
684 \end{table*}
685
686 For sites which simultaneously contain charges and quadrupoles, the
687 self-interaction includes a cross-interaction between these two
688 multipole orders. Symmetry prevents the charge-dipole and
689 dipole-quadrupole interactions from contributing to the
690 self-interaction. The functions that go into the self-neutralization
691 terms, $g(r), h(r), s(r), \mathrm{~and~} t(r)$ are successive
692 derivatives of the electrostatic kernel, $f(r)$ (either the undamped
693 $1/r$ or the damped $B_0(r)=\mathrm{erfc}(\alpha r)/r$ function) that
694 have been evaluated at the cutoff distance. For undamped
695 interactions, $f(r_c) = 1/r_c$, $g(r_c) = -1/r_c^{2}$, and so on. For
696 damped interactions, $f(r_c) = B_0(r_c)$, $g(r_c) = B_0'(r_c)$, and so
697 on. Appendix \ref{SmithFunc} contains recursion relations that allow
698 rapid evaluation of these derivatives.
699
700 \section{Interaction energies, forces, and torques}
701 The main result of this paper is a set of expressions for the
702 energies, forces and torques (up to quadrupole-quadrupole order) that
703 work for both the Taylor-shifted and Gradient-shifted approximations.
704 These expressions were derived using a set of generic radial
705 functions. Without using the shifting approximations mentioned above,
706 some of these radial functions would be identical, and the expressions
707 coalesce into the familiar forms for unmodified multipole-multipole
708 interactions. Table \ref{tab:tableenergy} maps between the generic
709 functions and the radial functions derived for both the Taylor-shifted
710 and Gradient-shifted methods. The energy equations are written in
711 terms of lab-frame representations of the dipoles, quadrupoles, and
712 the unit vector connecting the two objects,
713
714 % Energy in space coordinate form ----------------------------------------------------------------------------------------------
715 %
716 %
717 % u ca cb
718 %
719 \begin{align}
720 U_{C_{\bf a}C_{\bf b}}(r)=&
721 C_{\bf a} C_{\bf b} v_{01}(r) \label{uchch}
722 \\
723 %
724 % u ca db
725 %
726 U_{C_{\bf a}D_{\bf b}}(r)=&
727 C_{\bf a} \left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right) v_{11}(r)
728 \label{uchdip}
729 \\
730 %
731 % u ca qb
732 %
733 U_{C_{\bf a}Q_{\bf b}}(r)=& C_{\bf a } \Bigl[ \text{Tr}Q_{\bf b}
734 v_{21}(r) + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot
735 \hat{r} \right) v_{22}(r) \Bigr]
736 \label{uchquad}
737 \\
738 %
739 % u da cb
740 %
741 %U_{D_{\bf a}C_{\bf b}}(r)=&
742 %-\frac{C_{\bf b}}{4\pi \epsilon_0}
743 %\left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right) v_{11}(r) \label{udipch}
744 %\\
745 %
746 % u da db
747 %
748 U_{D_{\bf a}D_{\bf b}}(r)=&
749 -\Bigr[ \left( \mathbf{D}_{\mathbf {a}} \cdot
750 \mathbf{D}_{\mathbf{b}} \right) v_{21}(r)
751 +\left( \mathbf{D}_{\mathbf {a}} \cdot \hat{r} \right)
752 \left( \mathbf{D}_{\mathbf {b}} \cdot \hat{r} \right)
753 v_{22}(r) \Bigr]
754 \label{udipdip}
755 \\
756 %
757 % u da qb
758 %
759 \begin{split}
760 % 1
761 U_{D_{\bf a}Q_{\bf b}}(r) =&
762 -\Bigl[
763 \text{Tr}\mathbf{Q}_{\mathbf{b}}
764 \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right)
765 +2 ( \mathbf{D}_{\mathbf{a}} \cdot
766 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r} ) \Bigr] v_{31}(r) \\
767 % 2
768 &- \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right)
769 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} \right) v_{32}(r)
770 \label{udipquad}
771 \end{split}
772 \\
773 %
774 % u qa cb
775 %
776 %U_{Q_{\bf a}C_{\bf b}}(r)=&
777 %\frac{C_{\bf b }}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\bf a} v_{21}(r)
778 %\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{22}(r) \Bigr]
779 %\label{uquadch}
780 %\\
781 %
782 % u qa db
783 %
784 %\begin{split}
785 %1
786 %U_{Q_{\bf a}D_{\bf b}}(r)=&
787 %\frac{1}{4\pi \epsilon_0} \Bigl[
788 %\text{Tr}\mathbf{Q}_{\mathbf{a}}
789 %\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
790 %+2 ( \mathbf{D}_{\mathbf{b}} \cdot
791 %\mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r)\\
792 % 2
793 %&+\frac{1}{4\pi \epsilon_0}
794 %\left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
795 %\left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) v_{32}(r)
796 %\label{uquaddip}
797 %\end{split}
798 %\\
799 %
800 % u qa qb
801 %
802 \begin{split}
803 %1
804 U_{Q_{\bf a}Q_{\bf b}}(r)=&
805 \Bigl[
806 \text{Tr} \mathbf{Q}_{\mathbf{a}} \text{Tr} \mathbf{Q}_{\mathbf{b}}
807 +2
808 \mathbf{Q}_{\mathbf{a}} : \mathbf{Q}_{\mathbf{b}} \Bigr] v_{41}(r)
809 \\
810 % 2
811 &+\Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}}
812 \left( \hat{r} \cdot
813 \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} \right)
814 +\text{Tr}\mathbf{Q}_{\mathbf{b}}
815 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}}
816 \cdot \hat{r} \right) +4 (\hat{r} \cdot
817 \mathbf{Q}_{{\mathbf a}}\cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
818 \Bigr] v_{42}(r)
819 \\
820 % 4
821 &+
822 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right)
823 \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} \right) v_{43}(r).
824 \label{uquadquad}
825 \end{split}
826 \end{align}
827 %
828 Note that the energies of multipoles on site $\mathbf{b}$ interacting
829 with those on site $\mathbf{a}$ can be obtained by swapping indices
830 along with the sign of the intersite vector, $\hat{r}$.
831
832 %
833 %
834 % TABLE of radial functions ----------------------------------------------------------------------------------------------------------------
835 %
836
837 \begin{sidewaystable}
838 \caption{\label{tab:tableenergy}Radial functions used in the energy
839 and torque equations. The $f, g, h, s, t, \mathrm{and} u$
840 functions used in this table are defined in Appendices B and C.}
841 \begin{tabular}{|c|c|l|l|} \hline
842 Generic&Bare Coulomb&Taylor-Shifted&Gradient-Shifted
843 \\ \hline
844 %
845 %
846 %
847 %Ch-Ch&
848 $v_{01}(r)$ &
849 $\frac{1}{r}$ &
850 $f_0(r)$ &
851 $f(r)-f(r_c)-(r-r_c)g(r_c)$
852 \\
853 %
854 %
855 %
856 %Ch-Di&
857 $v_{11}(r)$ &
858 $-\frac{1}{r^2}$ &
859 $g_1(r)$ &
860 $g(r)-g(r_c)-(r-r_c)h(r_c)$ \\
861 %
862 %
863 %
864 %Ch-Qu/Di-Di&
865 $v_{21}(r)$ &
866 $-\frac{1}{r^3} $ &
867 $\frac{g_2(r)}{r} $ &
868 $\frac{g(r)}{r}-\frac{g(r_c)}{r_c} -(r-r_c)
869 \left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right)$ \\
870 $v_{22}(r)$ &
871 $\frac{3}{r^3} $ &
872 $\left(-\frac{g_2(r)}{r} + h_2(r) \right)$ &
873 $\left(-\frac{g(r)}{r}+h(r) \right)
874 -\left(-\frac{g(r_c)}{r_c}+h(r_c) \right)$ \\
875 &&& $ ~~~-(r-r_c) \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$
876 \\
877 %
878 %
879 %
880 %Di-Qu &
881 $v_{31}(r)$ &
882 $\frac{3}{r^4} $ &
883 $\left(-\frac{g_3(r)}{r^2} + \frac{h_3(r)}{r} \right)$ &
884 $\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r} \right)
885 -\left(-\frac{g(r_c)}{r_c^2}+\frac{h(r_c)}{r_c} \right) $\\
886 &&&$ ~~~ -(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)$
887 \\
888 %
889 $v_{32}(r)$ &
890 $-\frac{15}{r^4} $ &
891 $\left( \frac{3g_3(r)}{r^2} - \frac{3h_3(r)}{r} + s_3(r) \right)$ &
892 $\left( \frac{3g(r)}{r^2} - \frac{3h(r)}{r} + s(r) \right)
893 - \left( \frac{3g(r_c)}{r_c^2} - \frac{3h(r_c)}{r_c} + s(r_c) \right)$ \\
894 &&&$ ~~~ -(r-r_c) \left( \frac{-6g(r_c)}{r_c^3}+\frac{6h(r_c)}{r_c^2}-\frac{3s(r_c)}{r_c}+t(r_c) \right)$
895 \\
896 %
897 %
898 %
899 %Qu-Qu&
900 $v_{41}(r)$ &
901 $\frac{3}{r^5} $ &
902 $\left(-\frac{g_4(r)}{r^3} +\frac{h_4(r)}{r^2} \right) $ &
903 $\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)
904 - \left( -\frac{g(r_c)}{r_c^3} + \frac{h(r_c)}{r_c^2} \right)$ \\
905 &&&$ ~~~ -(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)$
906 \\
907 % 2
908 $v_{42}(r)$ &
909 $- \frac{15}{r^5} $ &
910 $\left( \frac{3g_4(r)}{r^3} - \frac{3h_4(r)}{r^2}+\frac{s_4(r)}{r} \right)$ &
911 $\left( \frac{3g(r)}{r^3} - \frac{3h(r)}{r^2}+\frac{s(r)}{r} \right)
912 -\left( \frac{3g(r_c)}{r_c^3} - \frac{3h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ \\
913 &&&$ ~~~ -(r-r_c) \left(- \frac{9g(r_c)}{r_c^4}+\frac{9h(r_c)}{r_c^3}
914 -\frac{4s(r_c)}{r_c^2} + \frac{t(r_c)}{r_c}\right)$
915 \\
916 % 3
917 $v_{43}(r)$ &
918 $ \frac{105}{r^5} $ &
919 $\left(-\frac{15g_4(r)}{r^3}+\frac{15h_4(r)}{r^2}-\frac{6s_4(r)}{r} + t_4(r)\right) $ &
920 $\left(-\frac{15g(r)}{r^3}+\frac{15h(r)}{r^2}-\frac{6s(r)}{r} + t(r)\right)$ \\
921 &&&$~~~ -\left(-\frac{15g(r_c)}{r_c^3}+\frac{15h(r_c)}{r_c^2}-\frac{6s(r_c)}{r_c} + t(r_c)\right)$ \\
922 &&&$~~~ -(r-r_c)\left(\frac{45g(r_c)}{r_c^4}-\frac{45h(r_c)}{r_c^3}+\frac{21s(r_c)}{r_c^2}
923 -\frac{6t(r_c)}{r_c}+u(r_c) \right)$ \\ \hline
924 \end{tabular}
925 \end{sidewaystable}
926 %
927 %
928 % FORCE TABLE of radial functions ----------------------------------------------------------------------------------------------------------------
929 %
930
931 \begin{sidewaystable}
932 \caption{\label{tab:tableFORCE}Radial functions used in the force equations.}
933 \begin{tabular}{|c|c|l|l|} \hline
934 Function&Definition&Taylor-Shifted&Gradient-Shifted
935 \\ \hline
936 %
937 %
938 %
939 $w_a(r)$&
940 $\frac{d v_{01}}{dr}$&
941 $g_0(r)$&
942 $g(r)-g(r_c)$ \\
943 %
944 %
945 $w_b(r)$ &
946 $\frac{d v_{11}}{dr} - \frac{v_{11}(r)}{r} $&
947 $\left( -\frac{g_1(r)}{r}+h_1(r) \right)$ &
948 $h(r)- h(r_c) - \frac{v_{11}(r)}{r} $ \\
949 %
950 $w_c(r)$ &
951 $\frac{v_{11}(r)}{r}$ &
952 $\frac{g_1(r)}{r} $ &
953 $\frac{v_{11}(r)}{r}$\\
954 %
955 %
956 $w_d(r)$&
957 $\frac{d v_{21}}{dr}$&
958 $\left( -\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right) $ &
959 $\left( -\frac{g(r)}{r^2} + \frac{h(r)}{r} \right)
960 -\left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right) $ \\
961 %
962 $w_e(r)$ &
963 $\left(-\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right)$ &
964 $\frac{v_{22}(r)}{r}$ &
965 $\frac{v_{22}(r)}{r}$ \\
966 %
967 %
968 $w_f(r)$&
969 $\frac{d v_{22}}{dr} - \frac{2v_{22}(r)}{r}$&
970 $\left( \frac{3g_2(r)}{r^2}-\frac{3h_2(r)}{r}+s_2(r) \right)$ &
971 $ \left( \frac{g(r)}{r^2}-\frac{h(r)}{r}+s(r) \right) $ \\
972 &&& $ ~~~- \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c)
973 \right)-\frac{2v_{22}(r)}{r}$\\
974 %
975 $w_g(r)$&
976 $\frac{v_{31}(r)}{r}$&
977 $ \left( -\frac{g_3(r)}{r^3}+\frac{h_3(r)}{r^2} \right)$&
978 $\frac{v_{31}(r)}{r}$\\
979 %
980 $w_h(r)$ &
981 $\frac{d v_{31}}{dr} -\frac{v_{31}(r)}{r}$&
982 $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ &
983 $ \left(\frac{2g(r)}{r^3} -\frac{2h(r)}{r^2} +\frac{s(r)}{r} \right) - \left(\frac{2g(r_c)}{r_c^3} -\frac{2h(r_c)}{r_c^2} +\frac{s(r_c)}{r_c} \right) $ \\
984 &&& $ ~~~ -\frac{v_{31}(r)}{r}$ \\
985 % 2
986 $w_i(r)$ &
987 $\frac{v_{32}(r)}{r}$ &
988 $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $ &
989 $\frac{v_{32}(r)}{r}$\\
990 %
991 $w_j(r)$ &
992 $\frac{d v_{32}}{dr} - \frac{3v_{32}}{r}$&
993 $\left(\frac{-15g_3(r)}{r^3} + \frac{15h_3(r)}{r^2} - \frac{6s_3(r)}{r} + t_3(r) \right) $ &
994 $\left(\frac{-6g(r)}{r^3} +\frac{6h(r)}{r^2} -\frac{3s(r)}{r} +t(r) \right)$ \\
995 &&& $~~~-\left(\frac{-6g(_cr)}{r_c^3} +\frac{6h(r_c)}{r_c^2}
996 -\frac{3s(r_c)}{r_c} +t(r_c) \right) -\frac{3v_{32}}{r}$ \\
997 %
998 $w_k(r)$ &
999 $\frac{d v_{41}}{dr} $ &
1000 $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ &
1001 $\left(\frac{3g(r)}{r^4} -\frac{3h(r)}{r^3} +\frac{s(r)}{r^2} \right)
1002 -\left(\frac{3g(r_c)}{r_c^4} -\frac{3h(r_c)}{r_c^3} +\frac{s(r_c)}{r_c^2} \right)$ \\
1003 %
1004 $w_l(r)$ &
1005 $\frac{d v_{42}}{dr} -\frac{2v_{42}(r)}{r}$ &
1006 $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1007 $\left(-\frac{9g(r)}{r^4} +\frac{9h(r)}{r^3} -\frac{4s(r)}{r^2} +\frac{t(r)}{r} \right)$ \\
1008 &&& $~~~ -\left(-\frac{9g(r_c)}{r_c^4} +\frac{9h(r_c)}{r_c^3} -\frac{4s(r_c)}{r_c^2} +\frac{t(r_c)}{r_c} \right)
1009 -\frac{2v_{42}(r)}{r}$\\
1010 %
1011 $w_m(r)$ &
1012 $\frac{d v_{43}}{dr} -\frac{4v_{43}(r)}{r}$&
1013 $\left(\frac{105g_4(r)}{r^4} - \frac{105h_4(r)}{r^3} + \frac{45s_4(r)}{r^2} - \frac{10t_4(r)}{r} +u_4(r) \right)$ &
1014 $\left(\frac{45g(r)}{r^4} -\frac{45h(r)}{r^3} +\frac{21s(r)}{r^2} -\frac{6t(r)}{r} +u(r) \right)$\\
1015 &&& $~~~- \left(\frac{45g(r_c)}{r_c^4} -\frac{45h(r_c)}{r_c^3}
1016 +\frac{21s(r_c)}{r_c^2} -\frac{6t(r_c)}{r_c} +u(r_c) \right) $\\
1017 &&& $~~~-\frac{4v_{43}(r)}{r}$ \\
1018 %
1019 $w_n(r)$ &
1020 $\frac{v_{42}(r)}{r}$ &
1021 $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2} \right)$ &
1022 $\frac{v_{42}(r)}{r}$\\
1023 %
1024 $w_o(r)$ &
1025 $\frac{v_{43}(r)}{r}$&
1026 $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1027 $\frac{v_{43}(r)}{r}$ \\ \hline
1028 %
1029
1030 \end{tabular}
1031 \end{sidewaystable}
1032 %
1033 %
1034 %
1035
1036 \subsection{Forces}
1037 The force on object $\bf{a}$, $\mathbf{F}_{\bf a}$, due to object
1038 $\bf{b}$ is the negative of the force on $\bf{b}$ due to $\bf{a}$. For
1039 a simple charge-charge interaction, these forces will point along the
1040 $\pm \hat{r}$ directions, where $\mathbf{r}=\mathbf{r}_b -
1041 \mathbf{r}_a $. Thus
1042 %
1043 \begin{equation}
1044 F_{\bf a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_{\bf a}C_{\bf b}}}{\partial r}
1045 \quad \text{and} \quad F_{\bf b \alpha}
1046 = - \hat{r}_\alpha \frac{\partial U_{C_{\bf a}C_{\bf b}}} {\partial r} .
1047 \end{equation}
1048 %
1049 We list below the force equations written in terms of lab-frame
1050 coordinates. The radial functions used in the two methods are listed
1051 in Table \ref{tab:tableFORCE}
1052 %
1053 %SPACE COORDINATES FORCE EQUATIONS
1054 %
1055 % **************************************************************************
1056 % f ca cb
1057 %
1058 \begin{align}
1059 \mathbf{F}_{{\bf a}C_{\bf a}C_{\bf b}} =&
1060 C_{\bf a} C_{\bf b} w_a(r) \hat{r} \\
1061 %
1062 %
1063 %
1064 \mathbf{F}_{{\bf a}C_{\bf a}D_{\bf b}} =&
1065 C_{\bf a} \Bigl[
1066 \left( \hat{r} \cdot \mathbf{D}_{\mathbf{b}} \right)
1067 w_b(r) \hat{r}
1068 + \mathbf{D}_{\mathbf{b}} w_c(r) \Bigr] \\
1069 %
1070 %
1071 %
1072 \mathbf{F}_{{\bf a}C_{\bf a}Q_{\bf b}} =&
1073 C_{\bf a } \Bigr[
1074 \text{Tr}\mathbf{Q}_{\bf b} w_d(r) \hat{r}
1075 + 2 \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} w_e(r)
1076 + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}
1077 \right) w_f(r) \hat{r} \Bigr] \\
1078 %
1079 %
1080 %
1081 % \begin{equation}
1082 % \mathbf{F}_{{\bf a}D_{\bf a}C_{\bf b}} =
1083 % -C_{\bf{b}} \Bigl[
1084 % \left( \hat{r} \cdot \mathbf{D}_{\mathbf{a}} \right) w_b(r) \hat{r}
1085 % + \mathbf{D}_{\mathbf{a}} w_c(r) \Bigr]
1086 % \end{equation}
1087 %
1088 %
1089 %
1090 \begin{split}
1091 \mathbf{F}_{{\bf a}D_{\bf a}D_{\bf b}} =&
1092 - \mathbf{D}_{\mathbf {a}} \cdot \mathbf{D}_{\mathbf{b}} w_d(r) \hat{r}
1093 + \left( \mathbf{D}_{\mathbf {a}}
1094 \left( \mathbf{D}_{\mathbf{b}} \cdot \hat{r} \right)
1095 + \mathbf{D}_{\mathbf {b}} \left( \mathbf{D}_{\mathbf{a}} \cdot \hat{r} \right) \right) w_e(r)\\
1096 % 2
1097 & - \left( \hat{r} \cdot \mathbf{D}_{\mathbf {a}} \right)
1098 \left( \hat{r} \cdot \mathbf{D}_{\mathbf {b}} \right) w_f(r) \hat{r}
1099 \end{split}\\
1100 %
1101 %
1102 %
1103 \begin{split}
1104 \mathbf{F}_{{\bf a}D_{\bf a}Q_{\bf b}} =& - \Bigl[
1105 \text{Tr}\mathbf{Q}_{\mathbf{b}} \mathbf{ D}_{\mathbf{a}}
1106 +2 \mathbf{D}_{\mathbf{a}} \cdot
1107 \mathbf{Q}_{\mathbf{b}} \Bigr] w_g(r)
1108 - \Bigl[
1109 \text{Tr}\mathbf{Q}_{\mathbf{b}}
1110 \left( \hat{r} \cdot \mathbf{D}_{\mathbf{a}} \right)
1111 +2 ( \mathbf{D}_{\mathbf{a}} \cdot
1112 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) \Bigr] w_h(r) \hat{r} \\
1113 % 3
1114 & - \Bigl[\mathbf{ D}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1115 +2 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}} ) (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} ) \Bigr]
1116 w_i(r)
1117 % 4
1118 -
1119 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}} )
1120 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) w_j(r) \hat{r} \end{split} \\
1121 %
1122 %
1123 % \begin{equation}
1124 % \mathbf{F}_{{\bf a}Q_{\bf a}C_{\bf b}} =
1125 % \frac{C_{\bf b }}{4\pi \epsilon_0} \Bigr[
1126 % \text{Tr}\mathbf{Q}_{\bf a} w_d(r) \hat{r}
1127 % + 2 \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} w_e(r)
1128 % + \left( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r} \right) w_f(r) \hat{r} \Bigr]
1129 % \end{equation}
1130 % %
1131 % \begin{equation}
1132 % \begin{split}
1133 % \mathbf{F}_{{\bf a}Q_{\bf a}D_{\bf b}} =
1134 % &\frac{1}{4\pi \epsilon_0} \Bigl[
1135 % \text{Tr}\mathbf{Q}_{\mathbf{a}} \mathbf{D}_{\mathbf{b}}
1136 % +2 \mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} \Bigr] w_g(r)
1137 % % 2
1138 % + \frac{1}{4\pi \epsilon_0} \Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}}
1139 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1140 % +2 (\mathbf{D}_{\mathbf{b}} \cdot
1141 % \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] w_h(r) \hat{r} \\
1142 % % 3
1143 % &+ \frac{1}{4\pi \epsilon_0} \Bigl[ \mathbf{D}_{\mathbf{b}}
1144 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1145 % +2 (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1146 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \Bigr] w_i(r)
1147 % % 4
1148 % +\frac{1}{4\pi \epsilon_0}
1149 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}})
1150 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) w_j(r) \hat{r}
1151 % \end{split}
1152 % \end{equation}
1153 %
1154 %
1155 %
1156 \begin{split}
1157 \mathbf{F}_{{\bf a}Q_{\bf a}Q_{\bf b}} =&
1158 \Bigl[
1159 \text{Tr}\mathbf{Q}_{\mathbf{a}} \text{Tr}\mathbf{Q}_{\mathbf{b}}
1160 + 2 \mathbf{Q}_{\mathbf{a}} : \mathbf{Q}_{\mathbf{b}} \Bigr] w_k(r) \hat{r} \\
1161 % 2
1162 &+ \Bigl[
1163 2\text{Tr}\mathbf{Q}_{\mathbf{b}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} )
1164 + 2\text{Tr}\mathbf{Q}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} )
1165 % 3
1166 +4 (\mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1167 + 4(\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}}) \Bigr] w_n(r) \\
1168 % 4
1169 &+ \Bigl[
1170 \text{Tr}\mathbf{Q}_{\mathbf{a}} (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1171 + \text{Tr}\mathbf{Q}_{\mathbf{b}}
1172 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1173 % 5
1174 +4 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot
1175 \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) \Bigr] w_l(r) \hat{r} \\
1176 %
1177 &+ \Bigl[
1178 + 2 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} )
1179 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1180 %6
1181 +2 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1182 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} ) \Bigr] w_o(r) \\
1183 % 7
1184 &+
1185 (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1186 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}} \cdot \hat{r}) w_m(r) \hat{r} \end{split}
1187 \end{align}
1188 Note that the forces for higher multipoles on site $\mathbf{a}$
1189 interacting with those of lower order on site $\mathbf{b}$ can be
1190 obtained by swapping indices in the expressions above.
1191
1192 %
1193 % Torques SECTION -----------------------------------------------------------------------------------------
1194 %
1195 \subsection{Torques}
1196
1197 %
1198 The torques for both the Taylor-Shifted as well as Gradient-Shifted
1199 methods are given in space-frame coordinates:
1200 %
1201 %
1202 \begin{align}
1203 \mathbf{\tau}_{{\bf b}C_{\bf a}D_{\bf b}} =&
1204 C_{\bf a} (\hat{r} \times \mathbf{D}_{\mathbf{b}}) v_{11}(r) \\
1205 %
1206 %
1207 %
1208 \mathbf{\tau}_{{\bf b}C_{\bf a}Q_{\bf b}} =&
1209 2C_{\bf a}
1210 \hat{r} \times ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{22}(r) \\
1211 %
1212 %
1213 %
1214 % \begin{equation}
1215 % \mathbf{\tau}_{{\bf a}D_{\bf a}C_{\bf b}} =
1216 % -\frac{C_{\bf b}}{4\pi \epsilon_0}
1217 % (\hat{r} \times \mathbf{D}_{\mathbf{a}}) v_{11}(r)
1218 % \end{equation}
1219 %
1220 %
1221 %
1222 \mathbf{\tau}_{{\bf a}D_{\bf a}D_{\bf b}} =&
1223 \mathbf{D}_{\mathbf {a}} \times \mathbf{D}_{\mathbf{b}} v_{21}(r)
1224 % 2
1225 -
1226 (\hat{r} \times \mathbf{D}_{\mathbf {a}} )
1227 (\hat{r} \cdot \mathbf{D}_{\mathbf {b}} ) v_{22}(r)\\
1228 %
1229 %
1230 %
1231 % \begin{equation}
1232 % \mathbf{\tau}_{{\bf b}D_{\bf a}D_{\bf b}} =
1233 % -\frac{1}{4\pi \epsilon_0} \mathbf{D}_{\mathbf {a}} \times \mathbf{D}_{\mathbf{b}} v_{21}(r)
1234 % % 2
1235 % +\frac{1}{4\pi \epsilon_0}
1236 % (\hat{r} \cdot \mathbf{D}_{\mathbf {a}} )
1237 % (\hat{r} \times \mathbf{D}_{\mathbf {b}} ) v_{22}(r)
1238 % \end{equation}
1239 %
1240 %
1241 %
1242 \mathbf{\tau}_{{\bf a}D_{\bf a}Q_{\bf b}} =&
1243 \Bigl[
1244 -\text{Tr}\mathbf{Q}_{\mathbf{b}}
1245 (\hat{r} \times \mathbf{D}_{\mathbf{a}} )
1246 +2 \mathbf{D}_{\mathbf{a}} \times
1247 (\mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1248 \Bigr] v_{31}(r)
1249 % 3
1250 - (\hat{r} \times \mathbf{D}_{\mathbf{a}} )
1251 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{32}(r)\\
1252 %
1253 %
1254 %
1255 \mathbf{\tau}_{{\bf b}D_{\bf a}Q_{\bf b}} =&
1256 \Bigl[
1257 +2 ( \mathbf{D}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} ) \times
1258 \hat{r}
1259 -2 \mathbf{D}_{\mathbf{a}} \times
1260 (\mathbf{Q}_{\mathbf{b}} \cdot \hat{r})
1261 \Bigr] v_{31}(r)
1262 % 2
1263 +
1264 (\hat{r} \cdot \mathbf{D}_{\mathbf{a}})
1265 (\hat{r} \cdot \mathbf{Q}_{\mathbf{b}}) \times \hat{r} v_{32}(r)\\
1266 %
1267 %
1268 %
1269 % \begin{equation}
1270 % \mathbf{\tau}_{{\bf a}Q_{\bf a}D_{\bf b}} =
1271 % \frac{1}{4\pi \epsilon_0} \Bigl[
1272 % -2 (\mathbf{D}_{\mathbf{b}} \cdot \mathbf{Q}_{\mathbf{a}} ) \times \hat{r}
1273 % +2 \mathbf{D}_{\mathbf{b}} \times
1274 % (\mathbf{Q}_{\mathbf{a}} \cdot \hat{r})
1275 % \Bigr] v_{31}(r)
1276 % % 3
1277 % - \frac{2}{4\pi \epsilon_0}
1278 % (\hat{r} \cdot \mathbf{D}_{\mathbf{b}} )
1279 % (\hat{r} \cdot \mathbf
1280 % {Q}_{{\mathbf a}}) \times \hat{r} v_{32}(r)
1281 % \end{equation}
1282 %
1283 %
1284 %
1285 % \begin{equation}
1286 % \mathbf{\tau}_{{\bf b}Q_{\bf a}D_{\bf b}} =
1287 % \frac{1}{4\pi \epsilon_0} \Bigl[
1288 % \text{Tr}\mathbf{Q}_{\mathbf{a}}
1289 % (\hat{r} \times \mathbf{D}_{\mathbf{b}} )
1290 % +2 \mathbf{D}_{\mathbf{b}} \times
1291 % ( \mathbf{Q}_{\mathbf{a}} \cdot \hat{r}) \Bigr] v_{31}(r)
1292 % % 2
1293 % +\frac{1}{4\pi \epsilon_0}
1294 % (\hat{r} \times \mathbf{D}_{\mathbf{b}} )
1295 % (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r}) v_{32}(r)
1296 % \end{equation}
1297 %
1298 %
1299 %
1300 \begin{split}
1301 \mathbf{\tau}_{{\bf a}Q_{\bf a}Q_{\bf b}} =&
1302 -4
1303 \mathbf{Q}_{{\mathbf a}} \times \mathbf{Q}_{{\mathbf b}}
1304 v_{41}(r) \\
1305 % 2
1306 &+
1307 \Bigl[-2\text{Tr}\mathbf{Q}_{\mathbf{b}}
1308 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \times \hat{r}
1309 +4 \hat{r} \times
1310 ( \mathbf{Q}_{{\mathbf a}} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1311 % 3
1312 -4 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} )\times
1313 ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r} ) \Bigr] v_{42}(r) \\
1314 % 4
1315 &+ 2
1316 \hat{r} \times ( \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1317 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{43}(r) \end{split}\\
1318 %
1319 %
1320 %
1321 \begin{split}
1322 \mathbf{\tau}_{{\bf b}Q_{\bf a}Q_{\bf b}} =
1323 &4
1324 \mathbf{Q}_{{\mathbf a}} \times \mathbf{Q}_{{\mathbf b}} v_{41}(r) \\
1325 % 2
1326 &+ \Bigl[- 2\text{Tr}\mathbf{Q}_{\mathbf{a}}
1327 (\hat{r} \cdot \mathbf{Q}_{{\mathbf b}} ) \times \hat{r}
1328 -4 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot
1329 \mathbf{Q}_{{\mathbf b}} ) \times
1330 \hat{r}
1331 +4 ( \hat{r} \cdot \mathbf{Q}_{{\mathbf a}} ) \times
1332 ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r})
1333 \Bigr] v_{42}(r) \\
1334 % 4
1335 &+2
1336 (\hat{r} \cdot \mathbf{Q}_{{\mathbf a}} \cdot \hat{r})
1337 \hat{r} \times ( \mathbf{Q}_{{\mathbf b}} \cdot \hat{r}) v_{43}(r)\end{split}
1338 \end{align}
1339 %
1340 Here, we have defined the matrix cross product in an identical form
1341 as in Ref. \onlinecite{Smith98}:
1342 \begin{equation}
1343 \left[\mathbf{A} \times \mathbf{B}\right]_\alpha = \sum_\beta
1344 \left[\mathbf{A}_{\alpha+1,\beta} \mathbf{B}_{\alpha+2,\beta}
1345 -\mathbf{A}_{\alpha+2,\beta} \mathbf{B}_{\alpha+2,\beta}
1346 \right]
1347 \end{equation}
1348 where $\alpha+1$ and $\alpha+2$ are regarded as cyclic
1349 permuations of the matrix indices.
1350
1351 All of the radial functions required for torques are identical with
1352 the radial functions previously computed for the interaction energies.
1353 These are tabulated for both shifted force methods in table
1354 \ref{tab:tableenergy}. The torques for higher multipoles on site
1355 $\mathbf{a}$ interacting with those of lower order on site
1356 $\mathbf{b}$ can be obtained by swapping indices in the expressions
1357 above.
1358
1359 \section{Related real-space methods}
1360 One can also formulate a shifted potential,
1361 \begin{equation}
1362 U^{\text{SP}} = U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) -
1363 U(\mathbf{r}_c, \hat{\mathbf{a}}, \hat{\mathbf{b}}),
1364 \label{eq:SP}
1365 \end{equation}
1366 obtained by projecting the image multipole onto the surface of the
1367 cutoff sphere. The shifted potential (SP) can be thought of as a
1368 simple extension to the original Wolf method. The energies and
1369 torques for the SP can be easily obtained by zeroing out the $(r-r_c)$
1370 terms in the final column of table \ref{tab:tableenergy}. SP forces
1371 (which retain discontinuities at the cutoff sphere) can be obtained by
1372 eliminating all functions that depend on $r_c$ in the last column of
1373 table \ref{tab:tableFORCE}. The self-energy contributions to the SP
1374 potential are identical to both the GSF and TSF methods.
1375
1376 \section{Comparison to known multipolar energies}
1377
1378 To understand how these new real-space multipole methods behave in
1379 computer simulations, it is vital to test against established methods
1380 for computing electrostatic interactions in periodic systems, and to
1381 evaluate the size and sources of any errors that arise from the
1382 real-space cutoffs. In this paper we test both TSF and GSF
1383 electrostatics against analytical methods for computing the energies
1384 of ordered multipolar arrays. In the following paper, we test the new
1385 methods against the multipolar Ewald sum for computing the energies,
1386 forces and torques for a wide range of typical condensed-phase
1387 (disordered) systems.
1388
1389 Because long-range electrostatic effects can be significant in
1390 crystalline materials, ordered multipolar arrays present one of the
1391 biggest challenges for real-space cutoff methods. The dipolar
1392 analogues to the Madelung constants were first worked out by Sauer,
1393 who computed the energies of ordered dipole arrays of zero
1394 magnetization and obtained a number of these constants.\cite{Sauer}
1395 This theory was developed more completely by Luttinger and
1396 Tisza\cite{LT,LT2} who tabulated energy constants for the Sauer arrays
1397 and other periodic structures.
1398
1399 To test the new electrostatic methods, we have constructed very large,
1400 $N=$ 16,000~(bcc) arrays of dipoles in the orientations described in
1401 Ref. \onlinecite{LT}. These structures include ``A'' lattices with
1402 nearest neighbor chains of antiparallel dipoles, as well as ``B''
1403 lattices with nearest neighbor strings of antiparallel dipoles if the
1404 dipoles are contained in a plane perpendicular to the dipole direction
1405 that passes through the dipole. We have also studied the minimum
1406 energy structure for the BCC lattice that was found by Luttinger \&
1407 Tisza. The total electrostatic energy for any of the arrays is given
1408 by:
1409 \begin{equation}
1410 E = C N^2 \mu^2
1411 \end{equation}
1412 where $C$ is the energy constant (equivalent to the Madelung
1413 constant), $N$ is the number of dipoles per unit volume, and $\mu$ is
1414 the strength of the dipole. Energy constants (converged to 1 part in
1415 $10^9$) are given in the supplemental information.
1416
1417 For the purposes of testing the energy expressions and the
1418 self-neutralization schemes, the primary quantity of interest is the
1419 analytic energy constant for the perfect arrays. Convergence to these
1420 constants are shown as a function of both the cutoff radius, $r_c$,
1421 and the damping parameter, $\alpha$ in Figs.
1422 \ref{fig:energyConstVsCutoff} and XXX. We have simultaneously tested a
1423 hard cutoff (where the kernel is simply truncated at the cutoff
1424 radius), as well as a shifted potential (SP) form which includes a
1425 potential-shifting and self-interaction term, but does not shift the
1426 forces and torques smoothly at the cutoff radius. The SP method is
1427 essentially an extension of the original Wolf method for multipoles.
1428
1429 \begin{figure}[!htbp]
1430 \includegraphics[width=4.5in]{energyConstVsCutoff}
1431 \caption{Convergence to the analytic energy constants as a function of
1432 cutoff radius (normalized by the lattice constant) for the different
1433 real-space methods. The two crystals shown here are the ``B'' array
1434 for bcc crystals with the dipoles along the 001 direction (upper),
1435 as well as the minimum energy bcc lattice (lower). The analytic
1436 energy constants are shown as a grey dashed line. The left panel
1437 shows results for the undamped kernel ($1/r$), while the damped
1438 error function kernel, $B_0(r)$ was used in the right panel. }
1439 \label{fig:energyConstVsCutoff}
1440 \end{figure}
1441
1442 The Hard cutoff exhibits oscillations around the analytic energy
1443 constants, and converges to incorrect energies when the complementary
1444 error function damping kernel is used. The shifted potential (SP) and
1445 gradient-shifted force (GSF) approximations converge to the correct
1446 energy smoothly by $r_c / 6 a$ even for the undamped case. This
1447 indicates that the correction provided by the self term is required
1448 for obtaining accurate energies. The Taylor-shifted force (TSF)
1449 approximation appears to perturb the potential too much inside the
1450 cutoff region to provide accurate measures of the energy constants.
1451
1452 {\it Quadrupolar} analogues to the Madelung constants were first
1453 worked out by Nagai and Nakamura who computed the energies of selected
1454 quadrupole arrays based on extensions to the Luttinger and Tisza
1455 approach.\cite{Nagai01081960,Nagai01091963} We have compared the
1456 energy constants for the lowest energy configurations for linear
1457 quadrupoles.
1458
1459 In analogy to the dipolar arrays, the total electrostatic energy for
1460 the quadrupolar arrays is:
1461 \begin{equation}
1462 E = C N \frac{3\bar{Q}^2}{4a^5}
1463 \end{equation}
1464 where $a$ is the lattice parameter, and $\bar{Q}$ is the effective
1465 quadrupole moment,
1466 \begin{equation}
1467 \bar{Q}^2 = 4 \left(3 Q : Q - (\text{Tr} Q)^2 \right)
1468 \end{equation}
1469 for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}.
1470 (For the traceless quadrupole tensor, $\Theta = 3 Q - \text{Tr} Q$,
1471 the effective moment, $\bar{Q}^2 = \frac{4}{3} \Theta : \Theta$.)
1472
1473 \section{Conclusion}
1474 We have presented two efficient real-space methods for computing the
1475 interactions between point multipoles. These methods have the benefit
1476 of smoothly truncating the energies, forces, and torques at the cutoff
1477 radius, making them attractive for both molecular dynamics (MD) and
1478 Monte Carlo (MC) simulations. We find that the Gradient-Shifted Force
1479 (GSF) and the Shifted-Potential (SP) methods converge rapidly to the
1480 correct lattice energies for ordered dipolar and quadrupolar arrays,
1481 while the Taylor-Shifted Force (TSF) is too severe an approximation to
1482 provide accurate convergence to lattice energies.
1483
1484 In most cases, GSF can obtain nearly quantitative agreement with the
1485 lattice energy constants with reasonably small cutoff radii. The only
1486 exception we have observed is for crystals which exhibit a bulk
1487 macroscopic dipole moment (e.g. Luttinger \& Tisza's $Z_1$ lattice).
1488 In this particular case, the multipole neutralization scheme can
1489 interfere with the correct computation of the energies. We note that
1490 the energies for these arrangements are typically much larger than for
1491 crystals with net-zero moments, so this is not expected to be an issue
1492 in most simulations.
1493
1494 In large systems, these new methods can be made to scale approximately
1495 linearly with system size, and detailed comparisons with the Ewald sum
1496 for a wide range of chemical environments follows in the second paper.
1497
1498 \begin{acknowledgments}
1499 JDG acknowledges helpful discussions with Christopher
1500 Fennell. Support for this project was provided by the National
1501 Science Foundation under grant CHE-0848243. Computational time was
1502 provided by the Center for Research Computing (CRC) at the
1503 University of Notre Dame.
1504 \end{acknowledgments}
1505
1506 \newpage
1507 \appendix
1508
1509 \section{Smith's $B_l(r)$ functions for damped-charge distributions}
1510 \label{SmithFunc}
1511 The following summarizes Smith's $B_l(r)$ functions and includes
1512 formulas given in his appendix.\cite{Smith98} The first function
1513 $B_0(r)$ is defined by
1514 %
1515 \begin{equation}
1516 B_0(r)=\frac{\text{erfc}(\alpha r)}{r} = \frac{2}{\sqrt{\pi}r}=
1517 \int_{\alpha r}^{\infty} \text{e}^{-s^2} ds .
1518 \end{equation}
1519 %
1520 The first derivative of this function is
1521 %
1522 \begin{equation}
1523 \frac{dB_0(r)}{dr}=-\frac{1}{r^2}\text{erfc}(\alpha r)
1524 -\frac{2\alpha}{r\sqrt{\pi}}\text{e}^{-{\alpha}^2r^2}
1525 \end{equation}
1526 %
1527 which can be used to define a function $B_1(r)$:
1528 %
1529 \begin{equation}
1530 B_1(r)=-\frac{1}{r}\frac{dB_0(r)}{dr}
1531 \end{equation}
1532 %
1533 In general, the recurrence relation,
1534 \begin{equation}
1535 B_l(r)=-\frac{1}{r}\frac{dB_{l-1}(r)}{dr}
1536 = \frac{1}{r^2} \left[ (2l-1)B_{l-1}(r) + \frac {(2\alpha^2)^l}{\alpha \sqrt{\pi}}
1537 \text{e}^{-{\alpha}^2r^2}
1538 \right] ,
1539 \end{equation}
1540 is very useful for building up higher derivatives. Using these
1541 formulas, we find:
1542 %
1543 \begin{align}
1544 \frac{dB_0}{dr}=&-rB_1(r) \\
1545 \frac{d^2B_0}{dr^2}=& - B_1(r) + r^2 B_2(r) \\
1546 \frac{d^3B_0}{dr^3}=& 3 r B_2(r) - r^3 B_3(r) \\
1547 \frac{d^4B_0}{dr^4}=& 3 B_2(r) - 6 r^2 B_3(r) + r^4 B_4(r) \\
1548 \frac{d^5B_0}{dr^5}=& - 15 r B_3(r) + 10 r^3 B_4(r) - r^5 B_5(r) .
1549 \end{align}
1550 %
1551 As noted by Smith, it is possible to approximate the $B_l(r)$
1552 functions,
1553 %
1554 \begin{equation}
1555 B_l(r)=\frac{(2l)!}{l!2^lr^{2l+1}} - \frac {(2\alpha^2)^{l+1}}{(2l+1)\alpha \sqrt{\pi}}
1556 +\text{O}(r) .
1557 \end{equation}
1558 \newpage
1559 \section{The $r$-dependent factors for TSF electrostatics}
1560
1561 Using the shifted damped functions $f_n(r)$ defined by:
1562 %
1563 \begin{equation}
1564 f_n(r)= B_0(r) -\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} B_0^{(m)}(r_c) ,
1565 \end{equation}
1566 %
1567 where the superscript $(m)$ denotes the $m^\mathrm{th}$ derivative. In
1568 this Appendix, we provide formulas for successive derivatives of this
1569 function. (If there is no damping, then $B_0(r)$ is replaced by
1570 $1/r$.) First, we find:
1571 %
1572 \begin{equation}
1573 \frac{\partial f_n}{\partial r_\alpha}=\hat{r}_\alpha \frac{d f_n}{d r} .
1574 \end{equation}
1575 %
1576 This formula clearly brings in derivatives of Smith's $B_0(r)$
1577 function, and we define higher-order derivatives as follows:
1578 %
1579 \begin{align}
1580 g_n(r)=& \frac{d f_n}{d r} =
1581 B_0^{(1)}(r) -\sum_{m=0}^{n} \frac {(r-r_c)^m}{m!} B_0^{(m+1)}(r_c) \\
1582 h_n(r)=& \frac{d^2f_n}{d r^2} =
1583 B_0^{(2)}(r) -\sum_{m=0}^{n-1} \frac {(r-r_c)^m}{m!} B_0^{(m+2)}(r_c) \\
1584 s_n(r)=& \frac{d^3f_n}{d r^3} =
1585 B_0^{(3)}(r) -\sum_{m=0}^{n-2} \frac {(r-r_c)^m}{m!} B_0^{(m+3)}(r_c) \\
1586 t_n(r)=& \frac{d^4f_n}{d r^4} =
1587 B_0^{(4)}(r) -\sum_{m=0}^{n-3} \frac {(r-r_c)^m}{m!} B_0^{(m+4)}(r_c) \\
1588 u_n(r)=& \frac{d^5f_n}{d r^5} =
1589 B_0^{(5)}(r) -\sum_{m=0}^{n-4} \frac {(r-r_c)^m}{m!} B_0^{(m+5)}(r_c) .
1590 \end{align}
1591 %
1592 We note that the last function needed (for quadrupole-quadrupole interactions) is
1593 %
1594 \begin{equation}
1595 u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1596 \end{equation}
1597 % The functions
1598 % needed are listed schematically below:
1599 % %
1600 % \begin{eqnarray}
1601 % f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1602 % g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1603 % h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1604 % s_2 \quad s_3 \quad &s_4 \nonumber \\
1605 % t_3 \quad &t_4 \nonumber \\
1606 % &u_4 \nonumber .
1607 % \end{eqnarray}
1608 The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and
1609 stored on a grid for values of $r$ from $0$ to $r_c$. Using these
1610 functions, we find
1611 %
1612 \begin{align}
1613 \frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\
1614 \frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r}
1615 +r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\
1616 \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =&
1617 \left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta +
1618 \delta_{ \beta \gamma} r_\alpha \right)
1619 \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\
1620 & + r_\alpha r_\beta r_\gamma
1621 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\
1622 \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial
1623 r_\gamma \partial r_\delta} =&
1624 \left( \delta_{\alpha \beta} \delta_{\gamma \delta}
1625 + \delta_{\alpha \gamma} \delta_{\beta \delta}
1626 +\delta_{ \beta \gamma} \delta_{\alpha \delta} \right)
1627 \left( - \frac{g_n}{r^3} + \frac{h_n}{r^2} \right) \nonumber \\
1628 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta
1629 + \text{5 permutations}
1630 \right) \left( \frac{3 g_n}{r^5} - \frac{3h_n}{r^4} + \frac{s_n}{r^3}
1631 \right) \nonumber \\
1632 &+ r_\alpha r_\beta r_\gamma r_\delta
1633 \left( -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5}
1634 + \frac{t_n}{r^4} \right)\\
1635 \frac{\partial^5 f_n}
1636 {\partial r_\alpha \partial r_\beta \partial r_\gamma \partial
1637 r_\delta \partial r_\epsilon} =&
1638 \left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon
1639 + \text{14 permutations} \right)
1640 \left( \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\
1641 &+ \left( \delta_{\alpha \beta} r_\gamma r_\delta r_\epsilon
1642 + \text{9 permutations}
1643 \right) \left(- \frac{15g_n}{r^7}+\frac{15h_n}{r^7} -\frac{6s_n}{r^5} +\frac{t_n}{r^4}
1644 \right) \nonumber \\
1645 &+ r_\alpha r_\beta r_\gamma r_\delta r_\epsilon
1646 \left( \frac{105g_n}{r^9} - \frac{105h_n}{r^8} + \frac{45s_n}{r^7}
1647 - \frac{10t_n}{r^6} +\frac{u_n}{r^5} \right) \label{eq:b13}
1648 \end{align}
1649 %
1650 %
1651 %
1652 \newpage
1653 \section{The $r$-dependent factors for GSF electrostatics}
1654
1655 In Gradient-shifted force electrostatics, the kernel is not expanded,
1656 rather the individual terms in the multipole interaction energies.
1657 For damped charges , this still brings into the algebra multiple
1658 derivatives of the Smith's $B_0(r)$ function. To denote these terms,
1659 we generalize the notation of the previous appendix. For either
1660 $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$ (damped),
1661 %
1662 \begin{align}
1663 g(r)=& \frac{df}{d r}\\
1664 h(r)=& \frac{dg}{d r} = \frac{d^2f}{d r^2} \\
1665 s(r)=& \frac{dh}{d r} = \frac{d^3f}{d r^3} \\
1666 t(r)=& \frac{ds}{d r} = \frac{d^4f}{d r^4} \\
1667 u(r)=& \frac{dt}{d r} = \frac{d^5f}{d r^5} .
1668 \end{align}
1669 %
1670 For undamped charges Table I lists these derivatives under the column
1671 ``Bare Coulomb.'' Equations \ref{eq:b9} to \ref{eq:b13} are still
1672 correct for GSF electrostatics if the subscript $n$ is eliminated.
1673
1674 \newpage
1675
1676 \bibliography{multipole}
1677
1678 \end{document}
1679 %
1680 % ****** End of file multipole.tex ******