ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole1.tex
(Generate patch)

Comparing trunk/multipole/multipole1.tex (file contents):
Revision 3988 by gezelter, Thu Jan 2 15:46:58 2014 UTC vs.
Revision 4183 by gezelter, Sat Jun 14 15:24:30 2014 UTC

# Line 44 | Line 44 | Sum. I. Taylor-shifted and Gradient-shifted electrosta
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}
47 > Sum. I. Shifted electrostatics for multipoles}
48  
49   \author{Madan Lamichhane}
50   \affiliation{Department of Physics, University
# Line 65 | Line 65 | of Notre Dame, Notre Dame, IN 46556}
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
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.  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.
70 >  on truncated Taylor expansions around the cutoff radius.  These
71 >  include a shifted potential (SP) that generalizes the Wolf method
72 >  for point multipoles, and Taylor-shifted force (TSF) and
73 >  gradient-shifted force (GSF) potentials that are both
74 >  generalizations of DSF electrostatics for multipoles. We find that
75 >  each of the distinct orientational contributions requires a separate
76 >  radial function to ensure that pairwise energies, forces and torques
77 >  all vanish at the cutoff radius. In this paper, we present energy,
78 >  force, and torque expressions for the new models, and compare these
79 >  real-space interaction models to exact results for ordered arrays of
80 >  multipoles.  We find that the GSF and SP methods converge rapidly to
81 >  the correct lattice energies for ordered dipolar and quadrupolar
82 >  arrays, while the Taylor-Shifted Force (TSF) is too severe an
83 >  approximation to provide accurate convergence to lattice energies.
84 >  Because real-space methods can be made to scale linearly with system
85 >  size, SP and GSF are attractive options for large Monte
86 >  Carlo and molecular dynamics simulations, respectively.
87   \end{abstract}
88  
89   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
# Line 133 | Line 143 | kernel to point multipoles.  We describe two distinct
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
146 > kernel to point multipoles.  We describe three distinct real-space
147 > interaction models for higher-order multipoles based on truncated
148   Taylor expansions that are carried out at the cutoff radius.  We are
149 < calling these models {\bf Taylor-shifted} and {\bf Gradient-shifted}
149 > calling these models {\bf Taylor-shifted} (TSF), {\bf
150 >  gradient-shifted} (GSF) and {\bf shifted potential} (SP)
151   electrostatics.  Because of differences in the initial assumptions,
152 < the two methods yield related, but somewhat different expressions for
153 < energies, forces, and torques.
152 > the two methods yield related, but distinct expressions for energies,
153 > forces, and torques.
154  
155   In this paper we outline the new methodology and give functional forms
156   for the energies, forces, and torques up to quadrupole-quadrupole
# Line 152 | Line 163 | V = \sum_i \sum_{j>i} V_\mathrm{pair}(r_{ij}, \Omega_i
163   An efficient real-space electrostatic method involves the use of a
164   pair-wise functional form,
165   \begin{equation}
166 < V = \sum_i \sum_{j>i} V_\mathrm{pair}(r_{ij}, \Omega_i, \Omega_j) +
167 < \sum_i V_i^\mathrm{correction}
166 > U = \sum_i \sum_{j>i} U_\mathrm{pair}(\mathbf{r}_{ij}, \Omega_i, \Omega_j) +
167 > \sum_i U_i^\mathrm{self}
168   \end{equation}
169   that is short-ranged and easily truncated at a cutoff radius,
170   \begin{equation}
171 <  V_\mathrm{pair}(r_{ij}, \Omega_i, \Omega_j) = \left\{
171 >  U_\mathrm{pair}(\mathbf{r}_{ij},\Omega_i, \Omega_j) = \left\{
172   \begin{array}{ll}
173 < V_\mathrm{approx} (r_{ij}, \Omega_i, \Omega_j) & \quad r \le r_c \\
174 < 0 & \quad r > r_c ,
173 > U_\mathrm{approx} (\mathbf{r}_{ij}, \Omega_i, \Omega_j) & \quad \left| \mathbf{r}_{ij} \right| \le r_c \\
174 > 0 & \quad \left| \mathbf{r}_{ij} \right|  > r_c ,
175   \end{array}
176   \right.
177   \end{equation}
178 < along with an easily computed correction term ($\sum_i
179 < V_i^\mathrm{correction}$) which has linear-scaling with the number of
178 > along with an easily computed self-interaction term ($\sum_i
179 > U_i^\mathrm{self}$) which scales linearly with the number of
180   particles.  Here $\Omega_i$ and $\Omega_j$ represent orientational
181 < coordinates of the two sites.  The computational efficiency, energy
181 > coordinates of the two sites, and $\mathbf{r}_{ij}$ is the vector
182 > between the two sites.  The computational efficiency, energy
183   conservation, and even some physical properties of a simulation can
184 < depend dramatically on how the $V_\mathrm{approx}$ function behaves at
184 > depend dramatically on how the $U_\mathrm{approx}$ function behaves at
185   the cutoff radius. The goal of any approximation method should be to
186   mimic the real behavior of the electrostatic interactions as closely
187   as possible without sacrificing the near-linear scaling of a cutoff
# Line 180 | Line 192 | computed within $r_c$. Damping using a complementary e
192   contained within the cutoff sphere surrounding each particle.  This is
193   accomplished by shifting the potential functions to generate image
194   charges on the surface of the cutoff sphere for each pair interaction
195 < computed within $r_c$. Damping using a complementary error
196 < function is applied to the potential to accelerate convergence. The
197 < potential for the DSF method, where $\alpha$ is the adjustable damping
186 < parameter, is given by
195 > computed within $r_c$. Damping using a complementary error function is
196 > applied to the potential to accelerate convergence. The interaction
197 > for a pair of charges ($C_i$ and $C_j$) in the DSF method,
198   \begin{equation*}
199 < V_\mathrm{DSF}(r) = C_i C_j \Biggr{[}
199 > U_\mathrm{DSF}(r_{ij}) = C_i C_j \Biggr{[}
200   \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
201   - \frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c} + \left(\frac{\mathrm{erfc}\left(\alpha r_c\right)}{r_c^2}
202   + \frac{2\alpha}{\pi^{1/2}}
# Line 193 | Line 204 | Note that in this potential and in all electrostatic q
204   \right)\left(r_{ij}-r_c\right)\ \Biggr{]}
205   \label{eq:DSFPot}
206   \end{equation*}
207 < Note that in this potential and in all electrostatic quantities that
208 < follow, the standard $1/4 \pi \epsilon_{0}$ has been omitted for
209 < clarity.
207 > where $\alpha$ is the adjustable damping parameter.  Note that in this
208 > potential and in all electrostatic quantities that follow, the
209 > standard $1/4 \pi \epsilon_{0}$ has been omitted for clarity.
210  
211   To insure net charge neutrality within each cutoff sphere, an
212   additional ``self'' term is added to the potential. This term is
# Line 244 | Line 255 | V_a(\mathbf r) =
255   a$.  Then the electrostatic potential of object $\bf a$ at
256   $\mathbf{r}$ is given by
257   \begin{equation}
258 < V_a(\mathbf r) =
258 > \phi_a(\mathbf r) =
259   \sum_{k \, \text{in \bf a}} \frac{q_k}{\lvert \mathbf{r} - \mathbf{r}_k \rvert}.
260   \end{equation}
261   The Taylor expansion in $r$ can be written using an implied summation
262   notation.  Here Greek indices are used to indicate space coordinates
263   ($x$, $y$, $z$) and the subscripts $k$ and $j$ are reserved for
264 < labelling specific charges in $\bf a$ and $\bf b$ respectively.  The
264 > labeling specific charges in $\bf a$ and $\bf b$ respectively.  The
265   Taylor expansion,
266   \begin{equation}
267   \frac{1}{\lvert \mathbf{r} - \mathbf{r}_k \rvert} =
# Line 263 | Line 274 | V_{\bf a}(\mathbf{r}) =\hat{M}_{\bf a} \frac{1}{r}
274   can then be used to express the electrostatic potential on $\bf a$ in
275   terms of multipole operators,
276   \begin{equation}
277 < V_{\bf a}(\mathbf{r}) =\hat{M}_{\bf a} \frac{1}{r}
277 > \phi_{\bf a}(\mathbf{r}) =\hat{M}_{\bf a} \frac{1}{r}
278   \end{equation}
279   where
280   \begin{equation}
# Line 276 | Line 287 | C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \\
287      a}\alpha\beta}$, respectively.  These are the primitive multipoles
288   which can be expressed as a distribution of charges,
289   \begin{align}
290 < C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \\
291 < D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha} ,\\
292 < Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k r_{k\alpha}  r_{k\beta} .
290 > C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \label{eq:charge} \\
291 > D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha}, \label{eq:dipole}\\
292 > Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k
293 > r_{k\alpha}  r_{k\beta} . \label{eq:quadrupole}
294   \end{align}
295   Note that the definition of the primitive quadrupole here differs from
296   the standard traceless form, and contains an additional Taylor-series
297 < based factor of $1/2$.
297 > based factor of $1/2$.  We are essentially treating the mass
298 > distribution with higher priority; the moment of inertia tensor,
299 > $\overleftrightarrow{\mathsf I}$, is diagonalized to obtain body-fixed
300 > axes, and the charge distribution may result in a quadrupole tensor
301 > that is not necessarily diagonal in the body frame.  Additional
302 > reasons for utilizing the primitive quadrupole are discussed in
303 > section \ref{sec:damped}.
304  
305   It is convenient to locate charges $q_j$ relative to the center of mass of  $\bf b$.  Then with $\bf{r}$ pointing from
306 < $\bf a$ to $\bf b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_b $), the interaction energy is given by
306 > $\bf a$ to $\bf b$ ($\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $), the interaction energy is given by
307   \begin{equation}
308   U_{\bf{ab}}(r)
309   = \hat{M}_a \sum_{j \, \text{in \bf b}} \frac {q_j}{\vert \bf{r}+\bf{r}_j \vert} .
# Line 303 | Line 321 | of $\bf a$ interacting with the same multipoles on $\b
321   \end{equation}
322   This form has the benefit of separating out the energies of
323   interaction into contributions from the charge, dipole, and quadrupole
324 < of $\bf a$ interacting with the same multipoles on $\bf b$.
324 > of $\bf a$ interacting with the same multipoles in $\bf b$.
325  
326   \subsection{Damped Coulomb interactions}
327 + \label{sec:damped}
328   In the standard multipole expansion, one typically uses the bare
329   Coulomb potential, with radial dependence $1/r$, as shown in
330   Eq.~(\ref{kernel}).  It is also quite common to use a damped Coulomb
# Line 321 | Line 340 | functions $B_l(r)$ are summarized in Appendix A.
340   either $1/r$ or $B_0(r)$, and all of the techniques can be applied to
341   bare or damped Coulomb kernels (or any other function) as long as
342   derivatives of these functions are known.  Smith's convenient
343 < functions $B_l(r)$ are summarized in Appendix A.
343 > functions $B_l(r)$, which are used for derivatives of the damped
344 > kernel, are summarized in Appendix A.  (N.B. there is one important
345 > distinction between the two kernels, which is the behavior of
346 > $\nabla^2 \frac{1}{r}$ compared with $\nabla^2 B_0(r)$.  The former is
347 > zero everywhere except for a delta function evaluated at the origin.
348 > The latter also has delta function behavior, but is non-zero for $r
349 > \neq 0$.  Thus the standard justification for using a traceless
350 > quadrupole tensor fails for the damped case.)
351  
352   The main goal of this work is to smoothly cut off the interaction
353   energy as well as forces and torques as $r\rightarrow r_c$.  To
# Line 418 | Line 444 | U= (\text{prefactor}) (\text{derivatives}) f_n(r)
444   In general, we can write
445   %
446   \begin{equation}
447 < U= (\text{prefactor}) (\text{derivatives}) f_n(r)
447 > U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
448   \label{generic}
449   \end{equation}
450   %
# Line 429 | Line 455 | space indices.
455   $Q_{{\bf a}\alpha\beta}Q_{{\bf b}\gamma\delta}$, the derivatives are
456   $\partial^4/\partial r_\alpha \partial r_\beta \partial
457   r_\gamma \partial r_\delta$, with implied summation combining the
458 < space indices.
458 > space indices.  Appendix \ref{radialTSF} contains details on the
459 > radial functions.
460  
461   In the formulas presented in the tables below, the placeholder
462   function $f(r)$ is used to represent the electrostatic kernel (either
# Line 439 | Line 466 | presented in tables \ref{tab:tableenergy} and \ref{tab
466   of the same index $n$.  The algebra required to evaluate energies,
467   forces and torques is somewhat tedious, so only the final forms are
468   presented in tables \ref{tab:tableenergy} and \ref{tab:tableFORCE}.
469 + One of the principal findings of our work is that the individual
470 + orientational contributions to the various multipole-multipole
471 + interactions must be treated with distinct radial functions, but each
472 + of these contributions is independently force shifted at the cutoff
473 + radius.  
474  
475   \subsection{Gradient-shifted force (GSF) electrostatics}
476   The second, and conceptually simpler approach to force-shifting
# Line 446 | Line 478 | U^{\text{shift}}(r)=U(r)-U(r_c)-(r-r_c)\hat{r}\cdot \n
478   expansion, and has a similar interaction energy for all multipole
479   orders:
480   \begin{equation}
481 < U^{\text{shift}}(r)=U(r)-U(r_c)-(r-r_c)\hat{r}\cdot \nabla U(r) \Big
482 < \lvert  _{r_c} .
481 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
482 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c)
483 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
484   \label{generic2}
485   \end{equation}
486 < Here the gradient for force shifting is evaluated for an image
487 < multipole projected onto the surface of the cutoff sphere (see fig
488 < \ref{fig:shiftedMultipoles}). No higher order terms $(r-r_c)^n$
489 < appear.  The primary difference between the TSF and GSF methods is the
490 < stage at which the Taylor Series is applied; in the Taylor-shifted
491 < approach, it is applied to the kernel itself.  In the Gradient-shifted
492 < approach, it is applied to individual radial interactions terms in the
493 < multipole expansion.  Energies from this method thus have the general
494 < form:
486 > where $\hat{\mathbf{r}}$ is the unit vector pointing between the two
487 > multipoles, and the sum describes a separate force-shifting that is
488 > applied to each orientational contribution to the energy.  Both the
489 > potential and the gradient for force shifting are evaluated for an
490 > image multipole projected onto the surface of the cutoff sphere (see
491 > fig \ref{fig:shiftedMultipoles}).  The image multipole retains the
492 > orientation ($\hat{\mathbf{b}}$) of the interacting multipole.  No
493 > higher order terms $(r-r_c)^n$ appear.  The primary difference between
494 > the TSF and GSF methods is the stage at which the Taylor Series is
495 > applied; in the Taylor-shifted approach, it is applied to the kernel
496 > itself.  In the Gradient-shifted approach, it is applied to individual
497 > radial interaction terms in the multipole expansion.  Energies from
498 > this method thus have the general form:
499   \begin{equation}
500   U= \sum  (\text{angular factor}) (\text{radial factor}).
501   \label{generic3}
502   \end{equation}
503  
504   Functional forms for both methods (TSF and GSF) can both be summarized
505 < using the form of Eq.~(\ref{generic3}).  The basic forms for the
505 > using the form of Eq.~\ref{generic3}).  The basic forms for the
506   energy, force, and torque expressions are tabulated for both shifting
507   approaches below -- for each separate orientational contribution, only
508   the radial factors differ between the two methods.
509  
510 < \subsection{\label{sec:level2}Body and space axes}
510 > \subsection{Generalization of the Wolf shifted potential (SP)}
511 > It is also possible to formulate an extension of the Wolf approach for
512 > multipoles by simply projecting the image multipole onto the surface
513 > of the cutoff sphere, and including the interactions with the central
514 > multipole and the image.  This effectively shifts the pair potential
515 > to zero at the cutoff radius,
516 > \begin{equation}
517 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
518 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
519 > \label{eq:SP}
520 > \end{equation}
521 > independent of the orientations of the two multipoles.  The sum again
522 > describes separate potential shifting that is applied to each
523 > orientational contribution to the energy.
524 >
525 > The shifted potential (SP) method is a simple truncation of the GSF
526 > method for each orientational contribution, leaving out the $(r-r_c)$
527 > terms that multiply the gradient. Functional forms for the
528 > shifted-potential (SP) method can also be summarized using the form of
529 > Eq.~\ref{generic3}.  The energy, force, and torque expressions are
530 > tabulated below for all three methods. As in the GSF and TSF methods,
531 > for each separate orientational contribution, only the radial factors
532 > differ between the SP, GSF, and TSF methods.
533  
475 [XXX Do we need this section in the main paper? or should it go in the
476 extra materials?]
534  
535 < So far, all energies and forces have been written in terms of fixed
536 < space coordinates.  Interaction energies are computed from the generic
537 < formulas Eq.~(\ref{generic}) and ~(\ref{generic2}) which combine
538 < orientational prefactors with radial functions.  Because objects $\bf
539 < a$ and $\bf b$ both translate and rotate during a molecular dynamics
540 < (MD) simulation, it is desirable to contract all $r$-dependent terms
541 < with dipole and quadrupole moments expressed in terms of their body
542 < axes.  To do so, we have followed the methodology of Allen and
543 < Germano,\cite{Allen:2006fk} which was itself based on earlier work by
544 < Price {\em et al.}\cite{Price:1984fk}
535 > \subsection{\label{sec:level2}Body and space axes}
536 > Although objects $\bf a$ and $\bf b$ rotate during a molecular
537 > dynamics (MD) simulation, their multipole tensors remain fixed in
538 > body-frame coordinates. While deriving force and torque expressions,
539 > it is therefore convenient to write the energies, forces, and torques
540 > in intermediate forms involving the vectors of the rotation matrices.
541 > We denote body axes for objects $\bf a$ and $\bf b$ using unit vectors
542 > $\hat{a}_m$ and $\hat{b}_m$, respectively, with the index $m=(123)$.
543 > In a typical simulation, the initial axes are obtained by
544 > diagonalizing the moment of inertia tensors for the objects.  (N.B.,
545 > the body axes are generally {\it not} the same as those for which the
546 > quadrupole moment is diagonal.)  The rotation matrices are then
547 > propagated during the simulation.
548  
549 < We denote body axes for objects $\bf a$ and $\bf b$ by unit vectors
490 < $\hat{a}_m$ and $\hat{b}_m$, respectively, with the index $m=(123)$
491 < referring to a convenient set of inertial body axes.  (N.B., these
492 < body axes are generally not the same as those for which the quadrupole
493 < moment is diagonal.)  Then,
494 < %
495 < \begin{eqnarray}
496 < \hat{a}_m= a_{mx}\hat{x} + a_{my}\hat{y} + a_{mz}\hat{z}  \\
497 < \hat{b}_m= b_{mx}\hat{x} + b_{my}\hat{y} + b_{mz}\hat{z}  .
498 < \end{eqnarray}
499 < Rotation matrices $\hat{\mathbf {a}}$ and $\hat{\mathbf {b}}$ can be
549 > The rotation matrices $\hat{\mathbf {a}}$ and $\hat{\mathbf {b}}$ can be
550   expressed using these unit vectors:
551   \begin{eqnarray}
552   \hat{\mathbf {a}} =
# Line 504 | Line 554 | expressed using these unit vectors:
554   \hat{a}_1 \\
555   \hat{a}_2 \\
556   \hat{a}_3
557 < \end{pmatrix}
508 < =
509 < \begin{pmatrix}
510 < a_{1x} \quad a_{1y} \quad a_{1z} \\
511 < a_{2x} \quad a_{2y} \quad a_{2z} \\
512 < a_{3x} \quad a_{3y} \quad a_{3z}
513 < \end{pmatrix}\\
557 > \end{pmatrix}, \qquad
558   \hat{\mathbf {b}} =
559   \begin{pmatrix}
560   \hat{b}_1 \\
561   \hat{b}_2 \\
562   \hat{b}_3
563   \end{pmatrix}
520 =
521 \begin{pmatrix}
522 b_{1x} \quad  b_{1y} \quad b_{1z} \\
523 b_{2x} \quad b_{2y} \quad b_{2z} \\
524 b_{3x} \quad b_{3y} \quad b_{3z}
525 \end{pmatrix}  .
564   \end{eqnarray}
565   %
566   These matrices convert from space-fixed $(xyz)$ to body-fixed $(123)$
567 < coordinates.  All contractions of prefactors with derivatives of
568 < functions can be written in terms of these matrices. It proves to be
569 < equally convenient to just write any contraction in terms of unit
570 < vectors $\hat{r}$, $\hat{a}_m$, and $\hat{b}_n$. In the torque
571 < expressions, it is useful to have the angular-dependent terms
572 < available in three different fashions, e.g. for the dipole-dipole
573 < contraction:
567 > coordinates.
568 >
569 > Allen and Germano,\cite{Allen:2006fk} following earlier work by Price
570 > {\em et al.},\cite{Price:1984fk} showed that if the interaction
571 > energies are written explicitly in terms of $\hat{r}$ and the body
572 > axes ($\hat{a}_m$, $\hat{b}_n$) :
573 > %
574 > \begin{equation}
575 > U(r, \{\hat{a}_m \cdot \hat{r} \},
576 > \{\hat{b}_n\cdot \hat{r} \},
577 > \{\hat{a}_m \cdot \hat{b}_n \}) .
578 > \label{ugeneral}
579 > \end{equation}
580 > %
581 > the forces come out relatively cleanly,
582 > %
583 > \begin{equation}
584 > \mathbf{F}_{\bf a}=-\mathbf{F}_{\bf b} =  \frac{\partial U}{\partial \mathbf{r}}
585 > = \frac{\partial U}{\partial r} \hat{r}
586 > + \sum_m \left[
587 > \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
588 > \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
589 > + \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
590 > \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
591 > \right] \label{forceequation}.
592 > \end{equation}
593 >
594 > The torques can also be found in a relatively similar
595 > manner,
596 > %
597 > \begin{eqnarray}
598 > \mathbf{\tau}_{\bf a} =
599 > \sum_m
600 > \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
601 > ( \hat{r} \times \hat{a}_m )
602 > -\sum_{mn}
603 > \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
604 > (\hat{a}_m \times \hat{b}_n) \\
605 > %
606 > \mathbf{\tau}_{\bf b} =
607 > \sum_m
608 > \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
609 > ( \hat{r} \times \hat{b}_m)
610 > +\sum_{mn}
611 > \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
612 > (\hat{a}_m \times \hat{b}_n) .
613 > \end{eqnarray}
614 >
615 > Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_a $
616 > is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk}
617 > We also made use of the identities,
618 > %
619 > \begin{align}
620 > \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
621 > =& \frac{1}{r} \left(  \hat{a}_m - (\hat{a}_m \cdot \hat{r})\hat{r}
622 > \right) \\
623 > \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
624 > =& \frac{1}{r} \left(  \hat{b}_m - (\hat{b}_m \cdot \hat{r})\hat{r}
625 > \right).
626 > \end{align}
627 >
628 > Many of the multipole contractions required can be written in one of
629 > three equivalent forms using the unit vectors $\hat{r}$, $\hat{a}_m$,
630 > and $\hat{b}_n$. In the torque expressions, it is useful to have the
631 > angular-dependent terms available in all three fashions, e.g. for the
632 > dipole-dipole contraction:
633   %
634   \begin{equation}
635   \mathbf{D}_{\mathbf {a}} \cdot \mathbf{D}_{\mathbf{b}}
636   = D_{\bf {a}\alpha} D_{\bf {b}\alpha} =
637 < \sum_{mn} {D_{\mathbf{a}m} \hat{a}_m \cdot \hat{b}_n D_{\mathbf{b}n}}
637 > \sum_{mn} {D_{\mathbf{a}m} \hat{a}_m \cdot \hat{b}_n D_{\mathbf{b}n}}.
638   \end{equation}
639   %
640   The first two forms are written using space coordinates.  The first
# Line 546 | Line 643 | contractions using space indices.
643   explicit sums over body indices and the dot products now indicate
644   contractions using space indices.
645  
646 + In computing our force and torque expressions, we carried out most of
647 + the work in body coordinates, and have transformed the expressions
648 + back to space-frame coordinates, which are reported below.  Interested
649 + readers may consult the supplemental information for this paper for
650 + the intermediate body-frame expressions.
651  
652   \subsection{The Self-Interaction \label{sec:selfTerm}}
653  
654   In addition to cutoff-sphere neutralization, the Wolf
655   summation~\cite{Wolf99} and the damped shifted force (DSF)
656 < extension~\cite{Fennell:2006zl} also included self-interactions that
656 > extension~\cite{Fennell:2006zl} also include self-interactions that
657   are handled separately from the pairwise interactions between
658   sites. The self-term is normally calculated via a single loop over all
659   sites in the system, and is relatively cheap to evaluate. The
# Line 563 | Line 665 | V_\textrm{self} = - \frac{1}{r_c} \sum_{{\bf a}=1}^N C
665   the cutoff sphere.  For a system of undamped charges, the total
666   self-term is
667   \begin{equation}
668 < V_\textrm{self} = - \frac{1}{r_c} \sum_{{\bf a}=1}^N C_{\bf a}^{2}
668 > U_\textrm{self} = - \frac{1}{r_c} \sum_{{\bf a}=1}^N C_{\bf a}^{2}.
669   \label{eq:selfTerm}
670   \end{equation}
671  
# Line 578 | Line 680 | V_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r
680   complexity to the Ewald sum).  For a system containing only damped
681   charges, the complete self-interaction can be written as
682   \begin{equation}
683 < V_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} +
683 > U_\textrm{self} = - \left(\frac{\textrm{erfc}(\alpha r_c)}{r_c} +
684    \frac{\alpha}{\sqrt{\pi}} \right) \sum_{{\bf a}=1}^N
685        C_{\bf a}^{2}.
686   \label{eq:dampSelfTerm}
687   \end{equation}
688  
689   The extension of DSF electrostatics to point multipoles requires
690 < treatment of {\it both} the self-neutralization and reciprocal
690 > treatment of the self-neutralization \textit{and} reciprocal
691   contributions to the self-interaction for higher order multipoles.  In
692   this section we give formulae for these interactions up to quadrupolar
693   order.
# Line 596 | Line 698 | obtained by Smith and Aguado and Madden for the applic
698   cutoff sphere, and averaging over the surface of the cutoff sphere.
699   Because the self term is carried out as a single sum over sites, the
700   reciprocal-space portion is identical to half of the self-term
701 < obtained by Smith and Aguado and Madden for the application of the
702 < Ewald sum to multipoles.\cite{Smith82,Smith98,Aguado03} For a given
703 < site which posesses a charge, dipole, and multipole, both types of
704 < contribution are given in table \ref{tab:tableSelf}.
701 > obtained by Smith, and also by Aguado and Madden for the application
702 > of the Ewald sum to multipoles.\cite{Smith82,Smith98,Aguado03} For a
703 > given site which posesses a charge, dipole, and quadrupole, both types
704 > of contribution are given in table \ref{tab:tableSelf}.
705  
706   \begin{table*}
707    \caption{\label{tab:tableSelf} Self-interaction contributions for
# Line 611 | Line 713 | Quadrupole & $2 \text{Tr}(\mathbf{Q}_{\bf a}^2) + \tex
713   Charge & $C_{\bf a}^2$ & $-f(r_c)$ & $-\frac{\alpha}{\sqrt{\pi}}$ \\
714   Dipole & $|\mathbf{D}_{\bf a}|^2$ & $\frac{1}{3} \left( h(r_c) +
715    \frac{2 g(r_c)}{r_c} \right)$ & $-\frac{2 \alpha^3}{3 \sqrt{\pi}}$\\
716 < Quadrupole & $2 \text{Tr}(\mathbf{Q}_{\bf a}^2) + \text{Tr}(\mathbf{Q}_{\bf a})^2$ &
716 > Quadrupole & $2 \mathbf{Q}_{\bf a}:\mathbf{Q}_{\bf a} + \text{Tr}(\mathbf{Q}_{\bf a})^2$ &
717   $- \frac{1}{15} \left( t(r_c)+ \frac{4 s(r_c)}{r_c} \right)$ &
718   $-\frac{4 \alpha^5}{5 \sqrt{\pi}}$ \\
719   Charge-Quadrupole & $-2 C_{\bf a} \text{Tr}(\mathbf{Q}_{\bf a})$ & $\frac{1}{3} \left(
# Line 637 | Line 739 | work for both the Taylor-shifted and Gradient-shifted
739   \section{Interaction energies, forces, and torques}
740   The main result of this paper is a set of expressions for the
741   energies, forces and torques (up to quadrupole-quadrupole order) that
742 < work for both the Taylor-shifted and Gradient-shifted approximations.
743 < These expressions were derived using a set of generic radial
744 < functions.  Without using the shifting approximations mentioned above,
745 < some of these radial functions would be identical, and the expressions
746 < coalesce into the familiar forms for unmodified multipole-multipole
747 < interactions.  Table \ref{tab:tableenergy} maps between the generic
748 < functions and the radial functions derived for both the Taylor-shifted
749 < and Gradient-shifted methods.  The energy equations are written in
750 < terms of lab-frame representations of the dipoles, quadrupoles, and
751 < the unit vector connecting the two objects,
742 > work for the Taylor-shifted, gradient-shifted, and shifted potential
743 > approximations.  These expressions were derived using a set of generic
744 > radial functions.  Without using the shifting approximations mentioned
745 > above, some of these radial functions would be identical, and the
746 > expressions coalesce into the familiar forms for unmodified
747 > multipole-multipole interactions.  Table \ref{tab:tableenergy} maps
748 > between the generic functions and the radial functions derived for the
749 > three methods.  The energy equations are written in terms of lab-frame
750 > representations of the dipoles, quadrupoles, and the unit vector
751 > connecting the two objects,
752  
753   % Energy in space coordinate form ----------------------------------------------------------------------------------------------
754   %
# Line 741 | Line 843 | +2 \text{Tr} \left(
843   U_{Q_{\bf a}Q_{\bf b}}(r)=&
844   \Bigl[
845   \text{Tr} \mathbf{Q}_{\mathbf{a}} \text{Tr} \mathbf{Q}_{\mathbf{b}}
846 < +2 \text{Tr} \left(
847 < \mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} \right) \Bigr] v_{41}(r)
846 > +2
847 > \mathbf{Q}_{\mathbf{a}} : \mathbf{Q}_{\mathbf{b}} \Bigr] v_{41}(r)
848   \\
849   % 2
850   &+\Bigl[ \text{Tr}\mathbf{Q}_{\mathbf{a}}
# Line 773 | Line 875 | along with the sign of the intersite vector, $\hat{r}$
875  
876   \begin{sidewaystable}
877    \caption{\label{tab:tableenergy}Radial functions used in the energy
878 <    and torque equations.  The $f, g, h, s, t, \mathrm{and} u$
879 <    functions used in this table are defined in Appendices B and C.}
880 < \begin{tabular}{|c|c|l|l|} \hline
881 < Generic&Bare Coulomb&Taylor-Shifted&Gradient-Shifted
878 >    and torque equations.  The $f, g, h, s, t, \mathrm{and~} u$
879 >    functions used in this table are defined in Appendices
880 >    \ref{radialTSF} and \ref{radialGSF}.  The gradient shifted (GSF)
881 >    functions include the shifted potential (SP)
882 >    contributions (\textit{cf.} Eqs. \ref{generic2} and
883 >    \ref{eq:SP}).}
884 > \begin{tabular}{|c|c|l|l|l|} \hline
885 > Generic&Bare Coulomb&Taylor-Shifted (TSF)&Shifted Potential (SP)&Gradient-Shifted (GSF)
886   \\ \hline
887   %
888   %
# Line 785 | Line 891 | $f(r)-f(r_c)-(r-r_c)g(r_c)$
891   $v_{01}(r)$ &
892   $\frac{1}{r}$ &
893   $f_0(r)$ &
894 < $f(r)-f(r_c)-(r-r_c)g(r_c)$
894 > $f(r)-f(r_c)$ &
895 > SP $-(r-r_c)g(r_c)$
896   \\
897   %
898   %
# Line 794 | Line 901 | $g(r)-g(r_c)-(r-r_c)h(r_c)$ \\
901   $v_{11}(r)$ &
902   $-\frac{1}{r^2}$ &
903   $g_1(r)$ &
904 < $g(r)-g(r_c)-(r-r_c)h(r_c)$ \\
904 > $g(r)-g(r_c)$ &
905 > SP $-(r-r_c)h(r_c)$ \\
906   %
907   %
908   %
# Line 802 | Line 910 | $\frac{g(r)}{r}-\frac{g(r_c)}{r_c} -(r-r_c)
910   $v_{21}(r)$ &
911   $-\frac{1}{r^3}  $ &
912   $\frac{g_2(r)}{r} $ &
913 < $\frac{g(r)}{r}-\frac{g(r_c)}{r_c} -(r-r_c)
914 < \left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right)$ \\
913 > $\frac{g(r)}{r}-\frac{g(r_c)}{r_c}$ &
914 > SP $-(r-r_c) \left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right)$ \\
915 > %
916 > %
917 > %
918   $v_{22}(r)$ &
919   $\frac{3}{r^3}  $ &
920   $\left(-\frac{g_2(r)}{r} + h_2(r) \right)$ &
921 < $\left(-\frac{g(r)}{r}+h(r) \right)
922 < -\left(-\frac{g(r_c)}{r_c}+h(r_c) \right)$  \\
812 < &&& $ ~~~-(r-r_c) \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$
813 < \\
921 > $\left(-\frac{g(r)}{r}+h(r) \right) -\left(-\frac{g(r_c)}{r_c}+h(r_c) \right)$
922 > & SP $-(r-r_c) \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\
923   %
924   %
925   %
# Line 818 | Line 927 | $\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r} \right)
927   $v_{31}(r)$ &
928   $\frac{3}{r^4}  $ &
929   $\left(-\frac{g_3(r)}{r^2} + \frac{h_3(r)}{r} \right)$ &
930 < $\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r} \right)
931 < -\left(-\frac{g(r_c)}{r_c^2}+\frac{h(r_c)}{r_c} \right) $\\
823 < &&&$ ~~~ -(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)$
824 < \\
930 > $\left( -\frac{g(r)}{r^2}+\frac{h(r)}{r}\right)-\left(-\frac{g(r_c)}{r_c^2}+\frac{h(r_c)}{r_c} \right)$
931 > & SP $-(r-r_c) \left(\frac{2g(r_c)}{r_c^3}-\frac{2h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$  \\
932   %
933 + %
934 + %
935   $v_{32}(r)$ &
936   $-\frac{15}{r^4}  $ &
937   $\left( \frac{3g_3(r)}{r^2} - \frac{3h_3(r)}{r} + s_3(r) \right)$ &
938 < $\left( \frac{3g(r)}{r^2} - \frac{3h(r)}{r} + s(r) \right)
939 < - \left( \frac{3g(r_c)}{r_c^2} - \frac{3h(r_c)}{r_c} + s(r_c) \right)$ \\
940 < &&&$ ~~~ -(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)$
941 < \\
938 > $\left( \frac{3g(r)}{r^2} - \frac{3h(r)}{r} + s(r) \right)$&
939 > SP $-(r-r_c) \left( \frac{-6g(r_c)}{r_c^3}+\frac{6h(r_c)}{r_c^2}\right.$ \\
940 > &&& $~~~-\left(\frac{3g(r_c)}{r_c^2} - \frac{3h(r_c)}{r_c} + s(r_c)\right)$ &
941 > $\phantom{SP-(r-r_c)}\left.-\frac{3s(r_c)}{r_c}+t(r_c) \right)$\\
942   %
943   %
944   %
# Line 837 | Line 946 | $\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)
946   $v_{41}(r)$ &
947   $\frac{3}{r^5} $ &
948   $\left(-\frac{g_4(r)}{r^3} +\frac{h_4(r)}{r^2} \right) $  &
949 < $\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)
950 < - \left( -\frac{g(r_c)}{r_c^3} + \frac{h(r_c)}{r_c^2} \right)$ \\
842 < &&&$ ~~~ -(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)$
949 > $\left( -\frac{g(r)}{r^3} + \frac{h(r)}{r^2} \right)- \left(-\frac{g(r_c)}{r_c^3} + \frac{h(r_c)}{r_c^2} \right)$ &
950 > SP $-(r-r_c) \left( \frac{3g(r_c)}{r_c^4}-\frac{3h(r_c)}{r_c^3}+\frac{s(r_c)}{r_c^2} \right)$
951   \\
952   % 2
953   $v_{42}(r)$ &
954   $- \frac{15}{r^5}   $ &
955   $\left( \frac{3g_4(r)}{r^3} - \frac{3h_4(r)}{r^2}+\frac{s_4(r)}{r} \right)$ &
956 < $\left( \frac{3g(r)}{r^3} - \frac{3h(r)}{r^2}+\frac{s(r)}{r} \right)
957 < -\left( \frac{3g(r_c)}{r_c^3} - \frac{3h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ \\
958 < &&&$ ~~~ -(r-r_c) \left(- \frac{9g(r_c)}{r_c^4}+\frac{9h(r_c)}{r_c^3}
959 < -\frac{4s(r_c)}{r_c^2} + \frac{t(r_c)}{r_c}\right)$
852 < \\
956 > $\left( \frac{3g(r)}{r^3} - \frac{3h(r)}{r^2}+\frac{s(r)}{r} \right)$ &
957 > SP$-(r-r_c) \left(- \frac{9g(r_c)}{r_c^4}+\frac{9h(r_c)}{r_c^3}\right.$  \\
958 > &&& $~~~-\left( \frac{3g(r_c)}{r_c^3} -  \frac{3h(r_c)}{r_c^2}+\frac{s(r_c)}{r_c} \right)$ &
959 > $\phantom{SP-(r-r_c)}\left. -\frac{4s(r_c)}{r_c^2} + \frac{t(r_c)}{r_c}\right)$\\
960   % 3
961 + %
962 + %
963   $v_{43}(r)$ &
964   $ \frac{105}{r^5}  $ &
965   $\left(-\frac{15g_4(r)}{r^3}+\frac{15h_4(r)}{r^2}-\frac{6s_4(r)}{r} + t_4(r)\right) $ &
966 < $\left(-\frac{15g(r)}{r^3}+\frac{15h(r)}{r^2}-\frac{6s(r)}{r} + t(r)\right)$ \\
967 < &&&$~~~ -\left(-\frac{15g(r_c)}{r_c^3}+\frac{15h(r_c)}{r_c^2}-\frac{6s(r_c)}{r_c} + t(r_c)\right)$ \\
968 < &&&$~~~ -(r-r_c)\left(\frac{45g(r_c)}{r_c^4}-\frac{45h(r_c)}{r_c^3}+\frac{21s(r_c)}{r_c^2}
969 < -\frac{6t(r_c)}{r_c}+u(r_c) \right)$ \\ \hline
966 > $ \left(-\frac{15g(r)}{r^3} +\frac{15h(r)}{r^2}-\frac{6s(r)}{r}+t(r)\right) $ &
967 > SP $-(r-r_c)\left(\frac{45g(r_c)}{r_c^4}-\frac{45h(r_c)}{r_c^3}\right.$\\
968 > &&& $~~~-\left(-\frac{15g(r_c)}{r_c^3}+\frac{15h(r_c)}{r_c^2}-\frac{6s(r_c)}{r_c}+ t(r_c)\right)$ &
969 > $\phantom{SP-(r-r_c)}\left.+\frac{21s(r_c)}{r_c^2}-\frac{6t(r_c)}{r_c}+u(r_c) \right)$\\
970 > \hline
971   \end{tabular}
972   \end{sidewaystable}
973   %
# Line 866 | Line 976 | -\frac{6t(r_c)}{r_c}+u(r_c) \right)$ \\ \hline
976   %
977  
978   \begin{sidewaystable}
979 < \caption{\label{tab:tableFORCE}Radial functions used in the force equations.}
980 < \begin{tabular}{|c|c|l|l|} \hline
981 < Function&Definition&Taylor-Shifted&Gradient-Shifted
979 > \caption{\label{tab:tableFORCE}Radial functions used in the force
980 >  equations. Gradient shifted (GSF) functions are constructed using the shifted
981 >    potential (SP) functions.  Some of these functions are simple
982 >    modifications of the functions found in table \ref{tab:tableenergy}}
983 > \begin{tabular}{|c|c|l|l|l|} \hline
984 > Function&Definition&Taylor-Shifted (TSF)& Shifted Potential (SP)
985 > &Gradient-Shifted (GSF)
986   \\ \hline
987   %
988   %
# Line 876 | Line 990 | $g(r)-g(r_c)$ \\
990   $w_a(r)$&
991   $\frac{d v_{01}}{dr}$&
992   $g_0(r)$&
993 < $g(r)-g(r_c)$ \\
993 > $g(r)$&
994 > SP $-g(r_c)$ \\
995   %
996   %
997   $w_b(r)$ &
998   $\frac{d v_{11}}{dr} - \frac{v_{11}(r)}{r} $&
999   $\left( -\frac{g_1(r)}{r}+h_1(r) \right)$ &
1000 < $h(r)- h(r_c) - \frac{v_{11}(r)}{r} $ \\
1000 > $h(r) - \frac{v_{11}(r)}{r} $ &
1001 > SP $- h(r_c)$ \\
1002   %
1003   $w_c(r)$ &
1004   $\frac{v_{11}(r)}{r}$ &
1005   $\frac{g_1(r)}{r} $ &
1006 + $\frac{v_{11}(r)}{r}$&
1007   $\frac{v_{11}(r)}{r}$\\
1008   %
1009   %
1010   $w_d(r)$&
1011   $\frac{d v_{21}}{dr}$&
1012   $\left( -\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right) $ &
1013 < $\left( -\frac{g(r)}{r^2} + \frac{h(r)}{r} \right)
1014 < -\left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right) $ \\
1013 > $\left( -\frac{g(r)}{r^2} + \frac{h(r)}{r} \right)$ &
1014 > SP $-\left( -\frac{g(r_c)}{r_c^2} + \frac{h(r_c)}{r_c} \right) $ \\
1015   %
1016   $w_e(r)$ &
1017   $\left(-\frac{g_2(r)}{r^2} + \frac{h_2(r)}{r} \right)$ &
1018   $\frac{v_{22}(r)}{r}$ &
1019 + $\frac{v_{22}(r)}{r}$ &
1020   $\frac{v_{22}(r)}{r}$ \\
1021   %
1022   %
1023   $w_f(r)$&
1024   $\frac{d v_{22}}{dr} - \frac{2v_{22}(r)}{r}$&
1025   $\left( \frac{3g_2(r)}{r^2}-\frac{3h_2(r)}{r}+s_2(r) \right)$ &
1026 <  $ \left( \frac{g(r)}{r^2}-\frac{h(r)}{r}+s(r) \right) $  \\
1027 <  &&& $ ~~~- \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c)
910 <     \right)-\frac{2v_{22}(r)}{r}$\\
1026 >  $ \left( \frac{g(r)}{r^2}-\frac{h(r)}{r}+s(r) \right) -\frac{2v_{22}(r)}{r}$&
1027 > SP $- \left( \frac{g(r_c)}{r_c^2}-\frac{h(r_c)}{r_c}+s(r_c) \right)$\\
1028   %
1029   $w_g(r)$&
1030   $\frac{v_{31}(r)}{r}$&
1031   $ \left( -\frac{g_3(r)}{r^3}+\frac{h_3(r)}{r^2} \right)$&
1032 + $\frac{v_{31}(r)}{r}$&
1033   $\frac{v_{31}(r)}{r}$\\
1034   %
1035   $w_h(r)$ &
1036   $\frac{d v_{31}}{dr} -\frac{v_{31}(r)}{r}$&
1037   $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $  &
1038 < $ \left(\frac{2g(r)}{r^3} -\frac{2h(r)}{r^2} +\frac{s(r)}{r} \right) - \left(\frac{2g(r_c)}{r_c^3} -\frac{2h(r_c)}{r_c^2} +\frac{s(r_c)}{r_c} \right) $ \\
1039 < &&& $ ~~~ -\frac{v_{31}(r)}{r}$ \\
1038 > $ \left(\frac{2g(r)}{r^3} -\frac{2h(r)}{r^2} +\frac{s(r)}{r} \right) -\frac{v_{31}(r)}{r}$ &
1039 > SP $ - \left(\frac{2g(r_c)}{r_c^3} -\frac{2h(r_c)}{r_c^2} +\frac{s(r_c)}{r_c} \right) $ \\
1040   % 2
1041   $w_i(r)$ &
1042   $\frac{v_{32}(r)}{r}$ &
1043   $\left(\frac{3g_3(r)}{r^3} -\frac{3h_3(r)}{r^2} +\frac{s_3(r)}{r} \right) $  &
1044 + $\frac{v_{32}(r)}{r}$&
1045   $\frac{v_{32}(r)}{r}$\\
1046   %
1047   $w_j(r)$ &
1048   $\frac{d v_{32}}{dr}  - \frac{3v_{32}}{r}$&
1049   $\left(\frac{-15g_3(r)}{r^3} + \frac{15h_3(r)}{r^2} - \frac{6s_3(r)}{r} + t_3(r) \right)  $ &
1050 < $\left(\frac{-6g(r)}{r^3} +\frac{6h(r)}{r^2} -\frac{3s(r)}{r} +t(r) \right)$ \\
1051 < &&& $~~~-\left(\frac{-6g(_cr)}{r_c^3} +\frac{6h(r_c)}{r_c^2}
1052 <  -\frac{3s(r_c)}{r_c} +t(r_c) \right) -\frac{3v_{32}}{r}$ \\
1050 > $\left(\frac{-6g(r)}{r^3} +\frac{6h(r)}{r^2} -\frac{3s(r)}{r} +t(r) \right) -\frac{3v_{32}}{r}$ &
1051 > SP $-\left(\frac{-6g(_cr)}{r_c^3} +\frac{6h(r_c)}{r_c^2}
1052 >  -\frac{3s(r_c)}{r_c} +t(r_c) \right)$ \\
1053   %
1054   $w_k(r)$ &
1055   $\frac{d v_{41}}{dr} $ &
1056   $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2}  \right)$ &
1057 < $\left(\frac{3g(r)}{r^4} -\frac{3h(r)}{r^3} +\frac{s(r)}{r^2}  \right)
1058 < -\left(\frac{3g(r_c)}{r_c^4} -\frac{3h(r_c)}{r_c^3} +\frac{s(r_c)}{r_c^2}  \right)$ \\
1057 > $\left(\frac{3g(r)}{r^4} -\frac{3h(r)}{r^3} +\frac{s(r)}{r^2}
1058 > \right)$ &
1059 > SP $-\left(\frac{3g(r_c)}{r_c^4} -\frac{3h(r_c)}{r_c^3} +\frac{s(r_c)}{r_c^2}  \right)$ \\
1060   %
1061   $w_l(r)$ &
1062   $\frac{d v_{42}}{dr} -\frac{2v_{42}(r)}{r}$ &
1063   $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1064 < $\left(-\frac{9g(r)}{r^4} +\frac{9h(r)}{r^3} -\frac{4s(r)}{r^2} +\frac{t(r)}{r} \right)$ \\
1065 < &&& $~~~ -\left(-\frac{9g(r_c)}{r_c^4} +\frac{9h(r_c)}{r_c^3} -\frac{4s(r_c)}{r_c^2} +\frac{t(r_c)}{r_c} \right)
1066 < -\frac{2v_{42}(r)}{r}$\\
1064 > $\left(-\frac{9g(r)}{r^4} +\frac{9h(r)}{r^3} -\frac{4s(r)}{r^2}
1065 >  +\frac{t(r)}{r} \right) -\frac{2v_{42}(r)}{r}$&
1066 > SP$-\left(-\frac{9g(r_c)}{r_c^4} +\frac{9h(r_c)}{r_c^3} -\frac{4s(r_c)}{r_c^2} +\frac{t(r_c)}{r_c} \right)$\\
1067   %
1068   $w_m(r)$ &
1069   $\frac{d v_{43}}{dr} -\frac{4v_{43}(r)}{r}$&
1070 < $\left(\frac{105g_4(r)}{r^4} - \frac{105h_4(r)}{r^3} + \frac{45s_4(r)}{r^2} - \frac{10t_4(r)}{r} +u_4(r) \right)$ &
1071 < $\left(\frac{45g(r)}{r^4} -\frac{45h(r)}{r^3} +\frac{21s(r)}{r^2} -\frac{6t(r)}{r} +u(r) \right)$\\
1072 < &&& $~~~- \left(\frac{45g(r_c)}{r_c^4} -\frac{45h(r_c)}{r_c^3}
1073 < +\frac{21s(r_c)}{r_c^2} -\frac{6t(r_c)}{r_c} +u(r_c) \right) $\\
1074 < &&& $~~~-\frac{4v_{43}(r)}{r}$  \\
1070 > $\left(\frac{105g_4(r)}{r^4} - \frac{105h_4(r)}{r^3} \right.$ &
1071 > $\left(\frac{45g(r)}{r^4} -\frac{45h(r)}{r^3} +\frac{21s(r)}{r^2}\right.$ &
1072 > SP $- \left(\frac{45g(r_c)}{r_c^4} -\frac{45h(r_c)}{r_c^3}\right.$ \\
1073 > && $~~~\left.+ \frac{45s_4(r)}{r^2} - \frac{10t_4(r)}{r} +u_4(r) \right)$
1074 > & $~~~\left. -\frac{6t(r)}{r} +u(r) \right) -\frac{4v_{43}(r)}{r}$ &
1075 > $\phantom{SP-} \left.+\frac{21s(r_c)}{r_c^2} -\frac{6t(r_c)}{r_c} +u(r_c) \right) $\\
1076   %
1077   $w_n(r)$ &
1078   $\frac{v_{42}(r)}{r}$ &
1079   $\left(\frac{3g_4(r)}{r^4} -\frac{3h_4(r)}{r^3} +\frac{s_4(r)}{r^2}  \right)$ &
1080 + $\frac{v_{42}(r)}{r}$&
1081   $\frac{v_{42}(r)}{r}$\\
1082   %
1083   $w_o(r)$ &
1084   $\frac{v_{43}(r)}{r}$&
1085   $\left(-\frac{15g_4(r)}{r^4} +\frac{15h_4(r)}{r^3} -\frac{6s_4(r)}{r^2} +\frac{t_4(r)}{r} \right)$ &
1086 + $\frac{v_{43}(r)}{r}$&
1087   $\frac{v_{43}(r)}{r}$  \\ \hline
1088   %
1089  
# Line 982 | Line 1105 | F_{\bf a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_
1105   \quad \text{and} \quad  F_{\bf b \alpha}
1106   = - \hat{r}_\alpha \frac{\partial U_{C_{\bf a}C_{\bf b}}} {\partial r}  .
1107   \end{equation}
985 %
986 Obtaining the force from the interaction energy expressions is the
987 same for higher-order multipole interactions -- the trick is to make
988 sure that all $r$-dependent derivatives are considered.  This is
989 straighforward if the interaction energies are written explicitly in
990 terms of $\hat{r}$ and the body axes ($\hat{a}_m$,
991 $\hat{b}_n$) :
992 %
993 \begin{equation}
994 U(r,\{\hat{a}_m \cdot \hat{r} \},
995 \{\hat{b}_n\cdot \hat{r} \},
996 \{\hat{a}_m \cdot \hat{b}_n \}) .
997 \label{ugeneral}
998 \end{equation}
999 %
1000 Allen and Germano,\cite{Allen:2006fk} showed that if the energy is
1001 written in this form, the forces come out relatively cleanly,
1002 %
1003 \begin{equation}
1004 \mathbf{F}_{\bf a}=-\mathbf{F}_{\bf b} =  \frac{\partial U}{\partial \mathbf{r}}
1005 = \frac{\partial U}{\partial r} \hat{r}
1006 + \sum_m \left[
1007 \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
1008 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
1009 + \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
1010 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
1011 \right] \label{forceequation}.
1012 \end{equation}
1013 %
1014 Note that our definition of $\mathbf{r}=\mathbf{r}_b - \mathbf{r}_b $
1015 is opposite in sign to that of Allen and Germano.\cite{Allen:2006fk}
1016 In simplifying the algebra, we have also used:
1017 %
1018 \begin{align}
1019 \frac { \partial (\hat{a}_m \cdot \hat{r})}{\partial \mathbf{r}}
1020 =& \frac{1}{r} \left(  \hat{a}_m - (\hat{a}_m \cdot \hat{r})\hat{r}
1021 \right) \\
1022 \frac { \partial (\hat{b}_m \cdot \hat{r})}{\partial \mathbf{r}}
1023 =& \frac{1}{r} \left(  \hat{b}_m - (\hat{b}_m \cdot \hat{r})\hat{r}
1024 \right) .
1025 \end{align}
1108   %
1109   We list below the force equations written in terms of lab-frame
1110 < coordinates.  The radial functions used in the two methods are listed
1110 > coordinates.  The radial functions used in the three methods are listed
1111   in Table \ref{tab:tableFORCE}
1112   %
1113   %SPACE COORDINATES FORCE EQUATIONS
# Line 1134 | Line 1216 | w_i(r)
1216   \begin{split}
1217   \mathbf{F}_{{\bf a}Q_{\bf a}Q_{\bf b}} =&
1218   \Bigl[
1219 < \text{Tr}\mathbf{Q}_{\mathbf{a}} \text{Tr}\mathbf{Q}_{\mathbf{b}} \hat{r}
1220 < + 2 \text{Tr} ( \mathbf{Q}_{\mathbf{a}} \cdot  \mathbf{Q}_{\mathbf{b}} ) \Bigr] w_k(r) \hat{r} \\
1219 > \text{Tr}\mathbf{Q}_{\mathbf{a}} \text{Tr}\mathbf{Q}_{\mathbf{b}}
1220 > + 2  \mathbf{Q}_{\mathbf{a}} :  \mathbf{Q}_{\mathbf{b}} \Bigr] w_k(r) \hat{r} \\
1221   % 2
1222   &+ \Bigl[
1223   2\text{Tr}\mathbf{Q}_{\mathbf{b}}  (\hat{r} \cdot \mathbf{Q}_{\mathbf{a}} )  
# Line 1171 | Line 1253 | When energies are written in the form of Eq.~({\ref{ug
1253   % Torques SECTION -----------------------------------------------------------------------------------------
1254   %
1255   \subsection{Torques}
1256 < When energies are written in the form of Eq.~({\ref{ugeneral}), then
1175 <  torques can be found in a relatively straightforward
1176 <  manner,\cite{Allen:2006fk}
1256 >
1257   %
1258 < \begin{eqnarray}
1259 < \mathbf{\tau}_{\bf a} =
1180 < \sum_m
1181 < \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{r})}
1182 < ( \hat{r} \times \hat{a}_m )
1183 < -\sum_{mn}
1184 < \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
1185 < (\hat{a}_m \times \hat{b}_n) \\
1186 < %
1187 < \mathbf{\tau}_{\bf b} =
1188 < \sum_m
1189 < \frac{\partial U}{\partial (\hat{b}_m \cdot \hat{r})}
1190 < ( \hat{r} \times \hat{b}_m)
1191 < +\sum_{mn}
1192 < \frac{\partial U}{\partial (\hat{a}_m \cdot \hat{b}_n)}
1193 < (\hat{a}_m \times \hat{b}_n) .
1194 < \end{eqnarray}
1258 > The torques for the three methods are given in space-frame
1259 > coordinates:
1260   %
1261   %
1197 The torques for both the Taylor-Shifted as well as Gradient-Shifted
1198 methods are given in space-frame coordinates:
1199 %
1200 %
1262   \begin{align}
1263   \mathbf{\tau}_{{\bf b}C_{\bf a}D_{\bf b}} =&
1264   C_{\bf a}  (\hat{r} \times  \mathbf{D}_{\mathbf{b}}) v_{11}(r) \\
# Line 1343 | Line 1404 | as in Ref. \onlinecite{Smith98}:
1404   \left[\mathbf{A}_{\alpha+1,\beta} \mathbf{B}_{\alpha+2,\beta}
1405    -\mathbf{A}_{\alpha+2,\beta} \mathbf{B}_{\alpha+2,\beta}
1406   \right]
1407 + \label{eq:matrixCross}
1408   \end{equation}
1409   where $\alpha+1$ and $\alpha+2$ are regarded as cyclic
1410   permuations of the matrix indices.
1411  
1412   All of the radial functions required for torques are identical with
1413   the radial functions previously computed for the interaction energies.
1414 < These are tabulated for both shifted force methods in table
1414 > These are tabulated for all three methods in table
1415   \ref{tab:tableenergy}.  The torques for higher multipoles on site
1416   $\mathbf{a}$ interacting with those of lower order on site
1417   $\mathbf{b}$ can be obtained by swapping indices in the expressions
# Line 1361 | Line 1423 | real-space cutoffs. In this paper we test Taylor-shift
1423   computer simulations, it is vital to test against established methods
1424   for computing electrostatic interactions in periodic systems, and to
1425   evaluate the size and sources of any errors that arise from the
1426 < real-space cutoffs. In this paper we test Taylor-shifted and
1427 < Gradient-shifted electrostatics against analytical methods for
1428 < computing the energies of ordered multipolar arrays.  In the following
1429 < paper, we test the new methods against the multipolar Ewald sum for
1430 < computing the energies, forces and torques for a wide range of typical
1431 < condensed-phase (disordered) systems.
1426 > real-space cutoffs. In this paper we test SP, TSF, and GSF
1427 > electrostatics against analytical methods for computing the energies
1428 > of ordered multipolar arrays.  In the following paper, we test the new
1429 > methods against the multipolar Ewald sum for computing the energies,
1430 > forces and torques for a wide range of typical condensed-phase
1431 > (disordered) systems.
1432  
1433   Because long-range electrostatic effects can be significant in
1434   crystalline materials, ordered multipolar arrays present one of the
# Line 1376 | Line 1438 | and other periodic structures.  We have repeated the L
1438   magnetization and obtained a number of these constants.\cite{Sauer}
1439   This theory was developed more completely by Luttinger and
1440   Tisza\cite{LT,LT2} who tabulated energy constants for the Sauer arrays
1441 < and other periodic structures.  We have repeated the Luttinger \&
1380 < Tisza series summations to much higher order and obtained the energy
1381 < constants (converged to one part in $10^9$) in table \ref{tab:LT}.
1441 > and other periodic structures.  
1442  
1443 < \begin{table*}[h]
1444 < \centering{
1445 <  \caption{Luttinger \& Tisza arrays and their associated
1446 <    energy constants. Type "A" arrays have nearest neighbor strings of
1447 <    antiparallel dipoles.  Type "B" arrays have nearest neighbor
1448 <    strings of antiparallel dipoles if the dipoles are contained in a
1449 <    plane perpendicular to the dipole direction that passes through
1390 <    the dipole.}
1391 < }
1392 < \label{tab:LT}
1393 < \begin{ruledtabular}
1394 < \begin{tabular}{cccc}
1395 < Array Type &  Lattice &   Dipole Direction &    Energy constants \\ \hline
1396 <   A     &      SC       &      001         &      -2.676788684 \\
1397 <   A     &      BCC      &      001         &       0 \\
1398 <   A     &      BCC      &      111         &      -1.770078733 \\
1399 <   A     &      FCC      &      001         &       2.166932835 \\
1400 <   A     &      FCC      &      011         &      -1.083466417 \\
1401 <   B     &      SC       &      001         &      -2.676788684 \\
1402 <   B     &      BCC      &      001         &      -1.338394342 \\
1403 <   B     &      BCC      &      111         &      -1.770078733 \\
1404 <   B     &      FCC      &      001         &      -1.083466417 \\
1405 <   B     &      FCC      &      011         &      -1.807573634 \\
1406 <  --     &      BCC      &    minimum       &      -1.985920929 \\
1407 < \end{tabular}
1408 < \end{ruledtabular}
1409 < \end{table*}
1410 <
1411 < In addition to the A and B arrays, there is an additional minimum
1443 > To test the new electrostatic methods, we have constructed very large,
1444 > $N=$ 16,000~(bcc) arrays of dipoles in the orientations described in
1445 > Ref. \onlinecite{LT}.  These structures include ``A'' lattices with
1446 > nearest neighbor chains of antiparallel dipoles, as well as ``B''
1447 > lattices with nearest neighbor strings of antiparallel dipoles if the
1448 > dipoles are contained in a plane perpendicular to the dipole direction
1449 > that passes through the dipole.  We have also studied the minimum
1450   energy structure for the BCC lattice that was found by Luttinger \&
1451 < Tisza.  The total electrostatic energy for any of the arrays is given
1452 < by:
1451 > Tisza.  The total electrostatic energy density for any of the arrays
1452 > is given by:
1453   \begin{equation}
1454    E = C N^2 \mu^2
1455   \end{equation}
1456 < where $C$ is the energy constant given in table \ref{tab:LT}, $N$ is
1457 < the number of dipoles per unit volume, and $\mu$ is the strength of
1458 < the dipole.
1456 > where $C$ is the energy constant (equivalent to the Madelung
1457 > constant), $N$ is the number of dipoles per unit volume, and $\mu$ is
1458 > the strength of the dipole. Energy constants (converged to 1 part in
1459 > $10^9$) are given in the supplemental information.
1460  
1461 < To test the new electrostatic methods, we have constructed very large,
1462 < $N$ = 8,000~(sc), 16,000~(bcc), or 32,000~(fcc) arrays of dipoles in
1463 < the orientations described in table \ref{tab:LT}.  For the purposes of
1464 < testing the energy expressions and the self-neutralization schemes,
1465 < the primary quantity of interest is the analytic energy constant for
1466 < the perfect arrays.  Convergence to these constants are shown as a
1467 < function of both the cutoff radius, $r_c$, and the damping parameter,
1468 < $\alpha$ in Figs.  \ref{fig:energyConstVsCutoff} and XXX. We have
1469 < simultaneously tested a hard cutoff (where the kernel is simply
1470 < truncated at the cutoff radius), as well as a shifted potential (SP)
1471 < form which includes a potential-shifting and self-interaction term,
1433 < but does not shift the forces and torques smoothly at the cutoff
1434 < radius.
1461 > \begin{figure}
1462 > \includegraphics[width=\linewidth]{Dipoles_rCutNew.pdf}
1463 > \caption{Convergence of the lattice energy constants as a function of
1464 >  cutoff radius (normalized by the lattice constant, $a$) for the new
1465 >  real-space methods.  Three dipolar crystal structures were sampled,
1466 >  and the analytic energy constants for the three lattices are
1467 >  indicated with grey dashed lines.  The left panel shows results for
1468 >  the undamped kernel ($1/r$), while the damped error function kernel,
1469 >  $B_0(r)$ was used in the right panel.}
1470 > \label{fig:Dipoles_rCut}
1471 > \end{figure}
1472  
1473   \begin{figure}
1474 < \includegraphics[width=4.5in]{energyConstVsCutoff}
1475 < \caption{Convergence to the analytic energy constants as a function of
1476 <  cutoff radius (normalized by the lattice constant) for the different
1477 <  real-space methods. The two crystals shown here are the ``B'' array
1478 <  for bcc crystals with the dipoles along the 001 direction (upper),
1479 <  as well as the minimum energy bcc lattice (lower).  The analytic
1480 <  energy constants are shown as a grey dashed line.  The left panel
1481 <  shows results for the undamped kernel ($1/r$), while the damped
1445 <  error function kernel, $B_0(r)$ was used in the right panel. }
1446 < \label{fig:energyConstVsCutoff}
1474 > \includegraphics[width=\linewidth]{Dipoles_alphaNew.pdf}
1475 > \caption{Convergence to the lattice energy constants as a function of
1476 >  the reduced damping parameter ($\alpha^* = \alpha a$) for the
1477 >  different real-space methods in the same three dipolar crystals in
1478 >  Figure \ref{fig:Dipoles_rCut}.  The left panel shows results for a
1479 >  relatively small cutoff radius ($r_c = 4.5 a$) while a larger cutoff
1480 >  radius ($r_c = 6 a$) was used in the right panel. }
1481 > \label{fig:Dipoles_alpha}
1482   \end{figure}
1483  
1484 < The Hard cutoff exhibits oscillations around the analytic energy
1485 < constants, and converges to incorrect energies when the complementary
1486 < error function damping kernel is used.  The shifted potential (SP) and
1487 < gradient-shifted force (GSF) approximations converge to the correct
1488 < energy smoothly by $r_c / 6 a$ even for the undamped case.  This
1489 < indicates that the correction provided by the self term is required
1490 < for obtaining accurate energies.  The Taylor-shifted force (TSF)
1491 < approximation appears to perturb the potential too much inside the
1457 < cutoff region to provide accurate measures of the energy constants.
1484 > For the purposes of testing the energy expressions and the
1485 > self-neutralization schemes, the primary quantity of interest is the
1486 > analytic energy constant for the perfect arrays.  Convergence to these
1487 > constants are shown as a function of both the cutoff radius, $r_c$,
1488 > and the damping parameter, $\alpha$ in Figs.\ref{fig:Dipoles_rCut}
1489 > and \ref{fig:Dipoles_alpha}. We have simultaneously tested a hard
1490 > cutoff (where the kernel is simply truncated at the cutoff radius) in
1491 > addition to the three new methods.
1492  
1493 + The hard cutoff exhibits oscillations around the analytic energy
1494 + constants, and converges to incorrect energies when the complementary
1495 + error function damping kernel is used.  The shifted potential (SP)
1496 + converges to the correct energy smoothly by $r_c = 4.5 a$ even for the
1497 + undamped case.  This indicates that the shifting and the correction
1498 + provided by the self term are  required for obtaining accurate energies.
1499 + The Taylor-shifted force (TSF) approximation appears to perturb the
1500 + potential too much inside the cutoff region to provide accurate
1501 + measures of the energy constants.  GSF is a compromise, converging to
1502 + the correct energies within $r_c = 6 a$.
1503  
1504   {\it Quadrupolar} analogues to the Madelung constants were first
1505   worked out by Nagai and Nakamura who computed the energies of selected
1506   quadrupole arrays based on extensions to the Luttinger and Tisza
1507 < approach.\cite{Nagai01081960,Nagai01091963} We have compared the
1464 < energy constants for the lowest energy configurations for linear
1465 < quadrupoles shown in table \ref{tab:NNQ}
1507 > approach.\cite{Nagai01081960,Nagai01091963}
1508  
1467 \begin{table*}
1468 \centering{
1469  \caption{Nagai and Nakamura Quadurpolar arrays}}
1470 \label{tab:NNQ}
1471 \begin{ruledtabular}
1472 \begin{tabular}{ccc}
1473 Lattice &   Quadrupole Direction &    Energy constants \\ \hline
1474  SC       &      111         &      -8.3 \\
1475  BCC      &      011         &      -21.7 \\
1476  FCC      &      111         &      -80.5
1477 \end{tabular}
1478 \end{ruledtabular}
1479 \end{table*}
1480
1509   In analogy to the dipolar arrays, the total electrostatic energy for
1510   the quadrupolar arrays is:
1511   \begin{equation}
1512 < E = C \frac{3}{4} N^2 Q^2
1512 > E = C N \frac{3\bar{Q}^2}{4a^5}
1513   \end{equation}
1514 < where $Q$ is the quadrupole moment.
1514 > where $a$ is the lattice parameter, and $\bar{Q}$ is the effective
1515 > quadrupole moment,
1516 > \begin{equation}
1517 > \bar{Q}^2 = 2 \left(3 Q : Q - (\text{Tr} Q)^2 \right)
1518 > \end{equation}
1519 > for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}.
1520 > (For the traceless quadrupole tensor, $\Theta = 3 Q - \text{Tr} Q$,
1521 > the effective moment, $\bar{Q}^2 = \frac{2}{3} \Theta : \Theta$.)
1522  
1523 + To test the new electrostatic methods for quadrupoles, we have
1524 + constructed very large, $N=$ 8,000~(sc), 16,000~(bcc), and
1525 + 32,000~(fcc) arrays of linear quadrupoles in the orientations
1526 + described in Ref. \onlinecite{Nagai01081960}.  We have compared the
1527 + energy constants for these low-energy configurations for linear
1528 + quadrupoles. Convergence to these constants are shown as a function of
1529 + both the cutoff radius, $r_c$, and the damping parameter, $\alpha$ in
1530 + Figs.~\ref{fig:Quadrupoles_rCut} and \ref{fig:Quadrupoles_alpha}.
1531 +
1532 + \begin{figure}
1533 + \includegraphics[width=\linewidth]{Quadrupoles_rcutConvergence.pdf}
1534 + \caption{Convergence of the lattice energy constants as a function of
1535 +  cutoff radius (normalized by the lattice constant, $a$) for the new
1536 +  real-space methods.  Three quadrupolar crystal structures were
1537 +  sampled, and the analytic energy constants for the three lattices
1538 +  are indicated with grey dashed lines.  The left panel shows results
1539 +  for the undamped kernel ($1/r$), while the damped error function
1540 +  kernel, $B_0(r)$ was used in the right panel.}
1541 + \label{fig:Quadrupoles_rCut}
1542 + \end{figure}
1543 +
1544 +
1545 + \begin{figure}[!htbp]  
1546 + \includegraphics[width=\linewidth]{Quadrupoles_newAlpha.pdf}
1547 + \caption{Convergence to the lattice energy constants as a function of
1548 +  the reduced damping parameter ($\alpha^* = \alpha a$) for the
1549 +  different real-space methods in the same three quadrupolar crystals
1550 +  in Figure \ref{fig:Quadrupoles_rCut}.  The left panel shows
1551 +  results for a relatively small cutoff radius ($r_c = 4.5 a$) while a
1552 +  larger cutoff radius ($r_c = 6 a$) was used in the right panel.  }
1553 + \label{fig:Quadrupoles_alpha}
1554 + \end{figure}
1555 +
1556 + Again, we find that the hard cutoff exhibits oscillations around the
1557 + analytic energy constants.  The shifted potential (SP) approximation
1558 + converges to the correct energy smoothly by $r_c = 3 a$ even for the
1559 + undamped case.  The Taylor-shifted force (TSF) approximation again
1560 + appears to perturb the potential too much inside the cutoff region to
1561 + provide accurate measures of the energy constants.  GSF again provides
1562 + a compromise between the two methods -- energies are converged by $r_c
1563 + = 4.5 a$, and the approximation is not as perturbative at short range
1564 + as TSF.
1565 +
1566 + It is also useful to understand the convergence to the lattice energy
1567 + constants as a function of the reduced damping parameter ($\alpha^* =
1568 + \alpha a$) for the different real-space methods.
1569 + Figures. \ref{fig:Dipoles_alpha} and \ref{fig:Quadrupoles_alpha} show
1570 + this comparison for the dipolar and quadrupolar lattices,
1571 + respectively.  All of the methods (except for TSF) have excellent
1572 + behavior for the undamped or weakly-damped cases.  All of the methods
1573 + can be forced to converge by increasing the value of $\alpha$, which
1574 + effectively shortens the overall range of the potential by equalizing
1575 + the truncation effects on the different orientational contributions.
1576 + In the second paper in the series, we discuss how large values of
1577 + $\alpha$ can perturb the force and torque vectors, but both
1578 + weakly-damped or over-damped electrostatics appear to generate
1579 + reasonable values for the total electrostatic energies under both the
1580 + SP and GSF approximations.
1581 +
1582   \section{Conclusion}
1583 < We have presented two efficient real-space methods for computing the
1584 < interactions between point multipoles.  These methods have the benefit
1585 < of smoothly truncating the energies, forces, and torques at the cutoff
1586 < radius, making them attractive for both molecular dynamics (MD) and
1587 < Monte Carlo (MC) simulations.  We find that the Gradient-Shifted Force
1588 < (GSF) and the Shifted-Potential (SP) methods converge rapidly to the
1589 < correct lattice energies for ordered dipolar and quadrupolar arrays,
1590 < while the Taylor-Shifted Force (TSF) is too severe an approximation to
1591 < provide accurate convergence to lattice energies.  
1583 > We have presented three efficient real-space methods for computing the
1584 > interactions between point multipoles.  One of these (SP) is a
1585 > multipolar generalization of Wolf's method that smoothly shifts
1586 > electrostatic energies to zero at the cutoff radius.  Two of these
1587 > methods (GSF and TSF) also smoothly truncate the forces and torques
1588 > (in addition to the energies) at the cutoff radius, making them
1589 > attractive for both molecular dynamics and Monte Carlo simulations. We
1590 > find that the Gradient-Shifted Force (GSF) and the Shifted-Potential
1591 > (SP) methods converge rapidly to the correct lattice energies for
1592 > ordered dipolar and quadrupolar arrays, while the Taylor-Shifted Force
1593 > (TSF) is too severe an approximation to provide accurate convergence
1594 > to lattice energies.
1595  
1596   In most cases, GSF can obtain nearly quantitative agreement with the
1597   lattice energy constants with reasonably small cutoff radii.  The only
# Line 1506 | Line 1603 | In large systems, these new methods can be made to sca
1603   crystals with net-zero moments, so this is not expected to be an issue
1604   in most simulations.
1605  
1606 + The techniques used here to derive the force, torque and energy
1607 + expressions can be extended to higher order multipoles, although some
1608 + of the objects (e.g. the matrix cross product in
1609 + Eq. \ref{eq:matrixCross}) will need to be generalized for higher-rank
1610 + tensors.  We also note that the definitions of the multipoles used
1611 + here are in a primitive form, and these need some care when comparing
1612 + with experiment or other computational techniques.
1613 +
1614   In large systems, these new methods can be made to scale approximately
1615   linearly with system size, and detailed comparisons with the Ewald sum
1616   for a wide range of chemical environments follows in the second paper.
# Line 1513 | Line 1618 | for a wide range of chemical environments follows in t
1618   \begin{acknowledgments}
1619    JDG acknowledges helpful discussions with Christopher
1620    Fennell. Support for this project was provided by the National
1621 <  Science Foundation under grant CHE-0848243. Computational time was
1621 >  Science Foundation under grant CHE-1362211. Computational time was
1622    provided by the Center for Research Computing (CRC) at the
1623    University of Notre Dame.
1624   \end{acknowledgments}
# Line 1552 | Line 1657 | is very useful for building up higher derivatives.  Us
1657   \text{e}^{-{\alpha}^2r^2}
1658   \right] ,
1659   \end{equation}
1660 < is very useful for building up higher derivatives.  Using these
1661 < formulas, we find:
1660 > is very useful for building up higher derivatives. As noted by Smith,
1661 > it is possible to approximate the $B_l(r)$ functions,
1662   %
1558 \begin{align}
1559 \frac{dB_0}{dr}=&-rB_1(r) \\
1560 \frac{d^2B_0}{dr^2}=&    - B_1(r) + r^2 B_2(r) \\
1561 \frac{d^3B_0}{dr^3}=&   3 r B_2(r) - r^3 B_3(r) \\
1562 \frac{d^4B_0}{dr^4}=&   3 B_2(r) - 6 r^2 B_3(r) + r^4 B_4(r) \\
1563 \frac{d^5B_0}{dr^5}=& - 15 r B_3(r) + 10 r^3 B_4(r) - r^5 B_5(r) .
1564 \end{align}
1565 %
1566 As noted by Smith, it is possible to approximate the $B_l(r)$
1567 functions,
1568 %
1663   \begin{equation}
1664   B_l(r)=\frac{(2l)!}{l!2^lr^{2l+1}} - \frac {(2\alpha^2)^{l+1}}{(2l+1)\alpha \sqrt{\pi}}
1665   +\text{O}(r) .
1666   \end{equation}
1667   \newpage
1668   \section{The $r$-dependent factors for TSF electrostatics}
1669 + \label{radialTSF}
1670  
1671   Using the shifted damped functions $f_n(r)$ defined by:
1672   %
# Line 1609 | Line 1704 | u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1704   \begin{equation}
1705   u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1706   \end{equation}
1707 <
1707 > % The functions
1708 > % needed are listed schematically below:
1709 > % %
1710 > % \begin{eqnarray}
1711 > % f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1712 > % g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1713 > % h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1714 > % s_2 \quad s_3 \quad &s_4 \nonumber \\
1715 > % t_3 \quad &t_4 \nonumber \\
1716 > % &u_4 \nonumber .
1717 > % \end{eqnarray}
1718   The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and
1719 < stored on a grid for values of $r$ from $0$ to $r_c$.  The functions
1720 < needed are listed schematically below:
1719 > stored on a grid for values of $r$ from $0$ to $r_c$.  Using these
1720 > functions, we find
1721   %
1617 \begin{eqnarray}
1618 f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1619 g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1620 h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1621 s_2 \quad s_3 \quad &s_4 \nonumber \\
1622 t_3 \quad &t_4 \nonumber \\
1623 &u_4 \nonumber .
1624 \end{eqnarray}
1625
1626 Using these functions, we find
1627 %
1722   \begin{align}
1723   \frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\
1724   \frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r}
1725   +r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\
1726 < \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta r_\gamma} =&
1726 > \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =&
1727   \left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta +
1728   \delta_{ \beta \gamma} r_\alpha \right)  
1729 < \left(  -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right)
1730 < + r_\alpha r_\beta r_\gamma
1729 > \left(  -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\
1730 > & + r_\alpha r_\beta r_\gamma
1731   \left(  \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\
1732 < \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta r_\gamma r_\delta} =&
1732 > \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial
1733 >  r_\gamma \partial r_\delta} =&
1734   \left( \delta_{\alpha \beta} \delta_{\gamma \delta}
1735   + \delta_{\alpha \gamma} \delta_{\beta \delta}
1736   +\delta_{ \beta \gamma} \delta_{\alpha \delta} \right)
# Line 1648 | Line 1743 | + \frac{t_n}{r^4} \right)\\
1743   \left(  -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5}
1744   + \frac{t_n}{r^4} \right)\\
1745   \frac{\partial^5 f_n}
1746 < {\partial r_\alpha \partial r_\beta r_\gamma r_\delta r_\epsilon} =&
1746 > {\partial r_\alpha \partial r_\beta \partial r_\gamma \partial
1747 >  r_\delta \partial r_\epsilon} =&
1748   \left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon
1749   + \text{14 permutations} \right)
1750   \left(  \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\
# Line 1665 | Line 1761 | - \frac{10t_n}{r^6} +\frac{u_n}{r^5} \right) \label{eq
1761   %
1762   \newpage
1763   \section{The $r$-dependent factors for GSF electrostatics}
1764 + \label{radialGSF}
1765  
1766   In Gradient-shifted force electrostatics, the kernel is not expanded,
1767 < rather the individual terms in the multipole interaction energies.
1768 < For damped charges , this still brings into the algebra multiple
1769 < derivatives of the Smith's $B_0(r)$ function.  To denote these terms,
1770 < we generalize the notation of the previous appendix.  For $f(r)=1/r$
1771 < (bare Coulomb) or $f(r)=B_0(r)$ (smeared charge)
1767 > and the expansion is carried out on the individual terms in the
1768 > multipole interaction energies. For damped charges, this still brings
1769 > multiple derivatives of the Smith's $B_0(r)$ function into the
1770 > algebra. To denote these terms, we generalize the notation of the
1771 > previous appendix. For either $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$
1772 > (damped),
1773   %
1774   \begin{align}
1775 < g(r)=& \frac{df}{d r}\\
1776 < h(r)=& \frac{dg}{d r} = \frac{d^2f}{d r^2} \\
1777 < s(r)=& \frac{dh}{d r} = \frac{d^3f}{d r^3} \\
1778 < t(r)=& \frac{ds}{d r} = \frac{d^4f}{d r^4} \\
1779 < u(r)=& \frac{dt}{d r} = \frac{d^5f}{d r^5} .
1775 > g(r) &= \frac{df}{d r} &&                      &&=-\frac{1}{r^2}
1776 > &&\mathrm{or~~~} -rB_1(r) \\
1777 > h(r) &= \frac{dg}{d r} &&= \frac{d^2f}{d r^2} &&= \frac{2}{r^3} &&\mathrm{or~~~}-B_1(r) + r^2 B_2(r) \\
1778 > s(r) &= \frac{dh}{d r} &&= \frac{d^3f}{d r^3} &&=-\frac{6}{r^4}&&\mathrm{or~~~}3rB_2(r) - r^3 B_3(r)\\
1779 > t(r) &= \frac{ds}{d r} &&= \frac{d^4f}{d r^4} &&= \frac{24}{r^5} &&\mathrm{or~~~} 3
1780 > B_2(r) - 6r^2 B_3(r) + r^4 B_4(r) \\
1781 > u(r) &= \frac{dt}{d r} &&= \frac{d^5f}{d r^5} &&=-\frac{120}{r^6} &&\mathrm{or~~~} -15
1782 > r B_3(r) + 10 r^3B_4(r) -r^5B_5(r).
1783   \end{align}
1784   %
1785 < For undamped charges, $f(r)=1/r$, Table I lists these derivatives
1786 < under the column ``Bare Coulomb.''  Equations \ref{eq:b9} to
1787 < \ref{eq:b13} are still correct for GSF electrostatics if the subscript
1687 < $n$ is eliminated.
1785 > For undamped charges, Table I lists these derivatives under the Bare
1786 > Coulomb column. Equations \ref{eq:b9} to \ref{eq:b13} are still
1787 > correct for GSF electrostatics if the subscript $n$ is eliminated.
1788  
1789   \newpage
1790  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines