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 4173 by gezelter, Thu Jun 5 17:09:34 2014 UTC

# Line 152 | Line 152 | V = \sum_i \sum_{j>i} V_\mathrm{pair}(r_{ij}, \Omega_i
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{correction}
155 > V = \sum_i \sum_{j>i} V_\mathrm{pair}(\mathbf{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\{
160 >  V_\mathrm{pair}(\mathbf{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 ,
162 > V_\mathrm{approx} (\mathbf{r}_{ij}, \Omega_i, \Omega_j) & \quad \left| \mathbf{r}_{ij} \right| \le r_c \\
163 > 0 & \quad \left| \mathbf{r}_{ij} \right|  > r_c ,
164   \end{array}
165   \right.
166   \end{equation}
167 < along with an easily computed correction term ($\sum_i
168 < V_i^\mathrm{correction}$) which has linear-scaling with the number of
167 > along with an easily computed self-interaction term ($\sum_i
168 > V_i^\mathrm{self}$) which scales linearly with the number of
169   particles.  Here $\Omega_i$ and $\Omega_j$ represent orientational
170 < coordinates of the two sites.  The computational efficiency, energy
170 > coordinates of the two sites, and $\mathbf{r}_{ij}$ is the vector
171 > between the two sites.  The computational efficiency, energy
172   conservation, and even some physical properties of a simulation can
173   depend dramatically on how the $V_\mathrm{approx}$ function behaves at
174   the cutoff radius. The goal of any approximation method should be to
# Line 180 | Line 181 | computed within $r_c$. Damping using a complementary e
181   contained within the cutoff sphere surrounding each particle.  This is
182   accomplished by shifting the potential functions to generate image
183   charges on the surface of the cutoff sphere for each pair interaction
184 < computed within $r_c$. Damping using a complementary error
185 < function is applied to the potential to accelerate convergence. The
186 < potential for the DSF method, where $\alpha$ is the adjustable damping
186 < parameter, is given by
184 > computed within $r_c$. Damping using a complementary error function is
185 > applied to the potential to accelerate convergence. The interaction
186 > for a pair of charges ($C_i$ and $C_j$) in the DSF method,
187   \begin{equation*}
188   V_\mathrm{DSF}(r) = C_i C_j \Biggr{[}
189   \frac{\mathrm{erfc}\left(\alpha r_{ij}\right)}{r_{ij}}
# Line 193 | Line 193 | Note that in this potential and in all electrostatic q
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.
196 > where $\alpha$ is the adjustable damping parameter.  Note that in this
197 > potential and in all electrostatic quantities that follow, the
198 > standard $1/4 \pi \epsilon_{0}$ has been omitted for 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
# Line 250 | Line 250 | labelling specific charges in $\bf a$ and $\bf b$ resp
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
253 > labeling 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} =
# Line 276 | Line 276 | C_{\bf a} =&\sum_{k \, \text{in \bf a}} q_k , \\
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 , \\
280 < D_{{\bf a}\alpha} =&\sum_{k \, \text{in \bf a}} q_k r_{k\alpha} ,\\
281 < Q_{{\bf a}\alpha\beta} =& \frac{1}{2} \sum_{k \, \text{in \bf a}} q_k r_{k\alpha}  r_{k\beta} .
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 $1/2$.
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
# Line 306 | Line 313 | In the standard multipole expansion, one typically use
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
# Line 321 | Line 329 | functions $B_l(r)$ are summarized in Appendix A.
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.
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
# Line 418 | Line 432 | U= (\text{prefactor}) (\text{derivatives}) f_n(r)
432   In general, we can write
433   %
434   \begin{equation}
435 < U= (\text{prefactor}) (\text{derivatives}) f_n(r)
435 > U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
436   \label{generic}
437   \end{equation}
438   %
# Line 439 | Line 453 | presented in tables \ref{tab:tableenergy} and \ref{tab
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
# Line 446 | Line 465 | U^{\text{shift}}(r)=U(r)-U(r_c)-(r-r_c)\hat{r}\cdot \n
465   expansion, and has a similar interaction energy for all multipole
466   orders:
467   \begin{equation}
468 < U^{\text{shift}}(r)=U(r)-U(r_c)-(r-r_c)\hat{r}\cdot \nabla U(r) \Big
469 < \lvert  _{r_c} .
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 < Here the gradient for force shifting is evaluated for an image
474 < multipole projected onto the surface of the cutoff sphere (see fig
475 < \ref{fig:shiftedMultipoles}). No higher order terms $(r-r_c)^n$
476 < appear.  The primary difference between the TSF and GSF methods is the
477 < stage at which the Taylor Series is applied; in the Taylor-shifted
478 < approach, it is applied to the kernel itself.  In the Gradient-shifted
479 < approach, it is applied to individual radial interactions terms in the
480 < multipole expansion.  Energies from this method thus have the general
481 < form:
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}
# Line 471 | Line 494 | the radial factors differ between the two methods.
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 < [XXX Do we need this section in the main paper? or should it go in the
476 < extra materials?]
477 <
478 < So far, all energies and forces have been written in terms of fixed
479 < space coordinates.  Interaction energies are computed from the generic
480 < formulas Eq.~(\ref{generic}) and ~(\ref{generic2}) which combine
481 < orientational prefactors with radial functions.  Because objects $\bf
482 < a$ and $\bf b$ both translate and rotate during a molecular dynamics
483 < (MD) simulation, it is desirable to contract all $r$-dependent terms
484 < with dipole and quadrupole moments expressed in terms of their body
485 < axes.  To do so, we have followed the methodology of Allen and
486 < Germano,\cite{Allen:2006fk} which was itself based on earlier work by
487 < Price {\em et al.}\cite{Price:1984fk}
488 <
489 < 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
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}} =
# Line 504 | Line 515 | expressed using these unit vectors:
515   \hat{a}_1 \\
516   \hat{a}_2 \\
517   \hat{a}_3
518 < \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}\\
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 < =
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}  .
526 < \end{eqnarray}
525 > \end{eqnarray}
526   %
527   These matrices convert from space-fixed $(xyz)$ to body-fixed $(123)$
528 < coordinates.  All contractions of prefactors with derivatives of
529 < functions can be written in terms of these matrices. It proves to be
530 < equally convenient to just write any contraction in terms of unit
531 < vectors $\hat{r}$, $\hat{a}_m$, and $\hat{b}_n$. In the torque
532 < expressions, it is useful to have the angular-dependent terms
533 < available in three different fashions, e.g. for the dipole-dipole
534 < contraction:
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}}
# Line 546 | Line 604 | contractions using space indices.
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
617 > extension~\cite{Fennell:2006zl} also include 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
# Line 598 | Line 661 | site which posesses a charge, dipole, and multipole, b
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
664 > site which posesses a charge, dipole, and quadrupole, both types of
665   contribution are given in table \ref{tab:tableSelf}.
666  
667   \begin{table*}
# Line 611 | Line 674 | Quadrupole & $2 \text{Tr}(\mathbf{Q}_{\bf a}^2) + \tex
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 & $2 \text{Tr}(\mathbf{Q}_{\bf a}^2) + \text{Tr}(\mathbf{Q}_{\bf a})^2$ &
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(
# Line 741 | Line 804 | +2 \text{Tr} \left(
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 \text{Tr} \left(
808 < \mathbf{Q}_{\mathbf{a}} \cdot \mathbf{Q}_{\mathbf{b}} \right) \Bigr] v_{41}(r)
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}}
# Line 981 | Line 1044 | F_{\bf a \alpha} = \hat{r}_\alpha \frac{\partial U_{C_
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}  .
984 \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}
1047   \end{equation}
1048   %
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}
1026 %
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}
# Line 1134 | Line 1156 | w_i(r)
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}} \hat{r}
1160 < + 2 \text{Tr} ( \mathbf{Q}_{\mathbf{a}} \cdot  \mathbf{Q}_{\mathbf{b}} ) \Bigr] w_k(r) \hat{r} \\
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}} )  
# Line 1171 | Line 1193 | When energies are written in the form of Eq.~({\ref{ug
1193   % Torques SECTION -----------------------------------------------------------------------------------------
1194   %
1195   \subsection{Torques}
1196 < 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}
1196 >
1197   %
1178 \begin{eqnarray}
1179 \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}
1195 %
1196 %
1198   The torques for both the Taylor-Shifted as well as Gradient-Shifted
1199   methods are given in space-frame coordinates:
1200   %
# Line 1355 | Line 1356 | above.
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 an extension of the Wolf approach for point
1361 + multipoles by simply projecting the image multipole onto the surface
1362 + of the cutoff sphere, and including the interactions with the central
1363 + multipole and the image.  This effectively shifts the total potential
1364 + to zero at the cutoff radius,
1365 + \begin{equation}
1366 + U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
1367 + U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
1368 + \label{eq:SP}
1369 + \end{equation}
1370 + where the sum describes separate potential shifting that is applied to
1371 + each orientational contribution to the energy.
1372 +
1373 + The energies and torques for the shifted potential (SP) can be easily
1374 + obtained by zeroing out the $(r-r_c)$ terms in the final column of
1375 + table \ref{tab:tableenergy}.  Forces for the SP method retain
1376 + discontinuities at the cutoff sphere, and can be obtained by
1377 + eliminating all functions that depend on $r_c$ in the last column of
1378 + table \ref{tab:tableFORCE}.  The self-energy contributions for the SP
1379 + potential are identical to both the GSF and TSF methods.
1380 +
1381   \section{Comparison to known multipolar energies}
1382  
1383   To understand how these new real-space multipole methods behave in
1384   computer simulations, it is vital to test against established methods
1385   for computing electrostatic interactions in periodic systems, and to
1386   evaluate the size and sources of any errors that arise from the
1387 < real-space cutoffs. In this paper we test Taylor-shifted and
1388 < Gradient-shifted electrostatics against analytical methods for
1389 < computing the energies of ordered multipolar arrays.  In the following
1390 < paper, we test the new methods against the multipolar Ewald sum for
1391 < computing the energies, forces and torques for a wide range of typical
1392 < condensed-phase (disordered) systems.
1387 > real-space cutoffs. In this paper we test both TSF and GSF
1388 > electrostatics against analytical methods for computing the energies
1389 > of ordered multipolar arrays.  In the following paper, we test the new
1390 > methods against the multipolar Ewald sum for computing the energies,
1391 > forces and torques for a wide range of typical condensed-phase
1392 > (disordered) systems.
1393  
1394   Because long-range electrostatic effects can be significant in
1395   crystalline materials, ordered multipolar arrays present one of the
# Line 1376 | Line 1399 | and other periodic structures.  We have repeated the L
1399   magnetization and obtained a number of these constants.\cite{Sauer}
1400   This theory was developed more completely by Luttinger and
1401   Tisza\cite{LT,LT2} who tabulated energy constants for the Sauer arrays
1402 < 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}.
1402 > and other periodic structures.  
1403  
1404 < \begin{table*}[h]
1405 < \centering{
1406 <  \caption{Luttinger \& Tisza arrays and their associated
1407 <    energy constants. Type "A" arrays have nearest neighbor strings of
1408 <    antiparallel dipoles.  Type "B" arrays have nearest neighbor
1409 <    strings of antiparallel dipoles if the dipoles are contained in a
1410 <    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
1404 > To test the new electrostatic methods, we have constructed very large,
1405 > $N=$ 16,000~(bcc) arrays of dipoles in the orientations described in
1406 > Ref. \onlinecite{LT}.  These structures include ``A'' lattices with
1407 > nearest neighbor chains of antiparallel dipoles, as well as ``B''
1408 > lattices with nearest neighbor strings of antiparallel dipoles if the
1409 > dipoles are contained in a plane perpendicular to the dipole direction
1410 > that passes through the dipole.  We have also studied the minimum
1411   energy structure for the BCC lattice that was found by Luttinger \&
1412   Tisza.  The total electrostatic energy for any of the arrays is given
1413   by:
1414   \begin{equation}
1415    E = C N^2 \mu^2
1416   \end{equation}
1417 < where $C$ is the energy constant given in table \ref{tab:LT}, $N$ is
1418 < the number of dipoles per unit volume, and $\mu$ is the strength of
1419 < the dipole.
1417 > where $C$ is the energy constant (equivalent to the Madelung
1418 > constant), $N$ is the number of dipoles per unit volume, and $\mu$ is
1419 > the strength of the dipole. Energy constants (converged to 1 part in
1420 > $10^9$) are given in the supplemental information.
1421  
1422 To test the new electrostatic methods, we have constructed very large,
1423 $N$ = 8,000~(sc), 16,000~(bcc), or 32,000~(fcc) arrays of dipoles in
1424 the orientations described in table \ref{tab:LT}.  For the purposes of
1425 testing the energy expressions and the self-neutralization schemes,
1426 the primary quantity of interest is the analytic energy constant for
1427 the perfect arrays.  Convergence to these constants are shown as a
1428 function of both the cutoff radius, $r_c$, and the damping parameter,
1429 $\alpha$ in Figs.  \ref{fig:energyConstVsCutoff} and XXX. We have
1430 simultaneously tested a hard cutoff (where the kernel is simply
1431 truncated at the cutoff radius), as well as a shifted potential (SP)
1432 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.
1435
1422   \begin{figure}
1423 < \includegraphics[width=4.5in]{energyConstVsCutoff}
1424 < \caption{Convergence to the analytic energy constants as a function of
1425 <  cutoff radius (normalized by the lattice constant) for the different
1426 <  real-space methods. The two crystals shown here are the ``B'' array
1427 <  for bcc crystals with the dipoles along the 001 direction (upper),
1428 <  as well as the minimum energy bcc lattice (lower).  The analytic
1429 <  energy constants are shown as a grey dashed line.  The left panel
1430 <  shows results for the undamped kernel ($1/r$), while the damped
1431 <  error function kernel, $B_0(r)$ was used in the right panel. }
1432 < \label{fig:energyConstVsCutoff}
1423 > \includegraphics[width=\linewidth]{Dipoles_rCutNew.pdf}
1424 > \caption{Convergence of the lattice energy constants as a function of
1425 >  cutoff radius (normalized by the lattice constant, $a$) for the new
1426 >  real-space methods.  Three dipolar crystal structures were sampled,
1427 >  and the analytic energy constants for the three lattices are
1428 >  indicated with grey dashed lines.  The left panel shows results for
1429 >  the undamped kernel ($1/r$), while the damped error function kernel,
1430 >  $B_0(r)$ was used in the right panel.}
1431 > \label{fig:Dipoles_rCut}
1432 > \end{figure}
1433 >
1434 > \begin{figure}
1435 > \includegraphics[width=\linewidth]{Dipoles_alphaNew.pdf}
1436 > \caption{Convergence to the lattice energy constants as a function of
1437 >  the reduced damping parameter ($\alpha^* = \alpha a$) for the
1438 >  different real-space methods in the same three dipolar crystals in
1439 >  Figure \ref{fig:Dipoles_rCut}.  The left panel shows results for a
1440 >  relatively small cutoff radius ($r_c = 4.5 a$) while a larger cutoff
1441 >  radius ($r_c = 6 a$) was used in the right panel. }
1442 > \label{fig:Dipoles_alpha}
1443   \end{figure}
1444  
1445 + For the purposes of testing the energy expressions and the
1446 + self-neutralization schemes, the primary quantity of interest is the
1447 + analytic energy constant for the perfect arrays.  Convergence to these
1448 + constants are shown as a function of both the cutoff radius, $r_c$,
1449 + and the damping parameter, $\alpha$ in Figs.  \ref{fig:Dipoles_rCut}
1450 + and \ref{fig:Dipoles_alpha}. We have simultaneously tested a hard
1451 + cutoff (where the kernel is simply truncated at the cutoff radius), as
1452 + well as a shifted potential (SP) form which includes a
1453 + potential-shifting and self-interaction term, but does not shift the
1454 + forces and torques smoothly at the cutoff radius.  The SP method is
1455 + essentially an extension of the original Wolf method for multipoles.
1456 +
1457   The Hard cutoff exhibits oscillations around the analytic energy
1458   constants, and converges to incorrect energies when the complementary
1459   error function damping kernel is used.  The shifted potential (SP) and
# Line 1455 | Line 1463 | cutoff region to provide accurate measures of the ener
1463   for obtaining accurate energies.  The Taylor-shifted force (TSF)
1464   approximation appears to perturb the potential too much inside the
1465   cutoff region to provide accurate measures of the energy constants.
1458
1459
1466   {\it Quadrupolar} analogues to the Madelung constants were first
1467   worked out by Nagai and Nakamura who computed the energies of selected
1468   quadrupole arrays based on extensions to the Luttinger and Tisza
1469   approach.\cite{Nagai01081960,Nagai01091963} We have compared the
1470   energy constants for the lowest energy configurations for linear
1471 < quadrupoles shown in table \ref{tab:NNQ}
1471 > quadrupoles.  
1472  
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
1473   In analogy to the dipolar arrays, the total electrostatic energy for
1474   the quadrupolar arrays is:
1475   \begin{equation}
1476 < E = C \frac{3}{4} N^2 Q^2
1476 > E = C N \frac{3\bar{Q}^2}{4a^5}
1477   \end{equation}
1478 < where $Q$ is the quadrupole moment.
1478 > where $a$ is the lattice parameter, and $\bar{Q}$ is the effective
1479 > quadrupole moment,
1480 > \begin{equation}
1481 > \bar{Q}^2 = 2 \left(3 Q : Q - (\text{Tr} Q)^2 \right)
1482 > \end{equation}
1483 > for the primitive quadrupole as defined in Eq. \ref{eq:quadrupole}.
1484 > (For the traceless quadrupole tensor, $\Theta = 3 Q - \text{Tr} Q$,
1485 > the effective moment, $\bar{Q}^2 = \frac{2}{3} \Theta : \Theta$.)
1486  
1487 + \begin{figure}
1488 + \includegraphics[width=\linewidth]{Quadrupoles_rcutConvergence.pdf}
1489 + \caption{Convergence of the lattice energy constants as a function of
1490 +  cutoff radius (normalized by the lattice constant, $a$) for the new
1491 +  real-space methods.  Three quadrupolar crystal structures were
1492 +  sampled, and the analytic energy constants for the three lattices
1493 +  are indicated with grey dashed lines.  The left panel shows results
1494 +  for the undamped kernel ($1/r$), while the damped error function
1495 +  kernel, $B_0(r)$ was used in the right panel.}
1496 + \label{fig:QuadrupolesrcutCovergence}
1497 + \end{figure}
1498 +
1499 +
1500 + \begin{figure}[!htbp]  
1501 + \includegraphics[width=3.5in]{Quadrupoles_alphaConvergence-crop.pdf}
1502 + \caption{Convergence to the analytic energy constants as a function of
1503 +  cutoff damping alpha for the different
1504 +  real-space methods for (a) dipolar and (b) quadrupolar crystals.The energy constants for hard, SP, GSF, TSF, and analytic methods are represented by the black sold-circle, red solid-square, green solid-diamond, and grey dashed line respectively.
1505 + The left panel shows results for the undamped kernel ($1/r$), while the damped
1506 +  error function kernel, $B_0(r)$ was used in the right panel. }
1507 + \label{fig:Quadrupoles_alphaCovergence-crop.pdf}
1508 + \end{figure}
1509 +
1510 +
1511   \section{Conclusion}
1512   We have presented two efficient real-space methods for computing the
1513   interactions between point multipoles.  These methods have the benefit
# Line 1513 | Line 1536 | for a wide range of chemical environments follows in t
1536   \begin{acknowledgments}
1537    JDG acknowledges helpful discussions with Christopher
1538    Fennell. Support for this project was provided by the National
1539 <  Science Foundation under grant CHE-0848243. Computational time was
1539 >  Science Foundation under grant CHE-1362211. Computational time was
1540    provided by the Center for Research Computing (CRC) at the
1541    University of Notre Dame.
1542   \end{acknowledgments}
# Line 1609 | Line 1632 | u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1632   \begin{equation}
1633   u_4(r)=B_0^{(5)}(r) - B_0^{(5)}(r_c) .
1634   \end{equation}
1635 <
1635 > % The functions
1636 > % needed are listed schematically below:
1637 > % %
1638 > % \begin{eqnarray}
1639 > % f_0 \quad f_1 \qquad \qquad \quad & \nonumber \\
1640 > % g_0 \quad g_1 \quad g_2 \quad g_3 \quad &g_4 \nonumber \\
1641 > % h_1 \quad h_2 \quad h_3 \quad &h_4 \nonumber \\
1642 > % s_2 \quad s_3 \quad &s_4 \nonumber \\
1643 > % t_3 \quad &t_4 \nonumber \\
1644 > % &u_4 \nonumber .
1645 > % \end{eqnarray}
1646   The functions $f_n(r)$ to $u_n(r)$ can be computed recursively and
1647 < stored on a grid for values of $r$ from $0$ to $r_c$.  The functions
1648 < needed are listed schematically below:
1647 > stored on a grid for values of $r$ from $0$ to $r_c$.  Using these
1648 > functions, we find
1649   %
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 %
1650   \begin{align}
1651   \frac{\partial f_n}{\partial r_\alpha} =&r_\alpha \frac {g_n}{r} \label{eq:b9}\\
1652   \frac{\partial^2 f_n}{\partial r_\alpha \partial r_\beta} =&\delta_{\alpha \beta}\frac {g_n}{r}
1653   +r_\alpha r_\beta \left( -\frac{g_n}{r^3} +\frac{h_n}{r^2}\right) \\
1654 < \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta r_\gamma} =&
1654 > \frac{\partial^3 f_n}{\partial r_\alpha \partial r_\beta \partial r_\gamma} =&
1655   \left( \delta_{\alpha \beta} r_\gamma + \delta_{\alpha \gamma} r_\beta +
1656   \delta_{ \beta \gamma} r_\alpha \right)  
1657 < \left(  -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right)
1658 < + r_\alpha r_\beta r_\gamma
1657 > \left(  -\frac{g_n}{r^3} +\frac{h_n}{r^2} \right) \nonumber \\
1658 > & + r_\alpha r_\beta r_\gamma
1659   \left(  \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \\
1660 < \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta r_\gamma r_\delta} =&
1660 > \frac{\partial^4 f_n}{\partial r_\alpha \partial r_\beta \partial
1661 >  r_\gamma \partial r_\delta} =&
1662   \left( \delta_{\alpha \beta} \delta_{\gamma \delta}
1663   + \delta_{\alpha \gamma} \delta_{\beta \delta}
1664   +\delta_{ \beta \gamma} \delta_{\alpha \delta} \right)
# Line 1648 | Line 1671 | + \frac{t_n}{r^4} \right)\\
1671   \left(  -\frac{15g_n}{r^7} + \frac{15h_n}{r^6} - \frac{6s_n}{r^5}
1672   + \frac{t_n}{r^4} \right)\\
1673   \frac{\partial^5 f_n}
1674 < {\partial r_\alpha \partial r_\beta r_\gamma r_\delta r_\epsilon} =&
1674 > {\partial r_\alpha \partial r_\beta \partial r_\gamma \partial
1675 >  r_\delta \partial r_\epsilon} =&
1676   \left( \delta_{\alpha \beta} \delta_{\gamma \delta} r_\epsilon
1677   + \text{14 permutations} \right)
1678   \left(  \frac{3g_n}{r^5}-\frac{3h_n}{r^4} +\frac{s_n}{r^3} \right) \nonumber \\
# Line 1670 | Line 1694 | we generalize the notation of the previous appendix.  
1694   rather the individual terms in the multipole interaction energies.
1695   For damped charges , this still brings into the algebra multiple
1696   derivatives of the Smith's $B_0(r)$ function.  To denote these terms,
1697 < we generalize the notation of the previous appendix.  For $f(r)=1/r$
1698 < (bare Coulomb) or $f(r)=B_0(r)$ (smeared charge)
1697 > we generalize the notation of the previous appendix.  For either
1698 > $f(r)=1/r$ (undamped) or $f(r)=B_0(r)$ (damped),
1699   %
1700   \begin{align}
1701   g(r)=& \frac{df}{d r}\\
# Line 1681 | Line 1705 | For undamped charges, $f(r)=1/r$, Table I lists these
1705   u(r)=& \frac{dt}{d r} = \frac{d^5f}{d r^5} .
1706   \end{align}
1707   %
1708 < For undamped charges, $f(r)=1/r$, Table I lists these derivatives
1709 < under the column ``Bare Coulomb.''  Equations \ref{eq:b9} to
1710 < \ref{eq:b13} are still correct for GSF electrostatics if the subscript
1687 < $n$ is eliminated.
1708 > For undamped charges Table I lists these derivatives under the column
1709 > ``Bare Coulomb.''  Equations \ref{eq:b9} to \ref{eq:b13} are still
1710 > correct for GSF electrostatics if the subscript $n$ is eliminated.
1711  
1712   \newpage
1713  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines