ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
(Generate patch)

Comparing trunk/multipole/multipole_2/multipole2.tex (file contents):
Revision 4171 by gezelter, Wed Jun 4 19:31:06 2014 UTC vs.
Revision 4184 by gezelter, Sat Jun 14 23:35:39 2014 UTC

# Line 35 | Line 35 | preprint,
35   %\linenumbers\relax % Commence numbering lines
36   \usepackage{amsmath}
37   \usepackage{times}
38 < \usepackage{mathptm}
38 > \usepackage{mathptmx}
39 > \usepackage{tabularx}
40   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
41   \usepackage{url}
42   \usepackage[english]{babel}
43  
44 + \newcolumntype{Y}{>{\centering\arraybackslash}X}
45  
46   \begin{document}
47  
48 < \preprint{AIP/123-QED}
48 > %\preprint{AIP/123-QED}
49  
50 < \title[Efficient electrostatics for condensed-phase multipoles]{Real space alternatives to the Ewald
51 < Sum. II. Comparison of Simulation Methodologies} % Force line breaks with \\
50 > \title{Real space alternatives to the Ewald
51 > Sum. II. Comparison of Methods} % Force line breaks with \\
52  
53   \author{Madan Lamichhane}
54   \affiliation{Department of Physics, University
# Line 65 | Line 67 | We have tested our recently developed shifted potentia
67               %  but any date may be explicitly specified
68  
69   \begin{abstract}
70 < We have tested our recently developed shifted potential, gradient-shifted force, and Taylor-shifted force methods for the higher-order multipoles against Ewald’s method in different types of liquid and crystalline system. In this paper, we have also investigated the conservation of total energy in the molecular dynamic simulation using all of these methods. The shifted potential method shows better agreement with the Ewald in the energy differences between different configurations as compared to the direct truncation. Both the gradient shifted force and Taylor-shifted force methods reproduce very good energy conservation. But the absolute energy, force and torque evaluated from the gradient shifted force method shows better result as compared to taylor-shifted force method. Hence the gradient-shifted force method suitably mimics the electrostatic interaction in the molecular dynamic simulation.
70 >  We have tested the real-space shifted potential (SP),
71 >  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
72 >  for multipoles that were developed in the first paper in this series
73 >  against the multipolar Ewald sum as a reference method. The tests
74 >  were carried out in a variety of condensed-phase environments which
75 >  were designed to test all levels of the multipole-multipole
76 >  interactions.  Comparisons of the energy differences between
77 >  configurations, molecular forces, and torques were used to analyze
78 >  how well the real-space models perform relative to the more
79 >  computationally expensive Ewald treatment.  We have also investigated the
80 >  energy conservation properties of the new methods in molecular
81 >  dynamics simulations using all of these methods. The SP method shows
82 >  excellent agreement with configurational energy differences, forces,
83 >  and torques, and would be suitable for use in Monte Carlo
84 >  calculations.  Of the two new shifted-force methods, the GSF
85 >  approach shows the best agreement with Ewald-derived energies,
86 >  forces, and torques and exhibits energy conservation properties that
87 >  make it an excellent choice for efficiently computing electrostatic
88 >  interactions in molecular dynamics simulations.
89   \end{abstract}
90  
91 < \pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
91 > %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92                               % Classification Scheme.
93 < \keywords{Electrostatics, Multipoles, Real-space}
93 > %\keywords{Electrostatics, Multipoles, Real-space}
94  
95   \maketitle
96  
# Line 100 | Line 120 | To simulate interfacial systems, Parry’s extension o
120   method may require modification to compute interactions for
121   interfacial molecular systems such as membranes and liquid-vapor
122   interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
123 < To simulate interfacial systems, Parry’s extension of the 3D Ewald sum
123 > To simulate interfacial systems, Parry's extension of the 3D Ewald sum
124   is appropriate for slab geometries.\cite{Parry:1975if} The inherent
125 < periodicity in the Ewald’s method can also be problematic for
126 < interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
127 < methods that were developed to handle two-dimensional (2D)
128 < electrostatic interactions in interfacial systems have not had similar
129 < particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
125 > periodicity in the Ewald method can also be problematic for molecular
126 > interfaces.\cite{Fennell:2006lq} Modified Ewald methods that were
127 > developed to handle two-dimensional (2D) electrostatic interactions in
128 > interfacial systems have not seen similar particle-mesh
129 > treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 >  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
131 > with system size.
132  
133   \subsection{Real-space methods}
134   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
135   method for calculating electrostatic interactions between point
136   charges. They argued that the effective Coulomb interaction in
137 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
138 < an ordered lattice (e.g. when computing the Madelung constant of an
139 < ionic solid), the material can be considered as a set of ions
137 > condensed phase systems is actually short ranged.\cite{Wolf92,Wolf95}
138 > For an ordered lattice (e.g., when computing the Madelung constant of
139 > an ionic solid), the material can be considered as a set of ions
140   interacting with neutral dipolar or quadrupolar ``molecules'' giving
141   an effective distance dependence for the electrostatic interactions of
142 < $r^{-5}$ (see figure \ref{fig:NaCl}.  For this reason, careful
143 < applications of Wolf's method are able to obtain accurate estimates of
144 < Madelung constants using relatively short cutoff radii.  Recently,
145 < Fukuda used neutralization of the higher order moments for the
146 < calculation of the electrostatic interaction of the point charges
147 < system.\cite{Fukuda:2013sf}
142 > $r^{-5}$ (see figure \ref{fig:schematic}).  For this reason, careful
143 > application of Wolf's method can obtain accurate estimates of Madelung
144 > constants using relatively short cutoff radii.  Recently, Fukuda used
145 > neutralization of the higher order moments for calculation of the
146 > electrostatic interactions in point charge
147 > systems.\cite{Fukuda:2013sf}
148  
149 < \begin{figure}[h!]
149 > \begin{figure}
150    \centering
151 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
152 <  \caption{Top: NaCl crystal showing how spherical truncation can
153 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
154 <    molecules interact with the central ion.  Bottom: A dipolar
155 <    crystal exhibiting similar behavior and illustrating how the
156 <    effective dipole-octupole interactions can be disrupted by
157 <    spherical truncation.}
158 <  \label{fig:NaCl}
151 >  \includegraphics[width=\linewidth]{schematic.pdf}
152 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
153 >    charges (in the smaller grey circle), so interactions are
154 >    effectively charge-multipole at longer distances.  With hard
155 >    cutoffs, motion of individual charges in and out of the cutoff
156 >    sphere can break the effective multipolar ordering.  Bottom:
157 >    dipolar crystals and fluids have a similar effective
158 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
159 >    orientational averaging helps to reduce the effective range of the
160 >    interactions in the fluid.  Placement of reversed image multipoles
161 >    on the surface of the cutoff sphere recovers the effective
162 >    higher-order multipole behavior.}
163 >  \label{fig:schematic}
164   \end{figure}
165  
166   The direct truncation of interactions at a cutoff radius creates
167 < truncation defects. Wolf \textit{et al.} further argued that
167 > truncation defects. Wolf \textit{et al.}  argued that
168   truncation errors are due to net charge remaining inside the cutoff
169   sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
170   placing an image charge on the surface of the cutoff sphere for every
# Line 158 | Line 184 | potential is found to be decreasing as $r^{-5}$. If on
184  
185   Considering the interaction of one central ion in an ionic crystal
186   with a portion of the crystal at some distance, the effective Columbic
187 < potential is found to be decreasing as $r^{-5}$. If one views the
188 < \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
187 > potential is found to decrease as $r^{-5}$. If one views the \ce{NaCl}
188 > crystal as a simple cubic (SC) structure with an octupolar
189   \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
190   rapidly to the Madelung energy than the dipolar
191   approximation.\cite{Wolf92} To find the correct Madelung constant,
192   Lacman suggested that the NaCl structure could be constructed in a way
193   that the finite crystal terminates with complete \ce{(NaCl)4}
194 < molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
195 < by opposite charges. Similarly for each pair of charges, there is an
196 < opposite pair of charge adjacent to it.  The central ion sees what is
197 < effectively a set of octupoles at large distances. These facts suggest
172 < that the Madelung constants are relatively short ranged for perfect
173 < ionic crystals.\cite{Wolf:1999dn}
194 > molecules.\cite{Lacman65} The central ion sees what is effectively a
195 > set of octupoles at large distances. These facts suggest that the
196 > Madelung constants are relatively short ranged for perfect ionic
197 > crystals.\cite{Wolf:1999dn}
198  
199   One can make a similar argument for crystals of point multipoles. The
200   Luttinger and Tisza treatment of energy constants for dipolar lattices
# Line 188 | Line 212 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
212   unstable.
213  
214   In ionic crystals, real-space truncation can break the effective
215 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
216 < swings in the electrostatic energy as the cutoff radius is increased
217 < (or as individual ions move back and forth across the boundary).  This
218 < is why the image charges were necessary for the Wolf sum to exhibit
219 < rapid convergence.  Similarly, the real-space truncation of point
220 < multipole interactions breaks higher order multipole arrangements, and
221 < image multipoles are required for real-space treatments of
222 < electrostatic energies.
215 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
216 > significant swings in the electrostatic energy as individual ions move
217 > back and forth across the boundary.  This is why the image charges are
218 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
219 > the real-space truncation of point multipole interactions breaks
220 > higher order multipole arrangements, and image multipoles are required
221 > for real-space treatments of electrostatic energies.
222 >
223 > The shorter effective range of electrostatic interactions is not
224 > limited to perfect crystals, but can also apply in disordered fluids.
225 > Even at elevated temperatures, there is, on average, local charge
226 > balance in an ionic liquid, where each positive ion has surroundings
227 > dominated by negaitve ions and vice versa.  The reversed-charge images
228 > on the cutoff sphere that are integral to the Wolf and DSF approaches
229 > retain the effective multipolar interactions as the charges traverse
230 > the cutoff boundary.
231  
232 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
233 + significant orientational averaging that additionally reduces the
234 + effect of long-range multipolar interactions.  The image multipoles
235 + that are introduced in the TSF, GSF, and SP methods mimic this effect
236 + and reduce the effective range of the multipolar interactions as
237 + interacting molecules traverse each other's cutoff boundaries.
238 +
239   % Because of this reason, although the nature of electrostatic
240   % interaction short ranged, the hard cutoff sphere creates very large
241   % fluctuation in the electrostatic energy for the perfect crystal. In
# Line 253 | Line 292 | multipoles up to octupolar
292   rough approximation.  Atomic sites can also be assigned point
293   multipoles and polarizabilities to increase the accuracy of the
294   molecular model.  Recently, water has been modeled with point
295 < multipoles up to octupolar
296 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
295 > multipoles up to octupolar order using the soft sticky
296 > dipole-quadrupole-octupole (SSDQO)
297 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
298   multipoles up to quadrupolar order have also been coupled with point
299   polarizabilities in the high-quality AMOEBA and iAMOEBA water
300 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
300 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
301   using point multipole with the real space truncation without
302   accounting for multipolar neutrality will create energy conservation
303   issues in molecular dynamics (MD) simulations.
# Line 297 | Line 337 | where the multipole operator for site $\bf a$,
337   \begin{equation}
338   U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
339   \end{equation}
340 < where the multipole operator for site $\bf a$,
341 < \begin{equation}
342 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
343 < +  Q_{{\bf a}\alpha\beta}
304 < \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
305 < \end{equation}
306 < is expressed in terms of the point charge, $C_{\bf a}$, dipole,
307 < $D_{{\bf a}\alpha}$, and quadrupole, $Q_{{\bf a}\alpha\beta}$, for
308 < object $\bf a$.  Note that in this work, we use the primitive
309 < quadrupole tensor, $Q_{a\alpha,\beta}=\frac{1}{2}\sum_{k\;in\;a}q_k
310 < r_{k\alpha}r_{k\beta}$ to represent point quadrupoles on a site.
340 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
341 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
342 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
343 > $\bf a$, etc.
344  
345 < Interactions between multipoles can be expressed as higher derivatives
346 < of the bare Coulomb potential, so one way of ensuring that the forces
347 < and torques vanish at the cutoff distance is to include a larger
348 < number of terms in the truncated Taylor expansion, e.g.,
349 < %
350 < \begin{equation}
351 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
352 < \end{equation}
353 < %
354 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
355 < Thus, for $f(r)=1/r$, we find
356 < %
357 < \begin{equation}
358 < f_1(r)=\frac{1}{r}- \frac{1}{R_c} + (r - R_c) \frac{1}{R_c^2} - \frac{(r-R_c)^2}{R_c^3} .
359 < \end{equation}
360 < This function is an approximate electrostatic potential that has
361 < vanishing second derivatives at the cutoff radius, making it suitable
362 < for shifting the forces and torques of charge-dipole interactions.
345 > % Interactions between multipoles can be expressed as higher derivatives
346 > % of the bare Coulomb potential, so one way of ensuring that the forces
347 > % and torques vanish at the cutoff distance is to include a larger
348 > % number of terms in the truncated Taylor expansion, e.g.,
349 > % %
350 > % \begin{equation}
351 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
352 > % \end{equation}
353 > % %
354 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
355 > % Thus, for $f(r)=1/r$, we find
356 > % %
357 > % \begin{equation}
358 > % f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
359 > % \end{equation}
360 > % This function is an approximate electrostatic potential that has
361 > % vanishing second derivatives at the cutoff radius, making it suitable
362 > % for shifting the forces and torques of charge-dipole interactions.
363  
364 < In general, the TSF potential for any multipole-multipole interaction
365 < can be written
364 > The TSF potential for any multipole-multipole interaction can be
365 > written
366   \begin{equation}
367   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
368   \label{generic}
369   \end{equation}
370 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
371 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
372 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
373 < energy, force, and torques, the required number of terms from Taylor
374 < series expansion in $f_n(r)$ must be performed for different
375 < multipole-multipole interactions.
370 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
371 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
372 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
373 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
374 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
375 > force, and torques, a Taylor expansion with $n$ terms must be
376 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
377  
378 < To carry out the same procedure for a damped electrostatic kernel, we
379 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
380 < Many of the derivatives of the damped kernel are well known from
381 < Smith's early work on multipoles for the Ewald
382 < summation.\cite{Smith82,Smith98}
378 > % To carry out the same procedure for a damped electrostatic kernel, we
379 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
380 > % Many of the derivatives of the damped kernel are well known from
381 > % Smith's early work on multipoles for the Ewald
382 > % summation.\cite{Smith82,Smith98}
383  
384 < Note that increasing the value of $n$ will add additional terms to the
385 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
386 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
387 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
388 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
389 < for computing multipole energies, forces, and torques, and smooth
390 < cutoffs of these quantities can be guaranteed as long as the number of
391 < terms in the Taylor series exceeds the derivative order required.
384 > % Note that increasing the value of $n$ will add additional terms to the
385 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
386 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
387 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
388 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
389 > % for computing multipole energies, forces, and torques, and smooth
390 > % cutoffs of these quantities can be guaranteed as long as the number of
391 > % terms in the Taylor series exceeds the derivative order required.
392  
393   For multipole-multipole interactions, following this procedure results
394 < in separate radial functions for each distinct orientational
395 < contribution to the potential, and ensures that the forces and torques
396 < from {\it each} of these contributions will vanish at the cutoff
397 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
398 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
394 > in separate radial functions for each of the distinct orientational
395 > contributions to the potential, and ensures that the forces and
396 > torques from each of these contributions will vanish at the cutoff
397 > radius.  For example, the direct dipole dot product
398 > ($\mathbf{D}_{\bf a}
399 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
400   dot products:
401   \begin{equation}
402 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
403 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
404 < -\frac{1}{4\pi \epsilon_0}
405 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
406 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
402 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
403 >  \mathbf{D}_{\bf a} \cdot
404 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
405 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
406 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
407   \end{equation}
408  
409 < The electrostatic forces and torques acting on the central multipole
410 < site due to another site within cutoff sphere are derived from
409 > For the Taylor shifted (TSF) method with the undamped kernel,
410 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
411 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
412 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
413 > connection to unmodified electrostatics as well as the smooth
414 > transition to zero in both these functions as $r\rightarrow r_c$.  The
415 > electrostatic forces and torques acting on the central multipole due
416 > to another site within the cutoff sphere are derived from
417   Eq.~\ref{generic}, accounting for the appropriate number of
418   derivatives. Complete energy, force, and torque expressions are
419   presented in the first paper in this series (Reference
420 < \citep{PaperI}).
420 > \onlinecite{PaperI}).
421  
422   \subsection{Gradient-shifted force (GSF)}
423  
424 < A second (and significantly simpler) method involves shifting the
425 < gradient of the raw coulomb potential for each particular multipole
424 > A second (and conceptually simpler) method involves shifting the
425 > gradient of the raw Coulomb potential for each particular multipole
426   order.  For example, the raw dipole-dipole potential energy may be
427   shifted smoothly by finding the gradient for two interacting dipoles
428   which have been projected onto the surface of the cutoff sphere
429   without changing their relative orientation,
430 < \begin{displaymath}
431 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
432 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
433 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
434 < \end{displaymath}
435 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
436 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
437 < the signs are reversed for the dipole that has been projected onto the
438 < cutoff sphere).  In many ways, this simpler approach is closer in
439 < spirit to the original shifted force method, in that it projects a
440 < neutralizing multipole (and the resulting forces from this multipole)
441 < onto a cutoff sphere. The resulting functional forms for the
442 < potentials, forces, and torques turn out to be quite similar in form
443 < to the Taylor-shifted approach, although the radial contributions are
444 < significantly less perturbed by the Gradient-shifted approach than
445 < they are in the Taylor-shifted method.
430 > \begin{equation}
431 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
432 > U_{D_{\bf a} D_{\bf b}}(r_c)
433 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
434 >  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
435 > \end{equation}
436 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
437 >  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
438 > (although the signs are reversed for the dipole that has been
439 > projected onto the cutoff sphere).  In many ways, this simpler
440 > approach is closer in spirit to the original shifted force method, in
441 > that it projects a neutralizing multipole (and the resulting forces
442 > from this multipole) onto a cutoff sphere. The resulting functional
443 > forms for the potentials, forces, and torques turn out to be quite
444 > similar in form to the Taylor-shifted approach, although the radial
445 > contributions are significantly less perturbed by the gradient-shifted
446 > approach than they are in the Taylor-shifted method.
447  
448 + For the gradient shifted (GSF) method with the undamped kernel,
449 + $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
450 + $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
451 + Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
452 + because the Taylor expansion retains only one term, they are
453 + significantly less perturbed than the TSF functions.
454 +
455   In general, the gradient shifted potential between a central multipole
456   and any multipolar site inside the cutoff radius is given by,
457   \begin{equation}
458 < U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
459 < U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
460 < \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert  _{r_c} \right]
458 >  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
459 >    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
460 >    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
461   \label{generic2}
462   \end{equation}
463   where the sum describes a separate force-shifting that is applied to
464 < each orientational contribution to the energy.
464 > each orientational contribution to the energy.  In this expression,
465 > $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
466 > ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
467 > represent the orientations the multipoles.
468  
469   The third term converges more rapidly than the first two terms as a
470   function of radius, hence the contribution of the third term is very
471   small for large cutoff radii.  The force and torque derived from
472 < equation \ref{generic2} are consistent with the energy expression and
473 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
472 > Eq. \ref{generic2} are consistent with the energy expression and
473 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
474   can be considered generalizations of the original DSF method for
475   higher order multipole interactions. GSF and TSF are also identical up
476   to the charge-dipole interaction but generate different expressions in
477   the energy, force and torque for higher order multipole-multipole
478   interactions. Complete energy, force, and torque expressions for the
479   GSF potential are presented in the first paper in this series
480 < (Reference \citep{PaperI})
480 > (Reference~\onlinecite{PaperI}).
481  
482  
483   \subsection{Shifted potential (SP) }
# Line 439 | Line 490 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
490   interactions with the central multipole and the image. This
491   effectively shifts the total potential to zero at the cutoff radius,
492   \begin{equation}
493 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
493 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
494 > U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
495   \label{eq:SP}
496   \end{equation}          
497   where the sum describes separate potential shifting that is done for
498   each orientational contribution to the energy (e.g. the direct dipole
499   product contribution is shifted {\it separately} from the
500   dipole-distance terms in dipole-dipole interactions).  Note that this
501 < is not a simple shifting of the total potential at $R_c$. Each radial
501 > is not a simple shifting of the total potential at $r_c$. Each radial
502   contribution is shifted separately.  One consequence of this is that
503   multipoles that reorient after leaving the cutoff sphere can re-enter
504   the cutoff sphere without perturbing the total energy.
505  
506 < The potential energy between a central multipole and other multipolar
507 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
508 < force and torque obtained from the shifted potential (SP) are
509 < discontinuous at $R_c$. Therefore, MD simulations will still
510 < experience energy drift while operating under the SP potential, but it
511 < may be suitable for Monte Carlo approaches where the configurational
512 < energy differences are the primary quantity of interest.
506 > For the shifted potential (SP) method with the undamped kernel,
507 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
508 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
509 > central multipole and other multipolar sites goes smoothly to zero as
510 > $r \rightarrow r_c$.  However, the force and torque obtained from the
511 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
512 > will still experience energy drift while operating under the SP
513 > potential, but it may be suitable for Monte Carlo approaches where the
514 > configurational energy differences are the primary quantity of
515 > interest.
516  
517 < \subsection{The Self term}
517 > \subsection{The Self Term}
518   In the TSF, GSF, and SP methods, a self-interaction is retained for
519   the central multipole interacting with its own image on the surface of
520   the cutoff sphere.  This self interaction is nearly identical with the
521   self-terms that arise in the Ewald sum for multipoles.  Complete
522   expressions for the self terms are presented in the first paper in
523 < this series (Reference \citep{PaperI})  
523 > this series (Reference \onlinecite{PaperI}).
524  
525  
526   \section{\label{sec:methodology}Methodology}
# Line 477 | Line 532 | arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} Thi
532   real-space cutoffs.  In the first paper of this series, we compared
533   the dipolar and quadrupolar energy expressions against analytic
534   expressions for ordered dipolar and quadrupolar
535 < arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
536 < multipolar Ewald sum as a reference method for comparing energies,
537 < forces, and torques for molecular models that mimic disordered and
538 < ordered condensed-phase systems.  These test-cases include:
535 > arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
536 > used the multipolar Ewald sum as a reference method for comparing
537 > energies, forces, and torques for molecular models that mimic
538 > disordered and ordered condensed-phase systems.  The parameters used
539 > in the test cases are given in table~\ref{tab:pars}.
540  
541 < \begin{itemize}
542 < \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
543 < \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
544 < \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
545 < \item Soft Quadrupolar solids  ($\sigma = , \epsilon = , Q_{xx} = ...$)
546 < \item A mixed multipole model for water
547 < \item A mixed multipole models for water with dissolved ions
548 < \end{itemize}
549 < This last test case exercises all levels of the multipole-multipole
550 < interactions we have derived so far and represents the most complete
551 < test of the new methods.
541 > \begin{table}
542 > \label{tab:pars}
543 > \caption{The parameters used in the systems used to evaluate the new
544 >  real-space methods.  The most comprehensive test was a liquid
545 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
546 >  ions).  This test excercises all orders of the multipolar
547 >  interactions developed in the first paper.}
548 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
549 >             & \multicolumn{2}{c|}{LJ parameters} &
550 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
551 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
552 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
553 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
554 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
555 > \AA\textsuperscript{2})} \\ \hline
556 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
557 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
558 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
559 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
560 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
561 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
562 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
563 > \end{tabularx}
564 > \end{table}
565 > The systems consist of pure multipolar solids (both dipole and
566 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
567 > fluid composed of sites containing both dipoles and quadrupoles
568 > simultaneously, and a final test case that includes ions with point
569 > charges in addition to the multipolar fluid.  The solid-phase
570 > parameters were chosen so that the systems can explore some
571 > orientational freedom for the multipolar sites, while maintaining
572 > relatively strict translational order.  The SSDQ model used here is
573 > not a particularly accurate water model, but it does test
574 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
575 > interactions at roughly the same magnitudes. The last test case, SSDQ
576 > water with dissolved ions, exercises \textit{all} levels of the
577 > multipole-multipole interactions we have derived so far and represents
578 > the most complete test of the new methods.
579  
580   In the following section, we present results for the total
581   electrostatic energy, as well as the electrostatic contributions to
582   the force and torque on each molecule.  These quantities have been
583   computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
584 < and have been compared with the values obtaine from the multipolar
585 < Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
584 > and have been compared with the values obtained from the multipolar
585 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
586   between two configurations is the primary quantity that governs how
587   the simulation proceeds. These differences are the most imporant
588   indicators of the reliability of a method even if the absolute
# Line 510 | Line 593 | contributions to the forces and torques.
593   behavior of the simulation, so we also compute the electrostatic
594   contributions to the forces and torques.
595  
596 < \subsection{Model systems}
597 < To sample independent configurations of multipolar crystals, a body
598 < centered cubic (BCC) crystal which is a minimum energy structure for
599 < point dipoles was generated using 3,456 molecules.  The multipoles
600 < were translationally locked in their respective crystal sites for
601 < equilibration at a relatively low temperature (50K), so that dipoles
602 < or quadrupoles could freely explore all accessible orientations.  The
603 < translational constraints were removed, and the crystals were
521 < simulated for 10 ps in the microcanonical (NVE) ensemble with an
522 < average temperature of 50 K.  Configurations were sampled at equal
523 < time intervals for the comparison of the configurational energy
524 < differences.  The crystals were not simulated close to the melting
525 < points in order to avoid translational deformation away of the ideal
526 < lattice geometry.
596 > \subsection{Implementation}
597 > The real-space methods developed in the first paper in this series
598 > have been implemented in our group's open source molecular simulation
599 > program, OpenMD,\cite{openmd} which was used for all calculations in
600 > this work.  The complementary error function can be a relatively slow
601 > function on some processors, so all of the radial functions are
602 > precomputed on a fine grid and are spline-interpolated to provide
603 > values when required.  
604  
605 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
606 < system was created with 2048 molecules oriented randomly.  These were
605 > Using the same simulation code, we compare to a multipolar Ewald sum
606 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
607 > the Ewald sum is a re-implementation of the algorithm originally
608 > proposed by Smith that does not use the particle mesh or smoothing
609 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
610 > being compared are the electrostatic contributions to energies, force,
611 > and torques.  All other contributions to these quantities (i.e. from
612 > Lennard-Jones interactions) are removed prior to the comparisons.
613  
614 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
615 < temperature after equilibration.  We collected 250 different
616 < configurations in equal interval of time. For the ions mixed liquid
617 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
618 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
619 < at the same environment for 1ns and 250 configurations were
620 < collected. While comparing energies, forces, and torques with Ewald
621 < method, Lennard-Jones potentials were turned off and purely
622 < electrostatic interaction had been compared.
614 > The convergence parameter ($\alpha$) also plays a role in the balance
615 > of the real-space and reciprocal-space portions of the Ewald
616 > calculation.  Typical molecular mechanics packages set this to a value
617 > that depends on the cutoff radius and a tolerance (typically less than
618 > $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
619 > associated with increasing accuracy at the expense of computational
620 > time spent on the reciprocal-space portion of the
621 > summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
622 > 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
623 > Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
624  
625 + The real-space models have self-interactions that provide
626 + contributions to the energies only.  Although the self interaction is
627 + a rapid calculation, we note that in systems with fluctuating charges
628 + or point polarizabilities, the self-term is not static and must be
629 + recomputed at each time step.
630 +
631 + \subsection{Model systems}
632 + To sample independent configurations of the multipolar crystals, body
633 + centered cubic (bcc) crystals, which exhibit the minimum energy
634 + structures for point dipoles, were generated using 3,456 molecules.
635 + The multipoles were translationally locked in their respective crystal
636 + sites for equilibration at a relatively low temperature (50K) so that
637 + dipoles or quadrupoles could freely explore all accessible
638 + orientations.  The translational constraints were then removed, the
639 + systems were re-equilibrated, and the crystals were simulated for an
640 + additional 10 ps in the microcanonical (NVE) ensemble with an average
641 + temperature of 50 K.  The balance between moments of inertia and
642 + particle mass were chosen to allow orientational sampling without
643 + significant translational motion.  Configurations were sampled at
644 + equal time intervals in order to compare configurational energy
645 + differences.  The crystals were simulated far from the melting point
646 + in order to avoid translational deformation away of the ideal lattice
647 + geometry.
648 +
649 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
650 + simulations, each system was created with 2,048 randomly-oriented
651 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
652 + Each system was then simulated for 1 ns in the microcanonical (NVE)
653 + ensemble.  We collected 250 different configurations at equal time
654 + intervals. For the liquid system that included ionic species, we
655 + converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
656 + \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
657 + run under the same conditions for 1 ns. A total of 250 configurations
658 + were collected. In the following comparisons of energies, forces, and
659 + torques, the Lennard-Jones potentials were turned off and only the
660 + purely electrostatic quantities were compared with the same values
661 + obtained via the Ewald sum.
662 +
663   \subsection{Accuracy of Energy Differences, Forces and Torques}
664   The pairwise summation techniques (outlined above) were evaluated for
665   use in MC simulations by studying the energy differences between
# Line 550 | Line 672 | we used least square regressions analysiss for the six
672   should be identical for all methods.
673  
674   Since none of the real-space methods provide exact energy differences,
675 < we used least square regressions analysiss for the six different
675 > we used least square regressions analysis for the six different
676   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
677   with the multipolar Ewald reference method.  Unitary results for both
678   the correlation (slope) and correlation coefficient for these
# Line 561 | Line 683 | also been compared by using least squares regression a
683   configurations and 250 configurations were recorded for comparison.
684   Each system provided 31,125 energy differences for a total of 186,750
685   data points.  Similarly, the magnitudes of the forces and torques have
686 < also been compared by using least squares regression analyses. In the
686 > also been compared using least squares regression analysis. In the
687   forces and torques comparison, the magnitudes of the forces acting in
688   each molecule for each configuration were evaluated. For example, our
689   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 647 | Line 769 | model must allow for long simulation times with minima
769        
770   %        \label{fig:barGraph2}
771   %      \end{figure}
772 < %The correlation coefficient ($R^2$) and slope of the linear regression plots for the energy differences for all six different molecular systems is shown in figure 4a and 4b.The plot shows that the correlation coefficient improves for the SP cutoff method as compared to the undamped hard cutoff method in the case of SSDQC, SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar crystal and liquid, the correlation coefficient is almost unchanged and close to 1.  The correlation coefficient is smallest (0.696276 for $r_c$ = 9 $A^o$) for the SSDQC liquid because of the presence of charge-charge and charge-multipole interactions. Since the charge-charge and charge-multipole interaction is long ranged, there is huge deviation of correlation coefficient from 1. Similarly, the quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with compared to interactions in the other multipolar systems, thus the correlation coefficient very close to 1 even for hard cutoff method. The idea of placing image multipole on the surface of the cutoff sphere improves the correlation coefficient and makes it close to 1 for all types of multipolar systems. Similarly the slope is hugely deviated from the correct value for the lower order multipole-multipole interaction and slightly deviated for higher order multipole – multipole interaction. The SP method improves both correlation coefficient ($R^2$) and slope significantly in SSDQC and dipolar systems.  The Slope is found to be deviated more in dipolar crystal as compared to liquid which is associated with the large fluctuation in the electrostatic energy in crystal. The GSF also produced better values of correlation coefficient and slope with the proper selection of the damping alpha (Interested reader can consult accompanying supporting material). The TSF method gives good value of correlation coefficient for the dipolar crystal, dipolar liquid, SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the regression slopes are significantly deviated.
772 > %The correlation coefficient ($R^2$) and slope of the linear
773 > %regression plots for the energy differences for all six different
774 > %molecular systems is shown in figure 4a and 4b.The plot shows that
775 > %the correlation coefficient improves for the SP cutoff method as
776 > %compared to the undamped hard cutoff method in the case of SSDQC,
777 > %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
778 > %crystal and liquid, the correlation coefficient is almost unchanged
779 > %and close to 1.  The correlation coefficient is smallest (0.696276
780 > %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
781 > %charge-charge and charge-multipole interactions. Since the
782 > %charge-charge and charge-multipole interaction is long ranged, there
783 > %is huge deviation of correlation coefficient from 1. Similarly, the
784 > %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
785 > %compared to interactions in the other multipolar systems, thus the
786 > %correlation coefficient very close to 1 even for hard cutoff
787 > %method. The idea of placing image multipole on the surface of the
788 > %cutoff sphere improves the correlation coefficient and makes it close
789 > %to 1 for all types of multipolar systems. Similarly the slope is
790 > %hugely deviated from the correct value for the lower order
791 > %multipole-multipole interaction and slightly deviated for higher
792 > %order multipole – multipole interaction. The SP method improves both
793 > %correlation coefficient ($R^2$) and slope significantly in SSDQC and
794 > %dipolar systems.  The Slope is found to be deviated more in dipolar
795 > %crystal as compared to liquid which is associated with the large
796 > %fluctuation in the electrostatic energy in crystal. The GSF also
797 > %produced better values of correlation coefficient and slope with the
798 > %proper selection of the damping alpha (Interested reader can consult
799 > %accompanying supporting material). The TSF method gives good value of
800 > %correlation coefficient for the dipolar crystal, dipolar liquid,
801 > %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
802 > %regression slopes are significantly deviated.
803 >
804   \begin{figure}
805 <        \centering
806 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
807 <        \caption{The correlation coefficient and regression slope of configurational energy differences for a given method with compared with the reference Ewald method. The value of result equal to 1(dashed line) indicates energy difference is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle)}
808 <        \label{fig:slopeCorr_energy}
809 <    \end{figure}
810 < The combined correlation coefficient and slope for all six systems is shown in Figure ~\ref{fig:slopeCorr_energy}. The correlation coefficient for the undamped hard cutoff method is does not have good agreement with the Ewald because of the fluctuation of the electrostatic energy in the direct truncation method. This deviation in correlation coefficient is improved by using SP, GSF, and TSF method. But the TSF method worsens the regression slope stating that this method produces statistically more biased result as compared to Ewald. Also the GSF method slightly deviate slope but it can be alleviated by using proper value of damping alpha and cutoff radius. The SP method shows good agreement with Ewald method for all values of damping alpha and radii.
805 >  \centering
806 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
807 >  \caption{Statistical analysis of the quality of configurational
808 >    energy differences for the real-space electrostatic methods
809 >    compared with the reference Ewald sum.  Results with a value equal
810 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
811 >    from those obtained using the multipolar Ewald sum.  Different
812 >    values of the cutoff radius are indicated with different symbols
813 >    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
814 >    triangles).}
815 >  \label{fig:slopeCorr_energy}
816 > \end{figure}
817 >
818 > The combined correlation coefficient and slope for all six systems is
819 > shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
820 > reproduce the Ewald configurational energy differences with remarkable
821 > fidelity.  Undamped hard cutoffs introduce a significant amount of
822 > random scatter in the energy differences which is apparent in the
823 > reduced value of the correlation coefficient for this method.  This
824 > can be easily understood as configurations which exhibit small
825 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
826 > will see large energy jumps when hard cutoffs are used.  The
827 > orientations of the multipoles (particularly in the ordered crystals)
828 > mean that these energy jumps can go in either direction, producing a
829 > significant amount of random scatter, but no systematic error.
830 >
831 > The TSF method produces energy differences that are highly correlated
832 > with the Ewald results, but it also introduces a significant
833 > systematic bias in the values of the energies, particularly for
834 > smaller cutoff values. The TSF method alters the distance dependence
835 > of different orientational contributions to the energy in a
836 > non-uniform way, so the size of the cutoff sphere can have a large
837 > effect, particularly for the crystalline systems.
838 >
839 > Both the SP and GSF methods appear to reproduce the Ewald results with
840 > excellent fidelity, particularly for moderate damping ($\alpha =
841 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
842 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
843 > method with short cutoffs, all of the methods would be appropriate for
844 > use in Monte Carlo simulations.
845 >
846   \subsection{Magnitude of the force and torque vectors}
847 < The comparison of the magnitude of the combined forces and torques for the data accumulated from all system types are shown in Figure ~\ref{fig:slopeCorr_force}. The correlation and slope for the forces agree with the Ewald even for the hard cutoff method. For the system of molecules with higher order multipoles, the interaction is short ranged. Moreover, the force decays more rapidly than the electrostatic energy hence the hard cutoff method also produces good results. Although the pure cutoff gives the good match of the electrostatic force, the discontinuity in the force at the cutoff radius causes problem in the total energy conservation in MD simulations, which will be discussed in detail in subsection D. The correlation coefficient for GSF method also perfectly matches with Ewald but the slope is slightly deviated (due to extra term obtained from the angular differentiation). This deviation in the slope can be alleviated with proper selection of the damping alpha and radii ($\alpha = 0.2$ and $r_c = 12 A^o$ are good choice). The TSF method shows good agreement in the correlation coefficient but the slope is not good as compared to the Ewald.
847 >
848 > The comparisons of the magnitudes of the forces and torques for the
849 > data accumulated from all six systems are shown in Figures
850 > ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
851 > correlation and slope for the forces agree well with the Ewald sum
852 > even for the hard cutoffs.
853 >
854 > For systems of molecules with only multipolar interactions, the pair
855 > energy contributions are quite short ranged.  Moreover, the force
856 > decays more rapidly than the electrostatic energy, hence the hard
857 > cutoff method can also produce reasonable agreement for this quantity.
858 > Although the pure cutoff gives reasonably good electrostatic forces
859 > for pairs of molecules included within each other's cutoff spheres,
860 > the discontinuity in the force at the cutoff radius can potentially
861 > cause energy conservation problems as molecules enter and leave the
862 > cutoff spheres.  This is discussed in detail in section
863 > \ref{sec:conservation}.
864 >
865 > The two shifted-force methods (GSF and TSF) exhibit a small amount of
866 > systematic variation and scatter compared with the Ewald forces.  The
867 > shifted-force models intentionally perturb the forces between pairs of
868 > molecules inside each other's cutoff spheres in order to correct the
869 > energy conservation issues, and this perturbation is evident in the
870 > statistics accumulated for the molecular forces.  The GSF
871 > perturbations are minimal, particularly for moderate damping and
872 > commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
873 > reasonable agreement in the correlation coefficient but again the
874 > systematic error in the forces is concerning if replication of Ewald
875 > forces is desired.
876 >
877   \begin{figure}
878 <        \centering
879 <        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
880 <        \caption{The correlation coefficient and regression slope of the magnitude of the force for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\  = circle, 12 \AA\  = square 15 \AA\  = inverted triangle). }
881 <        \label{fig:slopeCorr_force}
882 <    \end{figure}
883 < The torques appears to be very influenced because of extra term generated when the potential energy is modified to get consistent force and torque.  The result shows that the torque from the hard cutoff method has good agreement with Ewald. As the potential is modified to make it consistent with the force and torque, the correlation and slope is deviated as shown in Figure~\ref{fig:slopeCorr_torque} for SP, GSF and TSF cutoff methods.  But the proper value of the damping alpha and radius can improve the agreement of the GSF with the Ewald method. The TSF method shows worst agreement in the slope as compared to Ewald even for larger cutoff radii.
878 >  \centering
879 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
880 >  \caption{Statistical analysis of the quality of the force vector
881 >    magnitudes for the real-space electrostatic methods compared with
882 >    the reference Ewald sum. Results with a value equal to 1 (dashed
883 >    line) indicate force magnitude values indistinguishable from those
884 >    obtained using the multipolar Ewald sum.  Different values of the
885 >    cutoff radius are indicated with different symbols (9\AA\ =
886 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
887 >  \label{fig:slopeCorr_force}
888 > \end{figure}
889 >
890 >
891   \begin{figure}
892 <        \centering
893 <        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
894 <        \caption{The correlation coefficient and regression slope of the magnitude of the torque for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle).}
895 <        \label{fig:slopeCorr_torque}
896 <    \end{figure}
892 >  \centering
893 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
894 >  \caption{Statistical analysis of the quality of the torque vector
895 >    magnitudes for the real-space electrostatic methods compared with
896 >    the reference Ewald sum. Results with a value equal to 1 (dashed
897 >    line) indicate force magnitude values indistinguishable from those
898 >    obtained using the multipolar Ewald sum.  Different values of the
899 >    cutoff radius are indicated with different symbols (9\AA\ =
900 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
901 >  \label{fig:slopeCorr_torque}
902 > \end{figure}
903 >
904 > The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
905 > significantly influenced by the choice of real-space method.  The
906 > torque expressions have the same distance dependence as the energies,
907 > which are naturally longer-ranged expressions than the inter-site
908 > forces.  Torques are also quite sensitive to orientations of
909 > neighboring molecules, even those that are near the cutoff distance.
910 >
911 > The results shows that the torque from the hard cutoff method
912 > reproduces the torques in quite good agreement with the Ewald sum.
913 > The other real-space methods can cause some deviations, but excellent
914 > agreement with the Ewald sum torques is recovered at moderate values
915 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
916 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
917 > in the slope when compared with the Ewald torques even for larger
918 > cutoff radii.  It appears that the severity of the perturbations in
919 > the TSF method are most in evidence for the torques.
920 >
921   \subsection{Directionality of the force and torque vectors}  
922 < The accurate evaluation of the direction of the force and torques are also important for the dynamic simulation.In our research, the direction data sets were computed from the purposed method and compared with Ewald using Fisher statistics and results are expressed in terms of circular variance ($Var(\theta$).The force and torque vectors from the purposed method followed Fisher probability distribution function expressed in equation~\ref{eq:pdf}. The circular variance for the force and torque vectors of each molecule in the 250 configurations for all system types is shown in Figure~\ref{fig:slopeCorr_circularVariance}. The direction of the force and torque vectors from hard and SP cutoff methods showed best directional agreement with the Ewald. The force and torque vectors from GSF method also showed good agreement with the Ewald method, which can also be improved by varying damping alpha and cutoff radius.For $\alpha = 0.2$ and $r_c = 12 A^o$, $ Var(\theta) $ for direction of the force was found to be 0.002061 and corresponding value of $\kappa $ was 485.20. Integration of equation ~\ref{eq:pdf} for that corresponding value of $\kappa$ showed that 95\% of force vectors are with in $6.37^o$. The TSF method is the poorest in evaluating accurate direction with compared to Hard, SP, and GSF methods. The circular variance for the direction of the torques is larger as compared to force. For same $\alpha = 0.2, r_c = 12 A^o$ and GSF method, the circular variance was 0.01415, which showed 95\% of torque vectors are within $16.75^o$.The direction of the force and torque vectors can be improved by varying $\alpha$ and $r_c$.
922 >
923 > The accurate evaluation of force and torque directions is just as
924 > important for molecular dynamics simulations as the magnitudes of
925 > these quantities. Force and torque vectors for all six systems were
926 > analyzed using Fisher statistics, and the quality of the vector
927 > directionality is shown in terms of circular variance
928 > ($\mathrm{Var}(\theta)$) in figure
929 > \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
930 > from the new real-space methods exhibit nearly-ideal Fisher probability
931 > distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
932 > exhibit the best vectorial agreement with the Ewald sum. The force and
933 > torque vectors from GSF method also show good agreement with the Ewald
934 > method, which can also be systematically improved by using moderate
935 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
936 > 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
937 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
938 > the corresponding Ewald forces. The TSF method produces the poorest
939 > agreement with the Ewald force directions.
940 >
941 > Torques are again more perturbed than the forces by the new real-space
942 > methods, but even here the variance is reasonably small.  For the same
943 > method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
944 > the circular variance was 0.01415, corresponds to a distribution which
945 > has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
946 > results. Again, the direction of the force and torque vectors can be
947 > systematically improved by varying $\alpha$ and $r_c$.
948  
949   \begin{figure}
950 <        \centering
951 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
952 <        \caption{The circular variance of the data sets of the
953 <          direction of the  force and torque vectors obtained from a
954 <          given method about reference Ewald method. The result equal
955 <          to 0 (dashed line) indicates direction of the vectors are
956 <          indistinguishable from the Ewald method. Here different
957 <          symbols represent different value of the cutoff radius (9
958 <          \AA\ = circle, 12 \AA\ = square, 15 \AA\  = inverted triangle)}
959 <        \label{fig:slopeCorr_circularVariance}
960 <    \end{figure}
688 < \subsection{Total energy conservation}
689 < We have tested the conservation of energy in the SSDQC liquid system
690 < by running system for 1ns in the Hard, SP, GSF and TSF method. The
691 < Hard cutoff method shows very high energy drifts 433.53
692 < KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol
693 < (measured by the SD from the slope) for the undamped case, which makes
694 < it completely unusable in MD simulations. The SP method also shows
695 < large value of energy drift 1.289 Kcal/Mol/ns/particle and energy
696 < fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method
697 < is due to the non-vanishing nature of the torque and force at the
698 < cutoff radius. We can improve the energy conservation in some extent
699 < by the proper selection of the damping alpha but the improvement is
700 < not good enough, which can be observed in Figure 9a and 9b .The GSF
701 < and TSF shows very low value of energy drift 0.09016, 0.07371
702 < KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol
703 < respectively for the undamped case. Since the absolute value of the
704 < evaluated electrostatic energy, force and torque from TSF method are
705 < deviated from the Ewald, it does not mimic MD simulations
706 < appropriately. The electrostatic energy, force and torque from the GSF
707 < method have very good agreement with the Ewald. In addition, the
708 < energy drift and energy fluctuation from the GSF method is much better
709 < than Ewald’s method for reciprocal space vector value ($k_f$) equal to
710 < 7 as shown in Figure~\ref{fig:energyDrift} and
711 < ~\ref{fig:fluctuation}. We can improve the total energy fluctuation
712 < and drift for the Ewald’s method by increasing size of the reciprocal
713 < space, which extremely increseses the simulation time. In our current
714 < simulation, the simulation time for the Hard, SP, and GSF methods are
715 < about 5.5 times faster than the Ewald method.
950 >  \centering
951 >  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
952 >  \caption{The circular variance of the direction of the force and
953 >    torque vectors obtained from the real-space methods around the
954 >    reference Ewald vectors. A variance equal to 0 (dashed line)
955 >    indicates direction of the force or torque vectors are
956 >    indistinguishable from those obtained from the Ewald sum. Here
957 >    different symbols represent different values of the cutoff radius
958 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
959 >  \label{fig:slopeCorr_circularVariance}
960 > \end{figure}
961  
962 < In Fig.~\ref{fig:energyDrift}, $\delta \mbox{E}_1$ is a measure of the
718 < linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
719 < over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
720 < standard deviation of the energy fluctuations in units of $\mbox{kcal
721 <  mol}^{-1}$ per particle. In the bottom plot, it is apparent that the
722 < energy drift is reduced by a significant amount (2 to 3 orders of
723 < magnitude improvement at all values of the damping coefficient) by
724 < chosing either of the shifted-force methods over the hard or SP
725 < methods.  We note that the two shifted-force method can give
726 < significantly better energy conservation than the multipolar Ewald sum
727 < with the same choice of real-space cutoffs.
962 > \subsection{Energy conservation\label{sec:conservation}}
963  
964 + We have tested the conservation of energy one can expect to see with
965 + the new real-space methods using the SSDQ water model with a small
966 + fraction of solvated ions. This is a test system which exercises all
967 + orders of multipole-multipole interactions derived in the first paper
968 + in this series and provides the most comprehensive test of the new
969 + methods.  A liquid-phase system was created with 2000 water molecules
970 + and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
971 + temperature of 300K.  After equilibration, this liquid-phase system
972 + was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
973 + a cutoff radius of 12\AA.  The value of the damping coefficient was
974 + also varied from the undamped case ($\alpha = 0$) to a heavily damped
975 + case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
976 + sample was also run using the multipolar Ewald sum with the same
977 + real-space cutoff.
978 +
979 + In figure~\ref{fig:energyDrift} we show the both the linear drift in
980 + energy over time, $\delta E_1$, and the standard deviation of energy
981 + fluctuations around this drift $\delta E_0$.  Both of the
982 + shifted-force methods (GSF and TSF) provide excellent energy
983 + conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
984 + while the hard cutoff is essentially unusable for molecular dynamics.
985 + SP provides some benefit over the hard cutoff because the energetic
986 + jumps that happen as particles leave and enter the cutoff sphere are
987 + somewhat reduced, but like the Wolf method for charges, the SP method
988 + would not be as useful for molecular dynamics as either of the
989 + shifted-force methods.
990 +
991 + We note that for all tested values of the cutoff radius, the new
992 + real-space methods can provide better energy conservation behavior
993 + than the multipolar Ewald sum, even when utilizing a relatively large
994 + $k$-space cutoff values.
995 +
996   \begin{figure}
997    \centering
998 <  \includegraphics[width=\textwidth]{newDrift.pdf}
998 >  \includegraphics[width=\textwidth]{newDrift_12.pdf}
999   \label{fig:energyDrift}        
1000 < \caption{Analysis of the energy conservation of the real space
1000 > \caption{Analysis of the energy conservation of the real-space
1001    electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
1002 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
1003 <  of energy fluctuations around this drift.  All simulations were of a
1004 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
1005 <  K starting from the same initial configuration.}
1002 >  energy over time (in kcal / mol / particle / ns) and $\delta
1003 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
1004 >  around this drift (in kcal / mol / particle).  All simulations were
1005 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
1006 >  300 K starting from the same initial configuration. All runs
1007 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
1008   \end{figure}
1009  
1010 +
1011   \section{CONCLUSION}
1012 < We have generalized the charged neutralized potential energy originally developed by the Wolf et al.\cite{Wolf:1999dn} for the charge-charge interaction to the charge-multipole and multipole-multipole interaction in the SP method for higher order multipoles. Also, we have developed GSF and TSF methods by implementing the modification purposed by Fennel and Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the higher order multipoles to ensure consistency and smooth truncation of the electrostatic energy, force, and torque for the spherical truncation. The SP methods for multipoles proved its suitability in MC simulations. On the other hand, the results from the GSF method produced good agreement with the Ewald's energy, force, and torque. Also, it shows very good energy conservation in MD simulations.
1013 < The direct truncation of any molecular system without multipole neutralization creates the fluctuation in the electrostatic energy. This fluctuation in the energy is very large for the case of crystal because of long range of multipole ordering (Refer paper I).\cite{PaperI} This is also significant in the case of the liquid because of the local multipole ordering in the molecules. If the net multipole within cutoff radius neutralized within cutoff sphere by placing image multiples on the surface of the sphere, this fluctuation in the energy reduced significantly. Also, the multipole neutralization in the generalized SP method showed very good agreement with the Ewald as compared to direct truncation for the evaluation of the $\triangle E$ between the configurations.
1014 < In MD simulations, the energy conservation is very important. The
1015 < conservation of the total energy can be ensured by  i) enforcing the
1016 < smooth truncation of the energy, force and torque in the cutoff radius
1017 < and ii) making the energy, force and torque consistent with each
1018 < other. The GSF and TSF methods ensure the consistency and smooth
1019 < truncation of the energy, force and torque at the cutoff radius, as a
1020 < result show very good total energy conservation. But the TSF method
751 < does not show good agreement in the absolute value of the
752 < electrostatic energy, force and torque with the Ewald.  The GSF method
753 < has mimicked Ewald’s force, energy and torque accurately and also
754 < conserved energy. Therefore, the GSF method is the suitable method for
755 < evaluating required force field in MD simulations. In addition, the
756 < energy drift and fluctuation from the GSF method is much better than
757 < Ewald’s method for finite-sized reciprocal space.
1012 > In the first paper in this series, we generalized the
1013 > charge-neutralized electrostatic energy originally developed by Wolf
1014 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1015 > up to quadrupolar order.  The SP method is essentially a
1016 > multipole-capable version of the Wolf model.  The SP method for
1017 > multipoles provides excellent agreement with Ewald-derived energies,
1018 > forces and torques, and is suitable for Monte Carlo simulations,
1019 > although the forces and torques retain discontinuities at the cutoff
1020 > distance that prevents its use in molecular dynamics.
1021  
1022 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1023 < that can be made extremely efficient using spline interpolations of
1024 < the radial functions.  They require no Fourier transforms or $k$-space
1025 < sums, and guarantee the smooth handling of energies, forces, and
1026 < torques as multipoles cross the real-space cutoff boundary.  
1022 > We also developed two natural extensions of the damped shifted-force
1023 > (DSF) model originally proposed by Fennel and
1024 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1025 > smooth truncation of energies, forces, and torques at the real-space
1026 > cutoff, and both converge to DSF electrostatics for point-charge
1027 > interactions.  The TSF model is based on a high-order truncated Taylor
1028 > expansion which can be relatively perturbative inside the cutoff
1029 > sphere.  The GSF model takes the gradient from an images of the
1030 > interacting multipole that has been projected onto the cutoff sphere
1031 > to derive shifted force and torque expressions, and is a significantly
1032 > more gentle approach.
1033  
1034 + Of the two newly-developed shifted force models, the GSF method
1035 + produced quantitative agreement with Ewald energy, force, and torques.
1036 + It also performs well in conserving energy in MD simulations.  The
1037 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1038 + place on a potential energy surface that is significantly perturbed
1039 + from Ewald-based electrostatics.  
1040 +
1041 + % The direct truncation of any electrostatic potential energy without
1042 + % multipole neutralization creates large fluctuations in molecular
1043 + % simulations.  This fluctuation in the energy is very large for the case
1044 + % of crystal because of long range of multipole ordering (Refer paper
1045 + % I).\cite{PaperI} This is also significant in the case of the liquid
1046 + % because of the local multipole ordering in the molecules. If the net
1047 + % multipole within cutoff radius neutralized within cutoff sphere by
1048 + % placing image multiples on the surface of the sphere, this fluctuation
1049 + % in the energy reduced significantly. Also, the multipole
1050 + % neutralization in the generalized SP method showed very good agreement
1051 + % with the Ewald as compared to direct truncation for the evaluation of
1052 + % the $\triangle E$ between the configurations.  In MD simulations, the
1053 + % energy conservation is very important. The conservation of the total
1054 + % energy can be ensured by i) enforcing the smooth truncation of the
1055 + % energy, force and torque in the cutoff radius and ii) making the
1056 + % energy, force and torque consistent with each other. The GSF and TSF
1057 + % methods ensure the consistency and smooth truncation of the energy,
1058 + % force and torque at the cutoff radius, as a result show very good
1059 + % total energy conservation. But the TSF method does not show good
1060 + % agreement in the absolute value of the electrostatic energy, force and
1061 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1062 + % energy and torque accurately and also conserved energy.
1063 +
1064 + The only cases we have found where the new GSF and SP real-space
1065 + methods can be problematic are those which retain a bulk dipole moment
1066 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1067 + materials, uniform weighting of the orientational contributions can be
1068 + important for converging the total energy.  In these cases, the
1069 + damping function which causes the non-uniform weighting can be
1070 + replaced by the bare electrostatic kernel, and the energies return to
1071 + the expected converged values.
1072 +
1073 + Based on the results of this work, the GSF method is a suitable and
1074 + efficient replacement for the Ewald sum for evaluating electrostatic
1075 + interactions in MD simulations.  Both methods retain excellent
1076 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1077 + energy drift and fluctuations from the GSF electrostatics are better
1078 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1079 + Because they use real-space cutoffs with moderate cutoff radii, the
1080 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1081 + increases.  Additionally, they can be made extremely efficient using
1082 + spline interpolations of the radial functions.  They require no
1083 + Fourier transforms or $k$-space sums, and guarantee the smooth
1084 + handling of energies, forces, and torques as multipoles cross the
1085 + real-space cutoff boundary.
1086 +
1087 + \begin{acknowledgments}
1088 +  JDG acknowledges helpful discussions with Christopher
1089 +  Fennell. Support for this project was provided by the National
1090 +  Science Foundation under grant CHE-1362211. Computational time was
1091 +  provided by the Center for Research Computing (CRC) at the
1092 +  University of Notre Dame.
1093 + \end{acknowledgments}
1094 +
1095   %\bibliographystyle{aip}
1096   \newpage
1097   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines