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 4181 by gezelter, Thu Jun 12 14:58:06 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 a reference method. The tests were carried out in a variety
74 >  of condensed-phase environments which were designed to test all
75 >  levels of the multipole-multipole interactions.  Comparisons of the
76 >  energy differences between configurations, molecular forces, and
77 >  torques were used to analyze how well the real-space models perform
78 >  relative to the more computationally expensive Ewald sum.  We have
79 >  also investigated the energy conservation properties of the new
80 >  methods in molecular dynamics simulations using all of these
81 >  methods. The SP method shows excellent agreement with
82 >  configurational energy differences, forces, and torques, and would
83 >  be suitable for use in Monte Carlo calculations.  Of the two new
84 >  shifted-force methods, the GSF approach shows the best agreement
85 >  with Ewald-derived energies, forces, and torques and exhibits energy
86 >  conservation properties that make it an excellent choice for
87 >  efficiently computing electrostatic interactions in molecular
88 >  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}
94  
# 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
# Line 113 | Line 133 | condensed systems is actually short ranged.\cite{Wolf9
133   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
134   method for calculating electrostatic interactions between point
135   charges. They argued that the effective Coulomb interaction in
136 < condensed systems is actually short ranged.\cite{Wolf92,Wolf95}.  For
137 < an ordered lattice (e.g. when computing the Madelung constant of an
136 > condensed systems is actually short ranged.\cite{Wolf92,Wolf95} For an
137 > ordered lattice (e.g., when computing the Madelung constant of an
138   ionic solid), the material can be considered as a set of ions
139   interacting with neutral dipolar or quadrupolar ``molecules'' giving
140   an effective distance dependence for the electrostatic interactions of
141 < $r^{-5}$ (see figure \ref{fig:NaCl}.  For this reason, careful
141 > $r^{-5}$ (see figure \ref{fig:schematic}).  For this reason, careful
142   applications of Wolf's method are able to obtain accurate estimates of
143   Madelung constants using relatively short cutoff radii.  Recently,
144   Fukuda used neutralization of the higher order moments for the
145   calculation of the electrostatic interaction of the point charges
146   system.\cite{Fukuda:2013sf}
147  
148 < \begin{figure}[h!]
148 > \begin{figure}
149    \centering
150 <  \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
151 <  \caption{Top: NaCl crystal showing how spherical truncation can
152 <    breaking effective charge ordering, and how complete \ce{(NaCl)4}
153 <    molecules interact with the central ion.  Bottom: A dipolar
154 <    crystal exhibiting similar behavior and illustrating how the
155 <    effective dipole-octupole interactions can be disrupted by
156 <    spherical truncation.}
157 <  \label{fig:NaCl}
150 >  \includegraphics[width=\linewidth]{schematic.pdf}
151 >  \caption{Top: Ionic systems exhibit local clustering of dissimilar
152 >    charges (in the smaller grey circle), so interactions are
153 >    effectively charge-multipole in order at longer distances.  With
154 >    hard cutoffs, motion of individual charges in and out of the
155 >    cutoff sphere can break the effective multipolar ordering.
156 >    Bottom: dipolar crystals and fluids have a similar effective
157 >    \textit{quadrupolar} ordering (in the smaller grey circles), and
158 >    orientational averaging helps to reduce the effective range of the
159 >    interactions in the fluid.  Placement of reversed image multipoles
160 >    on the surface of the cutoff sphere recovers the effective
161 >    higher-order multipole behavior.}
162 >  \label{fig:schematic}
163   \end{figure}
164  
165   The direct truncation of interactions at a cutoff radius creates
# Line 158 | Line 183 | potential is found to be decreasing as $r^{-5}$. If on
183  
184   Considering the interaction of one central ion in an ionic crystal
185   with a portion of the crystal at some distance, the effective Columbic
186 < potential is found to be decreasing as $r^{-5}$. If one views the
187 < \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
186 > potential is found to decrease as $r^{-5}$. If one views the \ce{NaCl}
187 > crystal as a simple cubic (SC) structure with an octupolar
188   \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
189   rapidly to the Madelung energy than the dipolar
190   approximation.\cite{Wolf92} To find the correct Madelung constant,
191   Lacman suggested that the NaCl structure could be constructed in a way
192   that the finite crystal terminates with complete \ce{(NaCl)4}
193 < molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
194 < by opposite charges. Similarly for each pair of charges, there is an
195 < opposite pair of charge adjacent to it.  The central ion sees what is
196 < 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}
193 > molecules.\cite{Lacman65} The central ion sees what is effectively a
194 > set of octupoles at large distances. These facts suggest that the
195 > Madelung constants are relatively short ranged for perfect ionic
196 > crystals.\cite{Wolf:1999dn}
197  
198   One can make a similar argument for crystals of point multipoles. The
199   Luttinger and Tisza treatment of energy constants for dipolar lattices
# Line 188 | Line 211 | multipolar arrangements (see Fig. \ref{fig:NaCl}), cau
211   unstable.
212  
213   In ionic crystals, real-space truncation can break the effective
214 < multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
215 < swings in the electrostatic energy as the cutoff radius is increased
216 < (or as individual ions move back and forth across the boundary).  This
217 < is why the image charges were necessary for the Wolf sum to exhibit
218 < rapid convergence.  Similarly, the real-space truncation of point
219 < multipole interactions breaks higher order multipole arrangements, and
220 < image multipoles are required for real-space treatments of
198 < electrostatic energies.
214 > multipolar arrangements (see Fig. \ref{fig:schematic}), causing
215 > significant swings in the electrostatic energy as individual ions move
216 > back and forth across the boundary.  This is why the image charges are
217 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
218 > the real-space truncation of point multipole interactions breaks
219 > higher order multipole arrangements, and image multipoles are required
220 > for real-space treatments of electrostatic energies.
221  
222 + The shorter effective range of electrostatic interactions is not
223 + limited to perfect crystals, but can also apply in disordered fluids.
224 + Even at elevated temperatures, there is, on average, local charge
225 + balance in an ionic liquid, where each positive ion has surroundings
226 + dominated by negaitve ions and vice versa.  The reversed-charge images
227 + on the cutoff sphere that are integral to the Wolf and DSF approaches
228 + retain the effective multipolar interactions as the charges traverse
229 + the cutoff boundary.
230 +
231 + In multipolar fluids (see Fig. \ref{fig:schematic}) there is
232 + significant orientational averaging that additionally reduces the
233 + effect of long-range multipolar interactions.  The image multipoles
234 + that are introduced in the TSF, GSF, and SP methods mimic this effect
235 + and reduce the effective range of the multipolar interactions as
236 + interacting molecules traverse each other's cutoff boundaries.
237 +
238   % Because of this reason, although the nature of electrostatic
239   % interaction short ranged, the hard cutoff sphere creates very large
240   % fluctuation in the electrostatic energy for the perfect crystal. In
# Line 253 | Line 291 | multipoles up to octupolar
291   rough approximation.  Atomic sites can also be assigned point
292   multipoles and polarizabilities to increase the accuracy of the
293   molecular model.  Recently, water has been modeled with point
294 < multipoles up to octupolar
295 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
294 > multipoles up to octupolar order using the soft sticky
295 > dipole-quadrupole-octupole (SSDQO)
296 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
297   multipoles up to quadrupolar order have also been coupled with point
298   polarizabilities in the high-quality AMOEBA and iAMOEBA water
299 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
299 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
300   using point multipole with the real space truncation without
301   accounting for multipolar neutrality will create energy conservation
302   issues in molecular dynamics (MD) simulations.
# Line 297 | Line 336 | where the multipole operator for site $\bf a$,
336   \begin{equation}
337   U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
338   \end{equation}
339 < where the multipole operator for site $\bf a$,
340 < \begin{equation}
341 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
342 < +  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.
339 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
340 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
341 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
342 > $\bf a$.
343  
344 < Interactions between multipoles can be expressed as higher derivatives
345 < of the bare Coulomb potential, so one way of ensuring that the forces
346 < and torques vanish at the cutoff distance is to include a larger
347 < number of terms in the truncated Taylor expansion, e.g.,
348 < %
349 < \begin{equation}
350 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
351 < \end{equation}
352 < %
353 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
354 < Thus, for $f(r)=1/r$, we find
355 < %
356 < \begin{equation}
357 < 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} .
358 < \end{equation}
359 < This function is an approximate electrostatic potential that has
360 < vanishing second derivatives at the cutoff radius, making it suitable
361 < for shifting the forces and torques of charge-dipole interactions.
344 > % Interactions between multipoles can be expressed as higher derivatives
345 > % of the bare Coulomb potential, so one way of ensuring that the forces
346 > % and torques vanish at the cutoff distance is to include a larger
347 > % number of terms in the truncated Taylor expansion, e.g.,
348 > % %
349 > % \begin{equation}
350 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
351 > % \end{equation}
352 > % %
353 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
354 > % Thus, for $f(r)=1/r$, we find
355 > % %
356 > % \begin{equation}
357 > % 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} .
358 > % \end{equation}
359 > % This function is an approximate electrostatic potential that has
360 > % vanishing second derivatives at the cutoff radius, making it suitable
361 > % for shifting the forces and torques of charge-dipole interactions.
362  
363 < In general, the TSF potential for any multipole-multipole interaction
364 < can be written
363 > The TSF potential for any multipole-multipole interaction can be
364 > written
365   \begin{equation}
366   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
367   \label{generic}
368   \end{equation}
369 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
370 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
371 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
372 < energy, force, and torques, the required number of terms from Taylor
373 < series expansion in $f_n(r)$ must be performed for different
374 < multipole-multipole interactions.
369 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
370 > of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
371 > charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
372 > and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
373 > quadrupole-quadrupole.  To ensure smooth convergence of the energy,
374 > force, and torques, a Taylor expansion with $n$ terms must be
375 > performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
376  
377 < To carry out the same procedure for a damped electrostatic kernel, we
378 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
379 < Many of the derivatives of the damped kernel are well known from
380 < Smith's early work on multipoles for the Ewald
381 < summation.\cite{Smith82,Smith98}
377 > % To carry out the same procedure for a damped electrostatic kernel, we
378 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
379 > % Many of the derivatives of the damped kernel are well known from
380 > % Smith's early work on multipoles for the Ewald
381 > % summation.\cite{Smith82,Smith98}
382  
383 < Note that increasing the value of $n$ will add additional terms to the
384 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
385 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
386 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
387 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
388 < for computing multipole energies, forces, and torques, and smooth
389 < cutoffs of these quantities can be guaranteed as long as the number of
390 < terms in the Taylor series exceeds the derivative order required.
383 > % Note that increasing the value of $n$ will add additional terms to the
384 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
385 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
386 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
387 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
388 > % for computing multipole energies, forces, and torques, and smooth
389 > % cutoffs of these quantities can be guaranteed as long as the number of
390 > % terms in the Taylor series exceeds the derivative order required.
391  
392   For multipole-multipole interactions, following this procedure results
393 < in separate radial functions for each distinct orientational
394 < contribution to the potential, and ensures that the forces and torques
395 < from {\it each} of these contributions will vanish at the cutoff
396 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
397 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
393 > in separate radial functions for each of the distinct orientational
394 > contributions to the potential, and ensures that the forces and
395 > torques from each of these contributions will vanish at the cutoff
396 > radius.  For example, the direct dipole dot product
397 > ($\mathbf{D}_{\bf a}
398 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
399   dot products:
400   \begin{equation}
401 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
402 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
403 < -\frac{1}{4\pi \epsilon_0}
404 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
405 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
401 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
402 >  \mathbf{D}_{\bf a} \cdot
403 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
404 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
405 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
406   \end{equation}
407  
408 < The electrostatic forces and torques acting on the central multipole
409 < site due to another site within cutoff sphere are derived from
408 > For the Taylor shifted (TSF) method with the undamped kernel,
409 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
410 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
411 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
412 > connection to unmodified electrostatics as well as the smooth
413 > transition to zero in both these functions as $r\rightarrow r_c$.  The
414 > electrostatic forces and torques acting on the central multipole due
415 > to another site within cutoff sphere are derived from
416   Eq.~\ref{generic}, accounting for the appropriate number of
417   derivatives. Complete energy, force, and torque expressions are
418   presented in the first paper in this series (Reference
419 < \citep{PaperI}).
419 > \onlinecite{PaperI}).
420  
421   \subsection{Gradient-shifted force (GSF)}
422  
423 < A second (and significantly simpler) method involves shifting the
424 < gradient of the raw coulomb potential for each particular multipole
423 > A second (and conceptually simpler) method involves shifting the
424 > gradient of the raw Coulomb potential for each particular multipole
425   order.  For example, the raw dipole-dipole potential energy may be
426   shifted smoothly by finding the gradient for two interacting dipoles
427   which have been projected onto the surface of the cutoff sphere
428   without changing their relative orientation,
429 < \begin{displaymath}
430 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
431 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
432 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
433 < \end{displaymath}
434 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
435 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
436 < the signs are reversed for the dipole that has been projected onto the
437 < cutoff sphere).  In many ways, this simpler approach is closer in
438 < spirit to the original shifted force method, in that it projects a
439 < neutralizing multipole (and the resulting forces from this multipole)
440 < onto a cutoff sphere. The resulting functional forms for the
441 < potentials, forces, and torques turn out to be quite similar in form
442 < to the Taylor-shifted approach, although the radial contributions are
443 < significantly less perturbed by the Gradient-shifted approach than
444 < they are in the Taylor-shifted method.
429 > \begin{equation}
430 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
431 > U_{D_{\bf a} D_{\bf b}}(r_c)
432 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
433 >  \vec{\nabla} U_{D_{\bf a}D_{\bf b}}(r) \Big \lvert _{r_c}
434 > \end{equation}
435 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
436 >  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
437 > (although the signs are reversed for the dipole that has been
438 > projected onto the cutoff sphere).  In many ways, this simpler
439 > approach is closer in spirit to the original shifted force method, in
440 > that it projects a neutralizing multipole (and the resulting forces
441 > from this multipole) onto a cutoff sphere. The resulting functional
442 > forms for the potentials, forces, and torques turn out to be quite
443 > similar in form to the Taylor-shifted approach, although the radial
444 > contributions are significantly less perturbed by the gradient-shifted
445 > approach than they are in the Taylor-shifted method.
446  
447 + For the gradient shifted (GSF) method with the undamped kernel,
448 + $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
449 + $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
450 + Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
451 + because the Taylor expansion retains only one term, they are
452 + significantly less perturbed than the TSF functions.
453 +
454   In general, the gradient shifted potential between a central multipole
455   and any multipolar site inside the cutoff radius is given by,
456   \begin{equation}
# Line 419 | Line 466 | approach zero as $r \rightarrow R_c$.  Both the GSF an
466   function of radius, hence the contribution of the third term is very
467   small for large cutoff radii.  The force and torque derived from
468   equation \ref{generic2} are consistent with the energy expression and
469 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
469 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
470   can be considered generalizations of the original DSF method for
471   higher order multipole interactions. GSF and TSF are also identical up
472   to the charge-dipole interaction but generate different expressions in
473   the energy, force and torque for higher order multipole-multipole
474   interactions. Complete energy, force, and torque expressions for the
475   GSF potential are presented in the first paper in this series
476 < (Reference \citep{PaperI})
476 > (Reference~\onlinecite{PaperI})
477  
478  
479   \subsection{Shifted potential (SP) }
# Line 439 | Line 486 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
486   interactions with the central multipole and the image. This
487   effectively shifts the total potential to zero at the cutoff radius,
488   \begin{equation}
489 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
489 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
490 > U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
491   \label{eq:SP}
492   \end{equation}          
493   where the sum describes separate potential shifting that is done for
494   each orientational contribution to the energy (e.g. the direct dipole
495   product contribution is shifted {\it separately} from the
496   dipole-distance terms in dipole-dipole interactions).  Note that this
497 < is not a simple shifting of the total potential at $R_c$. Each radial
497 > is not a simple shifting of the total potential at $r_c$. Each radial
498   contribution is shifted separately.  One consequence of this is that
499   multipoles that reorient after leaving the cutoff sphere can re-enter
500   the cutoff sphere without perturbing the total energy.
501  
502 < The potential energy between a central multipole and other multipolar
503 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
504 < force and torque obtained from the shifted potential (SP) are
505 < discontinuous at $R_c$. Therefore, MD simulations will still
506 < experience energy drift while operating under the SP potential, but it
507 < may be suitable for Monte Carlo approaches where the configurational
508 < energy differences are the primary quantity of interest.
502 > For the shifted potential (SP) method with the undamped kernel,
503 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
504 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
505 > central multipole and other multipolar sites goes smoothly to zero as
506 > $r \rightarrow r_c$.  However, the force and torque obtained from the
507 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
508 > will still experience energy drift while operating under the SP
509 > potential, but it may be suitable for Monte Carlo approaches where the
510 > configurational energy differences are the primary quantity of
511 > interest.
512  
513 < \subsection{The Self term}
513 > \subsection{The Self Term}
514   In the TSF, GSF, and SP methods, a self-interaction is retained for
515   the central multipole interacting with its own image on the surface of
516   the cutoff sphere.  This self interaction is nearly identical with the
517   self-terms that arise in the Ewald sum for multipoles.  Complete
518   expressions for the self terms are presented in the first paper in
519 < this series (Reference \citep{PaperI})  
519 > this series (Reference \onlinecite{PaperI}).
520  
521  
522   \section{\label{sec:methodology}Methodology}
# Line 477 | Line 528 | arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} Thi
528   real-space cutoffs.  In the first paper of this series, we compared
529   the dipolar and quadrupolar energy expressions against analytic
530   expressions for ordered dipolar and quadrupolar
531 < arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
532 < multipolar Ewald sum as a reference method for comparing energies,
533 < forces, and torques for molecular models that mimic disordered and
534 < ordered condensed-phase systems.  These test-cases include:
531 > arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
532 > used the multipolar Ewald sum as a reference method for comparing
533 > energies, forces, and torques for molecular models that mimic
534 > disordered and ordered condensed-phase systems.  The parameters used
535 > in the test cases are given in table~\ref{tab:pars}.
536  
537 < \begin{itemize}
538 < \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
539 < \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
540 < \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
541 < \item Soft Quadrupolar solids  ($\sigma = , \epsilon = , Q_{xx} = ...$)
542 < \item A mixed multipole model for water
543 < \item A mixed multipole models for water with dissolved ions
544 < \end{itemize}
545 < This last test case exercises all levels of the multipole-multipole
546 < interactions we have derived so far and represents the most complete
547 < test of the new methods.
537 > \begin{table}
538 > \label{tab:pars}
539 > \caption{The parameters used in the systems used to evaluate the new
540 >  real-space methods.  The most comprehensive test was a liquid
541 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
542 >  ions).  This test excercises all orders of the multipolar
543 >  interactions developed in the first paper.}
544 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
545 >             & \multicolumn{2}{c|}{LJ parameters} &
546 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
547 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
548 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
549 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
550 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
551 > \AA\textsuperscript{2})} \\ \hline
552 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
553 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
554 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
555 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
556 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
557 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
558 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
559 > \end{tabularx}
560 > \end{table}
561 > The systems consist of pure multipolar solids (both dipole and
562 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
563 > fluid composed of sites containing both dipoles and quadrupoles
564 > simultaneously, and a final test case that includes ions with point
565 > charges in addition to the multipolar fluid.  The solid-phase
566 > parameters were chosen so that the systems can explore some
567 > orientational freedom for the multipolar sites, while maintaining
568 > relatively strict translational order.  The SSDQ model used here is
569 > not a particularly accurate water model, but it does test
570 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
571 > interactions at roughly the same magnitudes. The last test case, SSDQ
572 > water with dissolved ions, exercises \textit{all} levels of the
573 > multipole-multipole interactions we have derived so far and represents
574 > the most complete test of the new methods.
575  
576   In the following section, we present results for the total
577   electrostatic energy, as well as the electrostatic contributions to
578   the force and torque on each molecule.  These quantities have been
579   computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
580 < and have been compared with the values obtaine from the multipolar
581 < Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
580 > and have been compared with the values obtained from the multipolar
581 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
582   between two configurations is the primary quantity that governs how
583   the simulation proceeds. These differences are the most imporant
584   indicators of the reliability of a method even if the absolute
# Line 510 | Line 589 | contributions to the forces and torques.
589   behavior of the simulation, so we also compute the electrostatic
590   contributions to the forces and torques.
591  
592 < \subsection{Model systems}
593 < To sample independent configurations of multipolar crystals, a body
594 < centered cubic (BCC) crystal which is a minimum energy structure for
595 < point dipoles was generated using 3,456 molecules.  The multipoles
596 < were translationally locked in their respective crystal sites for
597 < equilibration at a relatively low temperature (50K), so that dipoles
598 < or quadrupoles could freely explore all accessible orientations.  The
599 < 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.
592 > \subsection{Implementation}
593 > The real-space methods developed in the first paper in this series
594 > have been implemented in our group's open source molecular simulation
595 > program, OpenMD,\cite{openmd} which was used for all calculations in
596 > this work.  The complementary error function can be a relatively slow
597 > function on some processors, so all of the radial functions are
598 > precomputed on a fine grid and are spline-interpolated to provide
599 > values when required.  
600  
601 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
602 < system was created with 2048 molecules oriented randomly.  These were
601 > Using the same simulation code, we compare to a multipolar Ewald sum
602 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
603 > the Ewald sum is a re-implementation of the algorithm originally
604 > proposed by Smith that does not use the particle mesh or smoothing
605 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
606 > being compared are the electrostatic contributions to energies, force,
607 > and torques.  All other contributions to these quantities (i.e. from
608 > Lennard-Jones interactions) are removed prior to the comparisons.
609  
610 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
611 < temperature after equilibration.  We collected 250 different
612 < configurations in equal interval of time. For the ions mixed liquid
613 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
614 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
615 < at the same environment for 1ns and 250 configurations were
616 < collected. While comparing energies, forces, and torques with Ewald
617 < method, Lennard-Jones potentials were turned off and purely
618 < electrostatic interaction had been compared.
610 > The convergence parameter ($\alpha$) also plays a role in the balance
611 > of the real-space and reciprocal-space portions of the Ewald
612 > calculation.  Typical molecular mechanics packages set this to a value
613 > that depends on the cutoff radius and a tolerance (typically less than
614 > $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
615 > associated with increasing accuracy at the expense of computational
616 > time spent on the reciprocal-space portion of the
617 > summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
618 > 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
619 > Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
620 >
621 > The real-space models have self-interactions that provide
622 > contributions to the energies only.  Although the self interaction is
623 > a rapid calculation, we note that in systems with fluctuating charges
624 > or point polarizabilities, the self-term is not static and must be
625 > recomputed at each time step.
626 >
627 > \subsection{Model systems}
628 > To sample independent configurations of the multipolar crystals, body
629 > centered cubic (bcc) crystals, which exhibit the minimum energy
630 > structures for point dipoles, were generated using 3,456 molecules.
631 > The multipoles were translationally locked in their respective crystal
632 > sites for equilibration at a relatively low temperature (50K) so that
633 > dipoles or quadrupoles could freely explore all accessible
634 > orientations.  The translational constraints were then removed, the
635 > systems were re-equilibrated, and the crystals were simulated for an
636 > additional 10 ps in the microcanonical (NVE) ensemble with an average
637 > temperature of 50 K.  The balance between moments of inertia and
638 > particle mass were chosen to allow orientational sampling without
639 > significant translational motion.  Configurations were sampled at
640 > equal time intervals in order to compare configurational energy
641 > differences.  The crystals were simulated far from the melting point
642 > in order to avoid translational deformation away of the ideal lattice
643 > geometry.
644  
645 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
646 + simulations, each system was created with 2,048 randomly-oriented
647 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
648 + Each system was then simulated for 1 ns in the microcanonical (NVE)
649 + ensemble.  We collected 250 different configurations at equal time
650 + intervals. For the liquid system that included ionic species, we
651 + converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
652 + \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
653 + run under the same conditions for 1 ns. A total of 250 configurations
654 + were collected. In the following comparisons of energies, forces, and
655 + torques, the Lennard-Jones potentials were turned off and only the
656 + purely electrostatic quantities were compared with the same values
657 + obtained via the Ewald sum.
658 +
659   \subsection{Accuracy of Energy Differences, Forces and Torques}
660   The pairwise summation techniques (outlined above) were evaluated for
661   use in MC simulations by studying the energy differences between
# Line 550 | Line 668 | we used least square regressions analysiss for the six
668   should be identical for all methods.
669  
670   Since none of the real-space methods provide exact energy differences,
671 < we used least square regressions analysiss for the six different
671 > we used least square regressions analysis for the six different
672   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
673   with the multipolar Ewald reference method.  Unitary results for both
674   the correlation (slope) and correlation coefficient for these
# Line 561 | Line 679 | also been compared by using least squares regression a
679   configurations and 250 configurations were recorded for comparison.
680   Each system provided 31,125 energy differences for a total of 186,750
681   data points.  Similarly, the magnitudes of the forces and torques have
682 < also been compared by using least squares regression analyses. In the
682 > also been compared using least squares regression analysis. In the
683   forces and torques comparison, the magnitudes of the forces acting in
684   each molecule for each configuration were evaluated. For example, our
685   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 647 | Line 765 | model must allow for long simulation times with minima
765        
766   %        \label{fig:barGraph2}
767   %      \end{figure}
768 < %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.
768 > %The correlation coefficient ($R^2$) and slope of the linear
769 > %regression plots for the energy differences for all six different
770 > %molecular systems is shown in figure 4a and 4b.The plot shows that
771 > %the correlation coefficient improves for the SP cutoff method as
772 > %compared to the undamped hard cutoff method in the case of SSDQC,
773 > %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
774 > %crystal and liquid, the correlation coefficient is almost unchanged
775 > %and close to 1.  The correlation coefficient is smallest (0.696276
776 > %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
777 > %charge-charge and charge-multipole interactions. Since the
778 > %charge-charge and charge-multipole interaction is long ranged, there
779 > %is huge deviation of correlation coefficient from 1. Similarly, the
780 > %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
781 > %compared to interactions in the other multipolar systems, thus the
782 > %correlation coefficient very close to 1 even for hard cutoff
783 > %method. The idea of placing image multipole on the surface of the
784 > %cutoff sphere improves the correlation coefficient and makes it close
785 > %to 1 for all types of multipolar systems. Similarly the slope is
786 > %hugely deviated from the correct value for the lower order
787 > %multipole-multipole interaction and slightly deviated for higher
788 > %order multipole – multipole interaction. The SP method improves both
789 > %correlation coefficient ($R^2$) and slope significantly in SSDQC and
790 > %dipolar systems.  The Slope is found to be deviated more in dipolar
791 > %crystal as compared to liquid which is associated with the large
792 > %fluctuation in the electrostatic energy in crystal. The GSF also
793 > %produced better values of correlation coefficient and slope with the
794 > %proper selection of the damping alpha (Interested reader can consult
795 > %accompanying supporting material). The TSF method gives good value of
796 > %correlation coefficient for the dipolar crystal, dipolar liquid,
797 > %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
798 > %regression slopes are significantly deviated.
799 >
800   \begin{figure}
801 <        \centering
802 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
803 <        \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)}
804 <        \label{fig:slopeCorr_energy}
805 <    \end{figure}
806 < 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.
801 >  \centering
802 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
803 >  \caption{Statistical analysis of the quality of configurational
804 >    energy differences for the real-space electrostatic methods
805 >    compared with the reference Ewald sum.  Results with a value equal
806 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
807 >    from those obtained using the multipolar Ewald sum.  Different
808 >    values of the cutoff radius are indicated with different symbols
809 >    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
810 >    triangles).}
811 >  \label{fig:slopeCorr_energy}
812 > \end{figure}
813 >
814 > The combined correlation coefficient and slope for all six systems is
815 > shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
816 > reproduce the Ewald configurational energy differences with remarkable
817 > fidelity.  Undamped hard cutoffs introduce a significant amount of
818 > random scatter in the energy differences which is apparent in the
819 > reduced value of the correlation coefficient for this method.  This
820 > can be easily understood as configurations which exhibit small
821 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
822 > will see large energy jumps when hard cutoffs are used.  The
823 > orientations of the multipoles (particularly in the ordered crystals)
824 > mean that these energy jumps can go in either direction, producing a
825 > significant amount of random scatter, but no systematic error.
826 >
827 > The TSF method produces energy differences that are highly correlated
828 > with the Ewald results, but it also introduces a significant
829 > systematic bias in the values of the energies, particularly for
830 > smaller cutoff values. The TSF method alters the distance dependence
831 > of different orientational contributions to the energy in a
832 > non-uniform way, so the size of the cutoff sphere can have a large
833 > effect, particularly for the crystalline systems.
834 >
835 > Both the SP and GSF methods appear to reproduce the Ewald results with
836 > excellent fidelity, particularly for moderate damping ($\alpha =
837 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
838 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
839 > method with short cutoffs, all of the methods would be appropriate for
840 > use in Monte Carlo simulations.
841 >
842   \subsection{Magnitude of the force and torque vectors}
843 < 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.
843 >
844 > The comparisons of the magnitudes of the forces and torques for the
845 > data accumulated from all six systems are shown in Figures
846 > ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
847 > correlation and slope for the forces agree well with the Ewald sum
848 > even for the hard cutoffs.
849 >
850 > For systems of molecules with only multipolar interactions, the pair
851 > energy contributions are quite short ranged.  Moreover, the force
852 > decays more rapidly than the electrostatic energy, hence the hard
853 > cutoff method can also produce reasonable agreement for this quantity.
854 > Although the pure cutoff gives reasonably good electrostatic forces
855 > for pairs of molecules included within each other's cutoff spheres,
856 > the discontinuity in the force at the cutoff radius can potentially
857 > cause energy conservation problems as molecules enter and leave the
858 > cutoff spheres.  This is discussed in detail in section
859 > \ref{sec:conservation}.
860 >
861 > The two shifted-force methods (GSF and TSF) exhibit a small amount of
862 > systematic variation and scatter compared with the Ewald forces.  The
863 > shifted-force models intentionally perturb the forces between pairs of
864 > molecules inside each other's cutoff spheres in order to correct the
865 > energy conservation issues, and this perturbation is evident in the
866 > statistics accumulated for the molecular forces.  The GSF
867 > perturbations are minimal, particularly for moderate damping and
868 > commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
869 > reasonable agreement in the correlation coefficient but again the
870 > systematic error in the forces is concerning if replication of Ewald
871 > forces is desired.
872 >
873   \begin{figure}
874 <        \centering
875 <        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
876 <        \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). }
877 <        \label{fig:slopeCorr_force}
878 <    \end{figure}
879 < 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.
874 >  \centering
875 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
876 >  \caption{Statistical analysis of the quality of the force vector
877 >    magnitudes for the real-space electrostatic methods compared with
878 >    the reference Ewald sum. Results with a value equal to 1 (dashed
879 >    line) indicate force magnitude values indistinguishable from those
880 >    obtained using the multipolar Ewald sum.  Different values of the
881 >    cutoff radius are indicated with different symbols (9\AA\ =
882 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
883 >  \label{fig:slopeCorr_force}
884 > \end{figure}
885 >
886 >
887   \begin{figure}
888 <        \centering
889 <        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
890 <        \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).}
891 <        \label{fig:slopeCorr_torque}
892 <    \end{figure}
888 >  \centering
889 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
890 >  \caption{Statistical analysis of the quality of the torque vector
891 >    magnitudes for the real-space electrostatic methods compared with
892 >    the reference Ewald sum. Results with a value equal to 1 (dashed
893 >    line) indicate force magnitude values indistinguishable from those
894 >    obtained using the multipolar Ewald sum.  Different values of the
895 >    cutoff radius are indicated with different symbols (9\AA\ =
896 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
897 >  \label{fig:slopeCorr_torque}
898 > \end{figure}
899 >
900 > The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
901 > significantly influenced by the choice of real-space method.  The
902 > torque expressions have the same distance dependence as the energies,
903 > which are naturally longer-ranged expressions than the inter-site
904 > forces.  Torques are also quite sensitive to orientations of
905 > neighboring molecules, even those that are near the cutoff distance.
906 >
907 > The results shows that the torque from the hard cutoff method
908 > reproduces the torques in quite good agreement with the Ewald sum.
909 > The other real-space methods can cause some deviations, but excellent
910 > agreement with the Ewald sum torques is recovered at moderate values
911 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
912 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
913 > in the slope when compared with the Ewald torques even for larger
914 > cutoff radii.  It appears that the severity of the perturbations in
915 > the TSF method are most in evidence for the torques.
916 >
917   \subsection{Directionality of the force and torque vectors}  
674 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 $r_c$.
918  
919 + The accurate evaluation of force and torque directions is just as
920 + important for molecular dynamics simulations as the magnitudes of
921 + these quantities. Force and torque vectors for all six systems were
922 + analyzed using Fisher statistics, and the quality of the vector
923 + directionality is shown in terms of circular variance
924 + ($\mathrm{Var}(\theta)$) in figure
925 + \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
926 + from the new real-space methods exhibit nearly-ideal Fisher probability
927 + distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
928 + exhibit the best vectorial agreement with the Ewald sum. The force and
929 + torque vectors from GSF method also show good agreement with the Ewald
930 + method, which can also be systematically improved by using moderate
931 + damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
932 + 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
933 + to a distribution with 95\% of force vectors within $6.37^\circ$ of
934 + the corresponding Ewald forces. The TSF method produces the poorest
935 + agreement with the Ewald force directions.
936 +
937 + Torques are again more perturbed than the forces by the new real-space
938 + methods, but even here the variance is reasonably small.  For the same
939 + method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
940 + the circular variance was 0.01415, corresponds to a distribution which
941 + has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
942 + results. Again, the direction of the force and torque vectors can be
943 + systematically improved by varying $\alpha$ and $r_c$.
944 +
945   \begin{figure}
946 <        \centering
947 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
948 <        \caption{The circular variance of the data sets of the
949 <          direction of the  force and torque vectors obtained from a
950 <          given method about reference Ewald method. The result equal
951 <          to 0 (dashed line) indicates direction of the vectors are
952 <          indistinguishable from the Ewald method. Here different
953 <          symbols represent different value of the cutoff radius (9
954 <          \AA\ = circle, 12 \AA\ = square, 15 \AA\  = inverted triangle)}
955 <        \label{fig:slopeCorr_circularVariance}
956 <    \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.
946 >  \centering
947 >  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
948 >  \caption{The circular variance of the direction of the force and
949 >    torque vectors obtained from the real-space methods around the
950 >    reference Ewald vectors. A variance equal to 0 (dashed line)
951 >    indicates direction of the force or torque vectors are
952 >    indistinguishable from those obtained from the Ewald sum. Here
953 >    different symbols represent different values of the cutoff radius
954 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
955 >  \label{fig:slopeCorr_circularVariance}
956 > \end{figure}
957  
958 < 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.
958 > \subsection{Energy conservation\label{sec:conservation}}
959  
960 + We have tested the conservation of energy one can expect to see with
961 + the new real-space methods using the SSDQ water model with a small
962 + fraction of solvated ions. This is a test system which exercises all
963 + orders of multipole-multipole interactions derived in the first paper
964 + in this series and provides the most comprehensive test of the new
965 + methods.  A liquid-phase system was created with 2000 water molecules
966 + and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
967 + temperature of 300K.  After equilibration, this liquid-phase system
968 + was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
969 + a cutoff radius of 12\AA.  The value of the damping coefficient was
970 + also varied from the undamped case ($\alpha = 0$) to a heavily damped
971 + case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
972 + sample was also run using the multipolar Ewald sum with the same
973 + real-space cutoff.
974 +
975 + In figure~\ref{fig:energyDrift} we show the both the linear drift in
976 + energy over time, $\delta E_1$, and the standard deviation of energy
977 + fluctuations around this drift $\delta E_0$.  Both of the
978 + shifted-force methods (GSF and TSF) provide excellent energy
979 + conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
980 + while the hard cutoff is essentially unusable for molecular dynamics.
981 + SP provides some benefit over the hard cutoff because the energetic
982 + jumps that happen as particles leave and enter the cutoff sphere are
983 + somewhat reduced, but like the Wolf method for charges, the SP method
984 + would not be as useful for molecular dynamics as either of the
985 + shifted-force methods.
986 +
987 + We note that for all tested values of the cutoff radius, the new
988 + real-space methods can provide better energy conservation behavior
989 + than the multipolar Ewald sum, even when utilizing a relatively large
990 + $k$-space cutoff values.
991 +
992   \begin{figure}
993    \centering
994 <  \includegraphics[width=\textwidth]{newDrift.pdf}
994 >  \includegraphics[width=\textwidth]{newDrift_12.pdf}
995   \label{fig:energyDrift}        
996 < \caption{Analysis of the energy conservation of the real space
996 > \caption{Analysis of the energy conservation of the real-space
997    electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
998 <  energy over time and $\delta \mathrm{E}_0$ is the standard deviation
999 <  of energy fluctuations around this drift.  All simulations were of a
1000 <  2000-molecule simulation of SSDQ water with 48 ionic charges at 300
1001 <  K starting from the same initial configuration.}
998 >  energy over time (in kcal / mol / particle / ns) and $\delta
999 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
1000 >  around this drift (in kcal / mol / particle).  All simulations were
1001 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
1002 >  300 K starting from the same initial configuration. All runs
1003 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
1004   \end{figure}
1005  
1006 +
1007   \section{CONCLUSION}
1008 < 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.
1009 < 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.
1010 < In MD simulations, the energy conservation is very important. The
1011 < conservation of the total energy can be ensured by  i) enforcing the
1012 < smooth truncation of the energy, force and torque in the cutoff radius
1013 < and ii) making the energy, force and torque consistent with each
1014 < other. The GSF and TSF methods ensure the consistency and smooth
1015 < truncation of the energy, force and torque at the cutoff radius, as a
1016 < 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.
1008 > In the first paper in this series, we generalized the
1009 > charge-neutralized electrostatic energy originally developed by Wolf
1010 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1011 > up to quadrupolar order.  The SP method is essentially a
1012 > multipole-capable version of the Wolf model.  The SP method for
1013 > multipoles provides excellent agreement with Ewald-derived energies,
1014 > forces and torques, and is suitable for Monte Carlo simulations,
1015 > although the forces and torques retain discontinuities at the cutoff
1016 > distance that prevents its use in molecular dynamics.
1017  
1018 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1019 < that can be made extremely efficient using spline interpolations of
1020 < the radial functions.  They require no Fourier transforms or $k$-space
1021 < sums, and guarantee the smooth handling of energies, forces, and
1022 < torques as multipoles cross the real-space cutoff boundary.  
1018 > We also developed two natural extensions of the damped shifted-force
1019 > (DSF) model originally proposed by Fennel and
1020 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1021 > smooth truncation of energies, forces, and torques at the real-space
1022 > cutoff, and both converge to DSF electrostatics for point-charge
1023 > interactions.  The TSF model is based on a high-order truncated Taylor
1024 > expansion which can be relatively perturbative inside the cutoff
1025 > sphere.  The GSF model takes the gradient from an images of the
1026 > interacting multipole that has been projected onto the cutoff sphere
1027 > to derive shifted force and torque expressions, and is a significantly
1028 > more gentle approach.
1029  
1030 + Of the two newly-developed shifted force models, the GSF method
1031 + produced quantitative agreement with Ewald energy, force, and torques.
1032 + It also performs well in conserving energy in MD simulations.  The
1033 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1034 + place on a potential energy surface that is significantly perturbed
1035 + from Ewald-based electrostatics.  
1036 +
1037 + % The direct truncation of any electrostatic potential energy without
1038 + % multipole neutralization creates large fluctuations in molecular
1039 + % simulations.  This fluctuation in the energy is very large for the case
1040 + % of crystal because of long range of multipole ordering (Refer paper
1041 + % I).\cite{PaperI} This is also significant in the case of the liquid
1042 + % because of the local multipole ordering in the molecules. If the net
1043 + % multipole within cutoff radius neutralized within cutoff sphere by
1044 + % placing image multiples on the surface of the sphere, this fluctuation
1045 + % in the energy reduced significantly. Also, the multipole
1046 + % neutralization in the generalized SP method showed very good agreement
1047 + % with the Ewald as compared to direct truncation for the evaluation of
1048 + % the $\triangle E$ between the configurations.  In MD simulations, the
1049 + % energy conservation is very important. The conservation of the total
1050 + % energy can be ensured by i) enforcing the smooth truncation of the
1051 + % energy, force and torque in the cutoff radius and ii) making the
1052 + % energy, force and torque consistent with each other. The GSF and TSF
1053 + % methods ensure the consistency and smooth truncation of the energy,
1054 + % force and torque at the cutoff radius, as a result show very good
1055 + % total energy conservation. But the TSF method does not show good
1056 + % agreement in the absolute value of the electrostatic energy, force and
1057 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1058 + % energy and torque accurately and also conserved energy.
1059 +
1060 + The only cases we have found where the new GSF and SP real-space
1061 + methods can be problematic are those which retain a bulk dipole moment
1062 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1063 + materials, uniform weighting of the orientational contributions can be
1064 + important for converging the total energy.  In these cases, the
1065 + damping function which causes the non-uniform weighting can be
1066 + replaced by the bare electrostatic kernel, and the energies return to
1067 + the expected converged values.
1068 +
1069 + Based on the results of this work, the GSF method is a suitable and
1070 + efficient replacement for the Ewald sum for evaluating electrostatic
1071 + interactions in MD simulations.  Both methods retain excellent
1072 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1073 + energy drift and fluctuations from the GSF electrostatics are better
1074 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1075 + Because they use real-space cutoffs with moderate cutoff radii, the
1076 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1077 + increases.  Additionally, they can be made extremely efficient using
1078 + spline interpolations of the radial functions.  They require no
1079 + Fourier transforms or $k$-space sums, and guarantee the smooth
1080 + handling of energies, forces, and torques as multipoles cross the
1081 + real-space cutoff boundary.
1082 +
1083 + \begin{acknowledgments}
1084 +  JDG acknowledges helpful discussions with Christopher
1085 +  Fennell. Support for this project was provided by the National
1086 +  Science Foundation under grant CHE-1362211. Computational time was
1087 +  provided by the Center for Research Computing (CRC) at the
1088 +  University of Notre Dame.
1089 + \end{acknowledgments}
1090 +
1091   %\bibliographystyle{aip}
1092   \newpage
1093   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines