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 4170 by gezelter, Wed Jun 4 18:48:27 2014 UTC vs.
Revision 4180 by gezelter, Wed Jun 11 21:03:11 2014 UTC

# Line 36 | Line 36 | preprint,
36   \usepackage{amsmath}
37   \usepackage{times}
38   \usepackage{mathptm}
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 114 | Line 134 | an ordered lattice (e.g. when computing the Madelung c
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
137 > an 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:NaCl}).  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
# Line 189 | Line 209 | swings in the electrostatic energy as the cutoff radiu
209  
210   In ionic crystals, real-space truncation can break the effective
211   multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
212 < swings in the electrostatic energy as the cutoff radius is increased
213 < (or as individual ions move back and forth across the boundary).  This
214 < is why the image charges were necessary for the Wolf sum to exhibit
215 < rapid convergence.  Similarly, the real-space truncation of point
216 < multipole interactions breaks higher order multipole arrangements, and
217 < image multipoles are required for real-space treatments of
198 < electrostatic energies.
212 > swings in the electrostatic energy as individual ions move back and
213 > forth across the boundary.  This is why the image charges are
214 > necessary for the Wolf sum to exhibit rapid convergence.  Similarly,
215 > the real-space truncation of point multipole interactions breaks
216 > higher order multipole arrangements, and image multipoles are required
217 > for real-space treatments of electrostatic energies.
218  
219   % Because of this reason, although the nature of electrostatic
220   % interaction short ranged, the hard cutoff sphere creates very large
# Line 253 | Line 272 | multipoles up to octupolar
272   rough approximation.  Atomic sites can also be assigned point
273   multipoles and polarizabilities to increase the accuracy of the
274   molecular model.  Recently, water has been modeled with point
275 < multipoles up to octupolar
276 < order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
275 > multipoles up to octupolar order using the soft sticky
276 > dipole-quadrupole-octupole (SSDQO)
277 > model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
278   multipoles up to quadrupolar order have also been coupled with point
279   polarizabilities in the high-quality AMOEBA and iAMOEBA water
280 < models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}.  But
280 > models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
281   using point multipole with the real space truncation without
282   accounting for multipolar neutrality will create energy conservation
283   issues in molecular dynamics (MD) simulations.
# Line 297 | Line 317 | where the multipole operator for site $\bf a$,
317   \begin{equation}
318   U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
319   \end{equation}
320 < where the multipole operator for site $\bf a$,
321 < \begin{equation}
322 < \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
323 < +  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.
320 > where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
321 > expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
322 >    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
323 > $\bf a$.
324  
325 < Interactions between multipoles can be expressed as higher derivatives
326 < of the bare Coulomb potential, so one way of ensuring that the forces
327 < and torques vanish at the cutoff distance is to include a larger
328 < number of terms in the truncated Taylor expansion, e.g.,
329 < %
330 < \begin{equation}
331 < f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert  _{R_c}  .
332 < \end{equation}
333 < %
334 < The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
335 < Thus, for $f(r)=1/r$, we find
336 < %
337 < \begin{equation}
338 < 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} .
339 < \end{equation}
340 < This function is an approximate electrostatic potential that has
341 < vanishing second derivatives at the cutoff radius, making it suitable
342 < for shifting the forces and torques of charge-dipole interactions.
325 > % Interactions between multipoles can be expressed as higher derivatives
326 > % of the bare Coulomb potential, so one way of ensuring that the forces
327 > % and torques vanish at the cutoff distance is to include a larger
328 > % number of terms in the truncated Taylor expansion, e.g.,
329 > % %
330 > % \begin{equation}
331 > % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
332 > % \end{equation}
333 > % %
334 > % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
335 > % Thus, for $f(r)=1/r$, we find
336 > % %
337 > % \begin{equation}
338 > % 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} .
339 > % \end{equation}
340 > % This function is an approximate electrostatic potential that has
341 > % vanishing second derivatives at the cutoff radius, making it suitable
342 > % for shifting the forces and torques of charge-dipole interactions.
343  
344 < In general, the TSF potential for any multipole-multipole interaction
345 < can be written
344 > The TSF potential for any multipole-multipole interaction can be
345 > written
346   \begin{equation}
347   U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
348   \label{generic}
349   \end{equation}
350 < with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
351 < charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
352 < $n=4$ for quadrupole-quadrupole.  To ensure smooth convergence of the
353 < energy, force, and torques, the required number of terms from Taylor
354 < series expansion in $f_n(r)$ must be performed for different
355 < multipole-multipole interactions.
350 > where $f_n(r)$ is a shifted kernel that is appropriate for the order
351 > of the interaction, with $n=0$ for charge-charge, $n=1$ for
352 > charge-dipole, $n=2$ for charge-quadrupole and dipole-dipole, $n=3$
353 > for dipole-quadrupole, and $n=4$ for quadrupole-quadrupole.  To ensure
354 > smooth convergence of the energy, force, and torques, a Taylor
355 > expansion with $n$ terms must be performed at cutoff radius ($r_c$) to
356 > obtain $f_n(r)$.
357  
358 < To carry out the same procedure for a damped electrostatic kernel, we
359 < replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
360 < Many of the derivatives of the damped kernel are well known from
361 < Smith's early work on multipoles for the Ewald
362 < summation.\cite{Smith82,Smith98}
358 > % To carry out the same procedure for a damped electrostatic kernel, we
359 > % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
360 > % Many of the derivatives of the damped kernel are well known from
361 > % Smith's early work on multipoles for the Ewald
362 > % summation.\cite{Smith82,Smith98}
363  
364 < Note that increasing the value of $n$ will add additional terms to the
365 < electrostatic potential, e.g., $f_2(r)$ includes orders up to
366 < $(r-R_c)^3/R_c^4$, and so on.  Successive derivatives of the $f_n(r)$
367 < functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
368 < f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
369 < for computing multipole energies, forces, and torques, and smooth
370 < cutoffs of these quantities can be guaranteed as long as the number of
371 < terms in the Taylor series exceeds the derivative order required.
364 > % Note that increasing the value of $n$ will add additional terms to the
365 > % electrostatic potential, e.g., $f_2(r)$ includes orders up to
366 > % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
367 > % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
368 > % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
369 > % for computing multipole energies, forces, and torques, and smooth
370 > % cutoffs of these quantities can be guaranteed as long as the number of
371 > % terms in the Taylor series exceeds the derivative order required.
372  
373   For multipole-multipole interactions, following this procedure results
374 < in separate radial functions for each distinct orientational
375 < contribution to the potential, and ensures that the forces and torques
376 < from {\it each} of these contributions will vanish at the cutoff
377 < radius.  For example, the direct dipole dot product ($\mathbf{D}_{i}
378 < \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
374 > in separate radial functions for each of the distinct orientational
375 > contributions to the potential, and ensures that the forces and
376 > torques from each of these contributions will vanish at the cutoff
377 > radius.  For example, the direct dipole dot product
378 > ($\mathbf{D}_{\bf a}
379 > \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
380   dot products:
381   \begin{equation}
382 < U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
383 < \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
384 < -\frac{1}{4\pi \epsilon_0}
385 < \left( \mathbf{D}_{i} \cdot \hat{r} \right)
386 < \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 <  \frac{g_2(r)}{r} \right)
382 > U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
383 >  \mathbf{D}_{\bf a} \cdot
384 > \mathbf{D}_{\bf b} \right) v_{21}(r) +
385 > \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
386 > \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
387   \end{equation}
388  
389 < The electrostatic forces and torques acting on the central multipole
390 < site due to another site within cutoff sphere are derived from
389 > For the Taylor shifted (TSF) method with the undamped kernel,
390 > $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
391 > \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
392 > - \frac{6}{r r_c^2}$.  In these functions, one can easily see the
393 > connection to unmodified electrostatics as well as the smooth
394 > transition to zero in both these functions as $r\rightarrow r_c$.  The
395 > electrostatic forces and torques acting on the central multipole due
396 > to another site within cutoff sphere are derived from
397   Eq.~\ref{generic}, accounting for the appropriate number of
398   derivatives. Complete energy, force, and torque expressions are
399   presented in the first paper in this series (Reference
400 < \citep{PaperI}).
400 > \onlinecite{PaperI}).
401  
402   \subsection{Gradient-shifted force (GSF)}
403  
404 < A second (and significantly simpler) method involves shifting the
405 < gradient of the raw coulomb potential for each particular multipole
404 > A second (and conceptually simpler) method involves shifting the
405 > gradient of the raw Coulomb potential for each particular multipole
406   order.  For example, the raw dipole-dipole potential energy may be
407   shifted smoothly by finding the gradient for two interacting dipoles
408   which have been projected onto the surface of the cutoff sphere
409   without changing their relative orientation,
410   \begin{displaymath}
411 < U_{D_{i}D_{j}}(r_{ij})  = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
412 <   - (r_{ij}-R_c) \hat{r}_{ij} \cdot
413 <  \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
411 > U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
412 > U_{D_{\bf a} D_{\bf b}}(r_c)
413 >   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
414 >  \vec{\nabla} U_{D_{\bf a}D_{\bf b}}(r) \Big \lvert _{r_c}
415   \end{displaymath}
416 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
417 < and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
418 < the signs are reversed for the dipole that has been projected onto the
419 < cutoff sphere).  In many ways, this simpler approach is closer in
420 < spirit to the original shifted force method, in that it projects a
421 < neutralizing multipole (and the resulting forces from this multipole)
422 < onto a cutoff sphere. The resulting functional forms for the
423 < potentials, forces, and torques turn out to be quite similar in form
424 < to the Taylor-shifted approach, although the radial contributions are
425 < significantly less perturbed by the Gradient-shifted approach than
426 < they are in the Taylor-shifted method.
416 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
417 >  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
418 > (although the signs are reversed for the dipole that has been
419 > projected onto the cutoff sphere).  In many ways, this simpler
420 > approach is closer in spirit to the original shifted force method, in
421 > that it projects a neutralizing multipole (and the resulting forces
422 > from this multipole) onto a cutoff sphere. The resulting functional
423 > forms for the potentials, forces, and torques turn out to be quite
424 > similar in form to the Taylor-shifted approach, although the radial
425 > contributions are significantly less perturbed by the gradient-shifted
426 > approach than they are in the Taylor-shifted method.
427  
428 + For the gradient shifted (GSF) method with the undamped kernel,
429 + $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
430 + $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
431 + Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
432 + because the Taylor expansion retains only one term, they are
433 + significantly less perturbed than the TSF functions.
434 +
435   In general, the gradient shifted potential between a central multipole
436   and any multipolar site inside the cutoff radius is given by,
437   \begin{equation}
# Line 419 | Line 447 | approach zero as $r \rightarrow R_c$.  Both the GSF an
447   function of radius, hence the contribution of the third term is very
448   small for large cutoff radii.  The force and torque derived from
449   equation \ref{generic2} are consistent with the energy expression and
450 < approach zero as $r \rightarrow R_c$.  Both the GSF and TSF methods
450 > approach zero as $r \rightarrow r_c$.  Both the GSF and TSF methods
451   can be considered generalizations of the original DSF method for
452   higher order multipole interactions. GSF and TSF are also identical up
453   to the charge-dipole interaction but generate different expressions in
454   the energy, force and torque for higher order multipole-multipole
455   interactions. Complete energy, force, and torque expressions for the
456   GSF potential are presented in the first paper in this series
457 < (Reference \citep{PaperI})
457 > (Reference~\onlinecite{PaperI})
458  
459  
460   \subsection{Shifted potential (SP) }
# Line 439 | Line 467 | U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
467   interactions with the central multipole and the image. This
468   effectively shifts the total potential to zero at the cutoff radius,
469   \begin{equation}
470 < U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
470 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
471 > U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
472   \label{eq:SP}
473   \end{equation}          
474   where the sum describes separate potential shifting that is done for
475   each orientational contribution to the energy (e.g. the direct dipole
476   product contribution is shifted {\it separately} from the
477   dipole-distance terms in dipole-dipole interactions).  Note that this
478 < is not a simple shifting of the total potential at $R_c$. Each radial
478 > is not a simple shifting of the total potential at $r_c$. Each radial
479   contribution is shifted separately.  One consequence of this is that
480   multipoles that reorient after leaving the cutoff sphere can re-enter
481   the cutoff sphere without perturbing the total energy.
482  
483 < The potential energy between a central multipole and other multipolar
484 < sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
485 < force and torque obtained from the shifted potential (SP) are
486 < discontinuous at $R_c$. Therefore, MD simulations will still
487 < experience energy drift while operating under the SP potential, but it
488 < may be suitable for Monte Carlo approaches where the configurational
489 < energy differences are the primary quantity of interest.
483 > For the shifted potential (SP) method with the undamped kernel,
484 > $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
485 > \frac{3}{r^3} - \frac{3}{r_c^3}$.  The potential energy between a
486 > central multipole and other multipolar sites goes smoothly to zero as
487 > $r \rightarrow r_c$.  However, the force and torque obtained from the
488 > shifted potential (SP) are discontinuous at $r_c$.  MD simulations
489 > will still experience energy drift while operating under the SP
490 > potential, but it may be suitable for Monte Carlo approaches where the
491 > configurational energy differences are the primary quantity of
492 > interest.
493  
494 < \subsection{The Self term}
494 > \subsection{The Self Term}
495   In the TSF, GSF, and SP methods, a self-interaction is retained for
496   the central multipole interacting with its own image on the surface of
497   the cutoff sphere.  This self interaction is nearly identical with the
498   self-terms that arise in the Ewald sum for multipoles.  Complete
499   expressions for the self terms are presented in the first paper in
500 < this series (Reference \citep{PaperI})  
500 > this series (Reference \onlinecite{PaperI}).
501  
502  
503   \section{\label{sec:methodology}Methodology}
# Line 477 | Line 509 | arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} Thi
509   real-space cutoffs.  In the first paper of this series, we compared
510   the dipolar and quadrupolar energy expressions against analytic
511   expressions for ordered dipolar and quadrupolar
512 < arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
513 < multipolar Ewald sum as a reference method for comparing energies,
514 < forces, and torques for molecular models that mimic disordered and
515 < ordered condensed-phase systems.  These test-cases include:
512 > arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
513 > used the multipolar Ewald sum as a reference method for comparing
514 > energies, forces, and torques for molecular models that mimic
515 > disordered and ordered condensed-phase systems.  The parameters used
516 > in the test cases are given in table~\ref{tab:pars}.
517  
518 < \begin{itemize}
519 < \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
520 < \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
521 < \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
522 < \item Soft Quadrupolar solids  ($\sigma = , \epsilon = , Q_{xx} = ...$)
523 < \item A mixed multipole model for water
524 < \item A mixed multipole models for water with dissolved ions
525 < \end{itemize}
526 < This last test case exercises all levels of the multipole-multipole
527 < interactions we have derived so far and represents the most complete
528 < test of the new methods.
518 > \begin{table}
519 > \label{tab:pars}
520 > \caption{The parameters used in the systems used to evaluate the new
521 >  real-space methods.  The most comprehensive test was a liquid
522 >  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
523 >  ions).  This test excercises all orders of the multipolar
524 >  interactions developed in the first paper.}
525 > \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
526 >             & \multicolumn{2}{c|}{LJ parameters} &
527 >             \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
528 > Test system & $\sigma$& $\epsilon$ & $C$ & $D$  &
529 > $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass  & $I_{xx}$ & $I_{yy}$ &
530 > $I_{zz}$ \\ \cline{6-8}\cline{10-12}
531 > & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
532 > \AA\textsuperscript{2})} \\ \hline
533 >    Soft Dipolar fluid & 3.051 & 0.152 &  & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
534 >    Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
535 > Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
536 > Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
537 >      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
538 >              \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
539 >              \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
540 > \end{tabularx}
541 > \end{table}
542 > The systems consist of pure multipolar solids (both dipole and
543 > quadrupole), pure multipolar liquids (both dipole and quadrupole), a
544 > fluid composed of sites containing both dipoles and quadrupoles
545 > simultaneously, and a final test case that includes ions with point
546 > charges in addition to the multipolar fluid.  The solid-phase
547 > parameters were chosen so that the systems can explore some
548 > orientational freedom for the multipolar sites, while maintaining
549 > relatively strict translational order.  The SSDQ model used here is
550 > not a particularly accurate water model, but it does test
551 > dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
552 > interactions at roughly the same magnitudes. The last test case, SSDQ
553 > water with dissolved ions, exercises \textit{all} levels of the
554 > multipole-multipole interactions we have derived so far and represents
555 > the most complete test of the new methods.
556  
557   In the following section, we present results for the total
558   electrostatic energy, as well as the electrostatic contributions to
559   the force and torque on each molecule.  These quantities have been
560   computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
561 < and have been compared with the values obtaine from the multipolar
562 < Ewald sum.  In Mote Carlo (MC) simulations, the energy differences
561 > and have been compared with the values obtained from the multipolar
562 > Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
563   between two configurations is the primary quantity that governs how
564   the simulation proceeds. These differences are the most imporant
565   indicators of the reliability of a method even if the absolute
# Line 510 | Line 570 | contributions to the forces and torques.
570   behavior of the simulation, so we also compute the electrostatic
571   contributions to the forces and torques.
572  
573 < \subsection{Model systems}
574 < To sample independent configurations of multipolar crystals, a body
575 < centered cubic (BCC) crystal which is a minimum energy structure for
576 < point dipoles was generated using 3,456 molecules.  The multipoles
577 < were translationally locked in their respective crystal sites for
578 < equilibration at a relatively low temperature (50K), so that dipoles
579 < or quadrupoles could freely explore all accessible orientations.  The
580 < 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.
573 > \subsection{Implementation}
574 > The real-space methods developed in the first paper in this series
575 > have been implemented in our group's open source molecular simulation
576 > program, OpenMD,\cite{openmd} which was used for all calculations in
577 > this work.  The complementary error function can be a relatively slow
578 > function on some processors, so all of the radial functions are
579 > precomputed on a fine grid and are spline-interpolated to provide
580 > values when required.  
581  
582 < For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
583 < system was created with 2048 molecules oriented randomly.  These were
582 > Using the same simulation code, we compare to a multipolar Ewald sum
583 > with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
584 > the Ewald sum is a re-implementation of the algorithm originally
585 > proposed by Smith that does not use the particle mesh or smoothing
586 > approximations.\cite{Smith82,Smith98} In all cases, the quantities
587 > being compared are the electrostatic contributions to energies, force,
588 > and torques.  All other contributions to these quantities (i.e. from
589 > Lennard-Jones interactions) are removed prior to the comparisons.
590  
591 < system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
592 < temperature after equilibration.  We collected 250 different
593 < configurations in equal interval of time. For the ions mixed liquid
594 < system, we converted 48 different molecules into 24 \ce{Na+} and 24
595 < \ce{Cl-} ions and equilibrated. After equilibration, the system was run
596 < at the same environment for 1ns and 250 configurations were
597 < collected. While comparing energies, forces, and torques with Ewald
598 < method, Lennard-Jones potentials were turned off and purely
599 < electrostatic interaction had been compared.
591 > The convergence parameter ($\alpha$) also plays a role in the balance
592 > of the real-space and reciprocal-space portions of the Ewald
593 > calculation.  Typical molecular mechanics packages set this to a value
594 > that depends on the cutoff radius and a tolerance (typically less than
595 > $1 \times 10^{-4}$ kcal/mol).  Smaller tolerances are typically
596 > associated with increasing accuracy at the expense of computational
597 > time spent on the reciprocal-space portion of the
598 > summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
599 > 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
600 > Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
601  
602 + The real-space models have self-interactions that provide
603 + contributions to the energies only.  Although the self interaction is
604 + a rapid calculation, we note that in systems with fluctuating charges
605 + or point polarizabilities, the self-term is not static and must be
606 + recomputed at each time step.
607 +
608 + \subsection{Model systems}
609 + To sample independent configurations of the multipolar crystals, body
610 + centered cubic (bcc) crystals, which exhibit the minimum energy
611 + structures for point dipoles, were generated using 3,456 molecules.
612 + The multipoles were translationally locked in their respective crystal
613 + sites for equilibration at a relatively low temperature (50K) so that
614 + dipoles or quadrupoles could freely explore all accessible
615 + orientations.  The translational constraints were then removed, the
616 + systems were re-equilibrated, and the crystals were simulated for an
617 + additional 10 ps in the microcanonical (NVE) ensemble with an average
618 + temperature of 50 K.  The balance between moments of inertia and
619 + particle mass were chosen to allow orientational sampling without
620 + significant translational motion.  Configurations were sampled at
621 + equal time intervals in order to compare configurational energy
622 + differences.  The crystals were simulated far from the melting point
623 + in order to avoid translational deformation away of the ideal lattice
624 + geometry.
625 +
626 + For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
627 + simulations, each system was created with 2,048 randomly-oriented
628 + molecules.  These were equilibrated at a temperature of 300K for 1 ns.
629 + Each system was then simulated for 1 ns in the microcanonical (NVE)
630 + ensemble.  We collected 250 different configurations at equal time
631 + intervals. For the liquid system that included ionic species, we
632 + converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
633 + \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
634 + run under the same conditions for 1 ns. A total of 250 configurations
635 + were collected. In the following comparisons of energies, forces, and
636 + torques, the Lennard-Jones potentials were turned off and only the
637 + purely electrostatic quantities were compared with the same values
638 + obtained via the Ewald sum.
639 +
640   \subsection{Accuracy of Energy Differences, Forces and Torques}
641   The pairwise summation techniques (outlined above) were evaluated for
642   use in MC simulations by studying the energy differences between
# Line 550 | Line 649 | we used least square regressions analysiss for the six
649   should be identical for all methods.
650  
651   Since none of the real-space methods provide exact energy differences,
652 < we used least square regressions analysiss for the six different
652 > we used least square regressions analysis for the six different
653   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
654   with the multipolar Ewald reference method.  Unitary results for both
655   the correlation (slope) and correlation coefficient for these
# Line 561 | Line 660 | also been compared by using least squares regression a
660   configurations and 250 configurations were recorded for comparison.
661   Each system provided 31,125 energy differences for a total of 186,750
662   data points.  Similarly, the magnitudes of the forces and torques have
663 < also been compared by using least squares regression analyses. In the
663 > also been compared using least squares regression analysis. In the
664   forces and torques comparison, the magnitudes of the forces acting in
665   each molecule for each configuration were evaluated. For example, our
666   dipolar liquid simulation contains 2048 molecules and there are 250
# Line 647 | Line 746 | model must allow for long simulation times with minima
746        
747   %        \label{fig:barGraph2}
748   %      \end{figure}
749 < %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.
749 > %The correlation coefficient ($R^2$) and slope of the linear
750 > %regression plots for the energy differences for all six different
751 > %molecular systems is shown in figure 4a and 4b.The plot shows that
752 > %the correlation coefficient improves for the SP cutoff method as
753 > %compared to the undamped hard cutoff method in the case of SSDQC,
754 > %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
755 > %crystal and liquid, the correlation coefficient is almost unchanged
756 > %and close to 1.  The correlation coefficient is smallest (0.696276
757 > %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
758 > %charge-charge and charge-multipole interactions. Since the
759 > %charge-charge and charge-multipole interaction is long ranged, there
760 > %is huge deviation of correlation coefficient from 1. Similarly, the
761 > %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
762 > %compared to interactions in the other multipolar systems, thus the
763 > %correlation coefficient very close to 1 even for hard cutoff
764 > %method. The idea of placing image multipole on the surface of the
765 > %cutoff sphere improves the correlation coefficient and makes it close
766 > %to 1 for all types of multipolar systems. Similarly the slope is
767 > %hugely deviated from the correct value for the lower order
768 > %multipole-multipole interaction and slightly deviated for higher
769 > %order multipole – multipole interaction. The SP method improves both
770 > %correlation coefficient ($R^2$) and slope significantly in SSDQC and
771 > %dipolar systems.  The Slope is found to be deviated more in dipolar
772 > %crystal as compared to liquid which is associated with the large
773 > %fluctuation in the electrostatic energy in crystal. The GSF also
774 > %produced better values of correlation coefficient and slope with the
775 > %proper selection of the damping alpha (Interested reader can consult
776 > %accompanying supporting material). The TSF method gives good value of
777 > %correlation coefficient for the dipolar crystal, dipolar liquid,
778 > %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
779 > %regression slopes are significantly deviated.
780 >
781   \begin{figure}
782 <        \centering
783 <        \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
784 <        \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)}
785 <        \label{fig:slopeCorr_energy}
786 <    \end{figure}
787 < 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.
782 >  \centering
783 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
784 >  \caption{Statistical analysis of the quality of configurational
785 >    energy differences for the real-space electrostatic methods
786 >    compared with the reference Ewald sum.  Results with a value equal
787 >    to 1 (dashed line) indicate $\Delta E$ values indistinguishable
788 >    from those obtained using the multipolar Ewald sum.  Different
789 >    values of the cutoff radius are indicated with different symbols
790 >    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
791 >    triangles).}
792 >  \label{fig:slopeCorr_energy}
793 > \end{figure}
794 >
795 > The combined correlation coefficient and slope for all six systems is
796 > shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
797 > reproduce the Ewald configurational energy differences with remarkable
798 > fidelity.  Undamped hard cutoffs introduce a significant amount of
799 > random scatter in the energy differences which is apparent in the
800 > reduced value of the correlation coefficient for this method.  This
801 > can be easily understood as configurations which exhibit small
802 > traversals of a few dipoles or quadrupoles out of the cutoff sphere
803 > will see large energy jumps when hard cutoffs are used.  The
804 > orientations of the multipoles (particularly in the ordered crystals)
805 > mean that these energy jumps can go in either direction, producing a
806 > significant amount of random scatter, but no systematic error.
807 >
808 > The TSF method produces energy differences that are highly correlated
809 > with the Ewald results, but it also introduces a significant
810 > systematic bias in the values of the energies, particularly for
811 > smaller cutoff values. The TSF method alters the distance dependence
812 > of different orientational contributions to the energy in a
813 > non-uniform way, so the size of the cutoff sphere can have a large
814 > effect, particularly for the crystalline systems.
815 >
816 > Both the SP and GSF methods appear to reproduce the Ewald results with
817 > excellent fidelity, particularly for moderate damping ($\alpha =
818 > 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
819 > 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
820 > method with short cutoffs, all of the methods would be appropriate for
821 > use in Monte Carlo simulations.
822 >
823   \subsection{Magnitude of the force and torque vectors}
824 < 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.
824 >
825 > The comparisons of the magnitudes of the forces and torques for the
826 > data accumulated from all six systems are shown in Figures
827 > ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
828 > correlation and slope for the forces agree well with the Ewald sum
829 > even for the hard cutoffs.
830 >
831 > For systems of molecules with only multipolar interactions, the pair
832 > energy contributions are quite short ranged.  Moreover, the force
833 > decays more rapidly than the electrostatic energy, hence the hard
834 > cutoff method can also produce reasonable agreement for this quantity.
835 > Although the pure cutoff gives reasonably good electrostatic forces
836 > for pairs of molecules included within each other's cutoff spheres,
837 > the discontinuity in the force at the cutoff radius can potentially
838 > cause energy conservation problems as molecules enter and leave the
839 > cutoff spheres.  This is discussed in detail in section
840 > \ref{sec:conservation}.
841 >
842 > The two shifted-force methods (GSF and TSF) exhibit a small amount of
843 > systematic variation and scatter compared with the Ewald forces.  The
844 > shifted-force models intentionally perturb the forces between pairs of
845 > molecules inside each other's cutoff spheres in order to correct the
846 > energy conservation issues, and this perturbation is evident in the
847 > statistics accumulated for the molecular forces.  The GSF
848 > perturbations are minimal, particularly for moderate damping and
849 > commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
850 > reasonable agreement in the correlation coefficient but again the
851 > systematic error in the forces is concerning if replication of Ewald
852 > forces is desired.
853 >
854   \begin{figure}
855 <        \centering
856 <        \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
857 <        \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). }
858 <        \label{fig:slopeCorr_force}
859 <    \end{figure}
860 < 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.
855 >  \centering
856 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
857 >  \caption{Statistical analysis of the quality of the force vector
858 >    magnitudes for the real-space electrostatic methods compared with
859 >    the reference Ewald sum. Results with a value equal to 1 (dashed
860 >    line) indicate force magnitude values indistinguishable from those
861 >    obtained using the multipolar Ewald sum.  Different values of the
862 >    cutoff radius are indicated with different symbols (9\AA\ =
863 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
864 >  \label{fig:slopeCorr_force}
865 > \end{figure}
866 >
867 >
868   \begin{figure}
869 <        \centering
870 <        \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
871 <        \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).}
872 <        \label{fig:slopeCorr_torque}
873 <    \end{figure}
869 >  \centering
870 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
871 >  \caption{Statistical analysis of the quality of the torque vector
872 >    magnitudes for the real-space electrostatic methods compared with
873 >    the reference Ewald sum. Results with a value equal to 1 (dashed
874 >    line) indicate force magnitude values indistinguishable from those
875 >    obtained using the multipolar Ewald sum.  Different values of the
876 >    cutoff radius are indicated with different symbols (9\AA\ =
877 >    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
878 >  \label{fig:slopeCorr_torque}
879 > \end{figure}
880 >
881 > The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
882 > significantly influenced by the choice of real-space method.  The
883 > torque expressions have the same distance dependence as the energies,
884 > which are naturally longer-ranged expressions than the inter-site
885 > forces.  Torques are also quite sensitive to orientations of
886 > neighboring molecules, even those that are near the cutoff distance.
887 >
888 > The results shows that the torque from the hard cutoff method
889 > reproduces the torques in quite good agreement with the Ewald sum.
890 > The other real-space methods can cause some deviations, but excellent
891 > agreement with the Ewald sum torques is recovered at moderate values
892 > of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
893 > radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
894 > in the slope when compared with the Ewald torques even for larger
895 > cutoff radii.  It appears that the severity of the perturbations in
896 > the TSF method are most in evidence for the torques.
897 >
898   \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 $r_c$.
899  
900 < \begin{figure}
901 <        \centering
902 <        \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
903 <        \caption{The circular variance of the data sets of the direction of the  force and torque vectors obtained from a given method about reference Ewald method. The result equal to 0 (dashed line) indicates direction of the vectors are 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)}
904 <        \label{fig:slopeCorr_circularVariance}
905 <    \end{figure}
906 < \subsection{Total energy conservation}
907 < We have tested the conservation of energy in the SSDQC liquid system by running system for 1ns in the Hard, SP, GSF and TSF method. The Hard cutoff method shows very high energy drifts 433.53 KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol (measured by the SD from the slope) for the undamped case, which makes it completely unusable in MD simulations. The SP method also shows large value of energy drift 1.289 Kcal/Mol/ns/particle and energy fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method is due to the non-vanishing nature of the torque and force at the cutoff radius. We can improve the energy conservation in some extent by the proper selection of the damping alpha but the improvement is not good enough, which can be observed in Figure 9a and 9b .The GSF and TSF shows very low value of energy drift 0.09016, 0.07371 KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol respectively for the undamped case. Since the absolute value of the evaluated electrostatic energy, force and torque from TSF method are deviated from the Ewald, it does not mimic MD simulations appropriately. The electrostatic energy, force and torque from the GSF method have very good agreement with the Ewald. In addition, the energy drift and energy fluctuation from the GSF method is much better than Ewald’s method for reciprocal space vector value ($k_f$) equal to 7 as shown in Figure~\ref{fig:energyDrift} and ~\ref{fig:fluctuation}. We can improve the total energy fluctuation and drift for the Ewald’s method by increasing size of the reciprocal space, which extremely increseses the simulation time. In our current simulation, the simulation time for the Hard, SP, and GSF methods are about 5.5 times faster than the Ewald method.
908 < \begin{figure}
909 <        \centering
910 <        \includegraphics[width=0.5 \textwidth]{log(energyDrift)-crop.pdf}
911 < \label{fig:energyDrift}        
912 <        \end{figure}
913 < \begin{figure}
914 <        \centering
915 <        \includegraphics[width=0.5 \textwidth]{logSD-crop.pdf}      
916 <        \caption{The plot showing (a) standard deviation, and (b) total energy drift in the total energy conservation plot for different values of the damping alpha for different cut off methods. }
917 <        \label{fig:fluctuation}
918 <    \end{figure}
900 > The accurate evaluation of force and torque directions is just as
901 > important for molecular dynamics simulations as the magnitudes of
902 > these quantities. Force and torque vectors for all six systems were
903 > analyzed using Fisher statistics, and the quality of the vector
904 > directionality is shown in terms of circular variance
905 > ($\mathrm{Var}(\theta)$) in figure
906 > \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
907 > from the new real-space methods exhibit nearly-ideal Fisher probability
908 > distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
909 > exhibit the best vectorial agreement with the Ewald sum. The force and
910 > torque vectors from GSF method also show good agreement with the Ewald
911 > method, which can also be systematically improved by using moderate
912 > damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
913 > 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
914 > to a distribution with 95\% of force vectors within $6.37^\circ$ of
915 > the corresponding Ewald forces. The TSF method produces the poorest
916 > agreement with the Ewald force directions.
917 >
918 > Torques are again more perturbed than the forces by the new real-space
919 > methods, but even here the variance is reasonably small.  For the same
920 > method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
921 > the circular variance was 0.01415, corresponds to a distribution which
922 > has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
923 > results. Again, the direction of the force and torque vectors can be
924 > systematically improved by varying $\alpha$ and $r_c$.
925 >
926 > \begin{figure}
927 >  \centering
928 >  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
929 >  \caption{The circular variance of the direction of the force and
930 >    torque vectors obtained from the real-space methods around the
931 >    reference Ewald vectors. A variance equal to 0 (dashed line)
932 >    indicates direction of the force or torque vectors are
933 >    indistinguishable from those obtained from the Ewald sum. Here
934 >    different symbols represent different values of the cutoff radius
935 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
936 >  \label{fig:slopeCorr_circularVariance}
937 > \end{figure}
938 >
939 > \subsection{Energy conservation\label{sec:conservation}}
940 >
941 > We have tested the conservation of energy one can expect to see with
942 > the new real-space methods using the SSDQ water model with a small
943 > fraction of solvated ions. This is a test system which exercises all
944 > orders of multipole-multipole interactions derived in the first paper
945 > in this series and provides the most comprehensive test of the new
946 > methods.  A liquid-phase system was created with 2000 water molecules
947 > and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
948 > temperature of 300K.  After equilibration, this liquid-phase system
949 > was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
950 > a cutoff radius of 12\AA.  The value of the damping coefficient was
951 > also varied from the undamped case ($\alpha = 0$) to a heavily damped
952 > case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
953 > sample was also run using the multipolar Ewald sum with the same
954 > real-space cutoff.
955 >
956 > In figure~\ref{fig:energyDrift} we show the both the linear drift in
957 > energy over time, $\delta E_1$, and the standard deviation of energy
958 > fluctuations around this drift $\delta E_0$.  Both of the
959 > shifted-force methods (GSF and TSF) provide excellent energy
960 > conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
961 > while the hard cutoff is essentially unusable for molecular dynamics.
962 > SP provides some benefit over the hard cutoff because the energetic
963 > jumps that happen as particles leave and enter the cutoff sphere are
964 > somewhat reduced, but like the Wolf method for charges, the SP method
965 > would not be as useful for molecular dynamics as either of the
966 > shifted-force methods.
967 >
968 > We note that for all tested values of the cutoff radius, the new
969 > real-space methods can provide better energy conservation behavior
970 > than the multipolar Ewald sum, even when utilizing a relatively large
971 > $k$-space cutoff values.
972 >
973 > \begin{figure}
974 >  \centering
975 >  \includegraphics[width=\textwidth]{newDrift_12.pdf}
976 > \label{fig:energyDrift}        
977 > \caption{Analysis of the energy conservation of the real-space
978 >  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
979 >  energy over time (in kcal / mol / particle / ns) and $\delta
980 >  \mathrm{E}_0$ is the standard deviation of energy fluctuations
981 >  around this drift (in kcal / mol / particle).  All simulations were
982 >  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
983 >  300 K starting from the same initial configuration. All runs
984 >  utilized the same real-space cutoff, $r_c = 12$\AA.}
985 > \end{figure}
986 >
987 >
988   \section{CONCLUSION}
989 < 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.
990 < 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.
991 < In MD simulations, the energy conservation is very important. The
992 < conservation of the total energy can be ensured by  i) enforcing the
993 < smooth truncation of the energy, force and torque in the cutoff radius
994 < and ii) making the energy, force and torque consistent with each
995 < other. The GSF and TSF methods ensure the consistency and smooth
996 < truncation of the energy, force and torque at the cutoff radius, as a
997 < result show very good total energy conservation. But the TSF method
705 < does not show good agreement in the absolute value of the
706 < electrostatic energy, force and torque with the Ewald.  The GSF method
707 < has mimicked Ewald’s force, energy and torque accurately and also
708 < conserved energy. Therefore, the GSF method is the suitable method for
709 < evaluating required force field in MD simulations. In addition, the
710 < energy drift and fluctuation from the GSF method is much better than
711 < Ewald’s method for finite-sized reciprocal space.
989 > In the first paper in this series, we generalized the
990 > charge-neutralized electrostatic energy originally developed by Wolf
991 > \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
992 > up to quadrupolar order.  The SP method is essentially a
993 > multipole-capable version of the Wolf model.  The SP method for
994 > multipoles provides excellent agreement with Ewald-derived energies,
995 > forces and torques, and is suitable for Monte Carlo simulations,
996 > although the forces and torques retain discontinuities at the cutoff
997 > distance that prevents its use in molecular dynamics.
998  
999 < Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
1000 < that can be made extremely efficient using spline interpolations of
1001 < the radial functions.  They require no Fourier transforms or $k$-space
1002 < sums, and guarantee the smooth handling of energies, forces, and
1003 < torques as multipoles cross the real-space cutoff boundary.  
999 > We also developed two natural extensions of the damped shifted-force
1000 > (DSF) model originally proposed by Fennel and
1001 > Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1002 > smooth truncation of energies, forces, and torques at the real-space
1003 > cutoff, and both converge to DSF electrostatics for point-charge
1004 > interactions.  The TSF model is based on a high-order truncated Taylor
1005 > expansion which can be relatively perturbative inside the cutoff
1006 > sphere.  The GSF model takes the gradient from an images of the
1007 > interacting multipole that has been projected onto the cutoff sphere
1008 > to derive shifted force and torque expressions, and is a significantly
1009 > more gentle approach.
1010  
1011 + Of the two newly-developed shifted force models, the GSF method
1012 + produced quantitative agreement with Ewald energy, force, and torques.
1013 + It also performs well in conserving energy in MD simulations.  The
1014 + Taylor-shifted (TSF) model provides smooth dynamics, but these take
1015 + place on a potential energy surface that is significantly perturbed
1016 + from Ewald-based electrostatics.  
1017 +
1018 + % The direct truncation of any electrostatic potential energy without
1019 + % multipole neutralization creates large fluctuations in molecular
1020 + % simulations.  This fluctuation in the energy is very large for the case
1021 + % of crystal because of long range of multipole ordering (Refer paper
1022 + % I).\cite{PaperI} This is also significant in the case of the liquid
1023 + % because of the local multipole ordering in the molecules. If the net
1024 + % multipole within cutoff radius neutralized within cutoff sphere by
1025 + % placing image multiples on the surface of the sphere, this fluctuation
1026 + % in the energy reduced significantly. Also, the multipole
1027 + % neutralization in the generalized SP method showed very good agreement
1028 + % with the Ewald as compared to direct truncation for the evaluation of
1029 + % the $\triangle E$ between the configurations.  In MD simulations, the
1030 + % energy conservation is very important. The conservation of the total
1031 + % energy can be ensured by i) enforcing the smooth truncation of the
1032 + % energy, force and torque in the cutoff radius and ii) making the
1033 + % energy, force and torque consistent with each other. The GSF and TSF
1034 + % methods ensure the consistency and smooth truncation of the energy,
1035 + % force and torque at the cutoff radius, as a result show very good
1036 + % total energy conservation. But the TSF method does not show good
1037 + % agreement in the absolute value of the electrostatic energy, force and
1038 + % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1039 + % energy and torque accurately and also conserved energy.
1040 +
1041 + The only cases we have found where the new GSF and SP real-space
1042 + methods can be problematic are those which retain a bulk dipole moment
1043 + at large distances (e.g. the $Z_1$ dipolar lattice).  In ferroelectric
1044 + materials, uniform weighting of the orientational contributions can be
1045 + important for converging the total energy.  In these cases, the
1046 + damping function which causes the non-uniform weighting can be
1047 + replaced by the bare electrostatic kernel, and the energies return to
1048 + the expected converged values.
1049 +
1050 + Based on the results of this work, the GSF method is a suitable and
1051 + efficient replacement for the Ewald sum for evaluating electrostatic
1052 + interactions in MD simulations.  Both methods retain excellent
1053 + fidelity to the Ewald energies, forces and torques.  Additionally, the
1054 + energy drift and fluctuations from the GSF electrostatics are better
1055 + than a multipolar Ewald sum for finite-sized reciprocal spaces.
1056 + Because they use real-space cutoffs with moderate cutoff radii, the
1057 + GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1058 + increases.  Additionally, they can be made extremely efficient using
1059 + spline interpolations of the radial functions.  They require no
1060 + Fourier transforms or $k$-space sums, and guarantee the smooth
1061 + handling of energies, forces, and torques as multipoles cross the
1062 + real-space cutoff boundary.
1063 +
1064 + \begin{acknowledgments}
1065 +  JDG acknowledges helpful discussions with Christopher
1066 +  Fennell. Support for this project was provided by the National
1067 +  Science Foundation under grant CHE-1362211. Computational time was
1068 +  provided by the Center for Research Computing (CRC) at the
1069 +  University of Notre Dame.
1070 + \end{acknowledgments}
1071 +
1072   %\bibliographystyle{aip}
1073   \newpage
1074   \bibliography{references}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines