ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4180
Committed: Wed Jun 11 21:03:11 2014 UTC (10 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 61360 byte(s)
Log Message:
Nearing the end

File Contents

# User Rev Content
1 mlamichh 4114 % ****** Start of file aipsamp.tex ******
2     %
3     % This file is part of the AIP files in the AIP distribution for REVTeX 4.
4     % Version 4.1 of REVTeX, October 2009
5     %
6     % Copyright (c) 2009 American Institute of Physics.
7     %
8     % See the AIP README file for restrictions and more information.
9     %
10     % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
11     % as well as the rest of the prerequisites for REVTeX 4.1
12     %
13     % It also requires running BibTeX. The commands are as follows:
14     %
15     % 1) latex aipsamp
16     % 2) bibtex aipsamp
17     % 3) latex aipsamp
18     % 4) latex aipsamp
19     %
20     % Use this file as a source of example code for your aip document.
21     % Use the file aiptemplate.tex as a template for your document.
22     \documentclass[%
23 gezelter 4167 aip,jcp,
24 mlamichh 4114 amsmath,amssymb,
25 gezelter 4167 preprint,
26     %reprint,%
27 mlamichh 4114 %author-year,%
28     %author-numerical,%
29     ]{revtex4-1}
30    
31     \usepackage{graphicx}% Include figure files
32     \usepackage{dcolumn}% Align table columns on decimal point
33 gezelter 4167 %\usepackage{bm}% bold math
34 mlamichh 4114 %\usepackage[mathlines]{lineno}% Enable numbering of text and display math
35     %\linenumbers\relax % Commence numbering lines
36     \usepackage{amsmath}
37 gezelter 4168 \usepackage{times}
38     \usepackage{mathptm}
39 gezelter 4175 \usepackage{tabularx}
40 gezelter 4167 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
41     \usepackage{url}
42     \usepackage[english]{babel}
43 mlamichh 4114
44 gezelter 4175 \newcolumntype{Y}{>{\centering\arraybackslash}X}
45 gezelter 4167
46 mlamichh 4114 \begin{document}
47    
48 gezelter 4175 %\preprint{AIP/123-QED}
49 mlamichh 4114
50 gezelter 4175 \title{Real space alternatives to the Ewald
51     Sum. II. Comparison of Methods} % Force line breaks with \\
52 mlamichh 4114
53     \author{Madan Lamichhane}
54     \affiliation{Department of Physics, University
55     of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
56    
57     \author{Kathie E. Newman}
58     \affiliation{Department of Physics, University
59     of Notre Dame, Notre Dame, IN 46556}
60    
61     \author{J. Daniel Gezelter}%
62     \email{gezelter@nd.edu.}
63     \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
64     }%
65    
66     \date{\today}% It is always \today, today,
67     % but any date may be explicitly specified
68    
69     \begin{abstract}
70 gezelter 4175 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 mlamichh 4114 \end{abstract}
90    
91 gezelter 4175 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92 mlamichh 4114 % Classification Scheme.
93 gezelter 4167 \keywords{Electrostatics, Multipoles, Real-space}
94    
95 mlamichh 4114 \maketitle
96    
97    
98 mlamichh 4166 \section{\label{sec:intro}Introduction}
99 gezelter 4167 Computing the interactions between electrostatic sites is one of the
100     most expensive aspects of molecular simulations, which is why there
101     have been significant efforts to develop practical, efficient and
102     convergent methods for handling these interactions. Ewald's method is
103     perhaps the best known and most accurate method for evaluating
104     energies, forces, and torques in explicitly-periodic simulation
105     cells. In this approach, the conditionally convergent electrostatic
106     energy is converted into two absolutely convergent contributions, one
107     which is carried out in real space with a cutoff radius, and one in
108     reciprocal space.\cite{Clarke:1986eu,Woodcock75}
109 mlamichh 4114
110 gezelter 4167 When carried out as originally formulated, the reciprocal-space
111     portion of the Ewald sum exhibits relatively poor computational
112     scaling, making it prohibitive for large systems. By utilizing
113     particle meshes and three dimensional fast Fourier transforms (FFT),
114     the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
115 gezelter 4168 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
116 gezelter 4167 the computational cost from $O(N^2)$ down to $O(N \log
117     N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
118    
119     Because of the artificial periodicity required for the Ewald sum, the
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 gezelter 4180 To simulate interfacial systems, Parry's extension of the 3D Ewald sum
124 gezelter 4167 is appropriate for slab geometries.\cite{Parry:1975if} The inherent
125     periodicity in the Ewald’s method can also be problematic for
126     interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
127     methods that were developed to handle two-dimensional (2D)
128     electrostatic interactions in interfacial systems have not had similar
129     particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
130     Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
131    
132 mlamichh 4166 \subsection{Real-space methods}
133 gezelter 4168 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
134     method for calculating electrostatic interactions between point
135 gezelter 4167 charges. They argued that the effective Coulomb interaction in
136     condensed systems is actually short ranged.\cite{Wolf92,Wolf95}. For
137 gezelter 4180 an ordered lattice (e.g., when computing the Madelung constant of an
138 gezelter 4167 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 gezelter 4180 $r^{-5}$ (see figure \ref{fig:NaCl}). For this reason, careful
142 gezelter 4167 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 mlamichh 4114
148     \begin{figure}[h!]
149 gezelter 4167 \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}
158     \end{figure}
159 mlamichh 4114
160 gezelter 4167 The direct truncation of interactions at a cutoff radius creates
161     truncation defects. Wolf \textit{et al.} further argued that
162     truncation errors are due to net charge remaining inside the cutoff
163     sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
164     placing an image charge on the surface of the cutoff sphere for every
165     real charge inside the cutoff. These charges are present for the
166     evaluation of both the pair interaction energy and the force, although
167     the force expression maintained a discontinuity at the cutoff sphere.
168     In the original Wolf formulation, the total energy for the charge and
169     image were not equal to the integral of their force expression, and as
170     a result, the total energy would not be conserved in molecular
171     dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
172     Fennel and Gezelter later proposed shifted force variants of the Wolf
173     method with commensurate force and energy expressions that do not
174 gezelter 4168 exhibit this problem.\cite{Fennell:2006lq} Related real-space
175     methods were also proposed by Chen \textit{et
176     al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
177     and by Wu and Brooks.\cite{Wu:044107}
178 gezelter 4167
179     Considering the interaction of one central ion in an ionic crystal
180     with a portion of the crystal at some distance, the effective Columbic
181     potential is found to be decreasing as $r^{-5}$. If one views the
182     \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
183     \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
184     rapidly to the Madelung energy than the dipolar
185     approximation.\cite{Wolf92} To find the correct Madelung constant,
186     Lacman suggested that the NaCl structure could be constructed in a way
187     that the finite crystal terminates with complete \ce{(NaCl)4}
188     molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
189     by opposite charges. Similarly for each pair of charges, there is an
190     opposite pair of charge adjacent to it. The central ion sees what is
191     effectively a set of octupoles at large distances. These facts suggest
192     that the Madelung constants are relatively short ranged for perfect
193     ionic crystals.\cite{Wolf:1999dn}
194    
195     One can make a similar argument for crystals of point multipoles. The
196     Luttinger and Tisza treatment of energy constants for dipolar lattices
197     utilizes 24 basis vectors that contain dipoles at the eight corners of
198     a unit cube. Only three of these basis vectors, $X_1, Y_1,
199     \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
200     zero net dipole and retain contributions only from higher order
201     multipoles. The effective interaction between a dipole at the center
202     of a crystal and a group of eight dipoles farther away is
203     significantly shorter ranged than the $r^{-3}$ that one would expect
204     for raw dipole-dipole interactions. Only in crystals which retain a
205     bulk dipole moment (e.g. ferroelectrics) does the analogy with the
206     ionic crystal break down -- ferroelectric dipolar crystals can exist,
207     while ionic crystals with net charge in each unit cell would be
208     unstable.
209    
210     In ionic crystals, real-space truncation can break the effective
211 gezelter 4168 multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
212 gezelter 4180 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 gezelter 4167
219     % Because of this reason, although the nature of electrostatic
220     % interaction short ranged, the hard cutoff sphere creates very large
221     % fluctuation in the electrostatic energy for the perfect crystal. In
222     % addition, the charge neutralized potential proposed by Wolf et
223     % al. converged to correct Madelung constant but still holds oscillation
224     % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
225     % oscillation in the energy around its fully converged value can be due
226     % to the non-neutralized value of the higher order moments within the
227     % cutoff sphere.
228    
229     The forces and torques acting on atomic sites are the fundamental
230     factors driving dynamics in molecular simulations. Fennell and
231     Gezelter proposed the damped shifted force (DSF) energy kernel to
232     obtain consistent energies and forces on the atoms within the cutoff
233 gezelter 4168 sphere. Both the energy and the force go smoothly to zero as an atom
234     aproaches the cutoff radius. The comparisons of the accuracy these
235     quantities between the DSF kernel and SPME was surprisingly
236     good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
237     calculating electrostatic interactions in molecular systems with
238     relatively uniform charge
239     densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
240 gezelter 4167
241 gezelter 4168 \subsection{The damping function}
242 gezelter 4167 The damping function used in our research has been discussed in detail
243     in the first paper of this series.\cite{PaperI} The radial kernel
244 gezelter 4168 $1/r$ for the interactions between point charges can be replaced by
245     the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
246     accelerate the rate of convergence, where $\alpha$ is a damping
247     parameter with units of inverse distance. Altering the value of
248     $\alpha$ is equivalent to changing the width of Gaussian charge
249     distributions that replace each point charge -- Gaussian overlap
250     integrals yield complementary error functions when truncated at a
251     finite distance.
252 mlamichh 4114
253 gezelter 4168 By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
254     cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
255     agreement with SPME for the interaction energies, forces and torques
256     for charge-charge interactions.\cite{Fennell:2006lq}
257 gezelter 4167
258 gezelter 4168 \subsection{Point multipoles in molecular modeling}
259     Coarse-graining approaches which treat entire molecular subsystems as
260     a single rigid body are now widely used. A common feature of many
261     coarse-graining approaches is simplification of the electrostatic
262     interactions between bodies so that fewer site-site interactions are
263     required to compute configurational energies. Many coarse-grained
264     molecular structures would normally consist of equal positive and
265     negative charges, and rather than use multiple site-site interactions,
266     the interaction between higher order multipoles can also be used to
267     evaluate a single molecule-molecule
268     interaction.\cite{Ren06,Essex10,Essex11}
269 mlamichh 4166
270 gezelter 4168 Because electrons in a molecule are not localized at specific points,
271     the assignment of partial charges to atomic centers is a relatively
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 gezelter 4180 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 gezelter 4168 multipoles up to quadrupolar order have also been coupled with point
279     polarizabilities in the high-quality AMOEBA and iAMOEBA water
280 gezelter 4180 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
281 gezelter 4168 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.
284 mlamichh 4166
285 gezelter 4168 In this paper we test a set of real-space methods that were developed
286     for point multipolar interactions. These methods extend the damped
287     shifted force (DSF) and Wolf methods originally developed for
288     charge-charge interactions and generalize them for higher order
289     multipoles. The detailed mathematical development of these methods has
290     been presented in the first paper in this series, while this work
291     covers the testing the energies, forces, torques, and energy
292     conservation properties of the methods in realistic simulation
293     environments. In all cases, the methods are compared with the
294     reference method, a full multipolar Ewald treatment.
295    
296    
297 mlamichh 4166 %\subsection{Conservation of total energy }
298 gezelter 4167 %To conserve the total energy in MD simulations, the energy, force, and torque on a central molecule due to another molecule should smoothly tends to zero as second molecule approaches to cutoff radius. In addition, the force should be derivable from the energy and vice versa. If only the first condition holds but not the second, the total energy does not conserve.\cite{Fennell:2006lq}. The hard cutoff method does not ensure the smooth transition of the energy, force, and torque at the cutoff radius.\cite{Wolf:1999dn} By placing image charge on the surface of the cutoff sphere, the smooth transition of the energy can be ensured but the force and torque remains discontinuous. Therefore the purposed methods should have smooth transition of the energy, force and torque to ensure the total energy conservation and the expression should be close to idea of placing image multipole on the surface of the cutoff sphere.
299 mlamichh 4166
300 gezelter 4168 \section{\label{sec:method}Review of Methods}
301     Any real-space electrostatic method that is suitable for MD
302     simulations should have the electrostatic energy, forces and torques
303     between two sites go smoothly to zero as the distance between the
304     sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
305     this continuity at the cutoff is essential for energy conservation in
306     MD simulations. The mathematical details of the shifted potential
307     (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
308     methods have been discussed in detail in the previous paper in this
309     series.\cite{PaperI} Here we briefly review the new methods and
310     describe their essential features.
311 mlamichh 4166
312 gezelter 4168 \subsection{Taylor-shifted force (TSF)}
313 mlamichh 4114
314 gezelter 4168 The electrostatic potential energy between point multipoles can be
315     expressed as the product of two multipole operators and a Coulombic
316     kernel,
317 mlamichh 4114 \begin{equation}
318 gezelter 4168 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
319 mlamichh 4114 \end{equation}
320 gezelter 4180 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 mlamichh 4166
325 gezelter 4180 % 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 gezelter 4168
344 gezelter 4180 The TSF potential for any multipole-multipole interaction can be
345     written
346 gezelter 4168 \begin{equation}
347     U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
348     \label{generic}
349     \end{equation}
350 gezelter 4180 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 gezelter 4168
358 gezelter 4180 % 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 gezelter 4168
364 gezelter 4180 % 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 gezelter 4168
373     For multipole-multipole interactions, following this procedure results
374 gezelter 4180 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 gezelter 4168 dot products:
381     \begin{equation}
382 gezelter 4180 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 gezelter 4168 \end{equation}
388    
389 gezelter 4180 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 gezelter 4168 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 gezelter 4175 \onlinecite{PaperI}).
401 gezelter 4168
402     \subsection{Gradient-shifted force (GSF)}
403    
404 gezelter 4180 A second (and conceptually simpler) method involves shifting the
405     gradient of the raw Coulomb potential for each particular multipole
406 gezelter 4168 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 gezelter 4180 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 gezelter 4168 \end{displaymath}
416 gezelter 4180 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 gezelter 4168
428 gezelter 4180 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 gezelter 4168 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}
438     U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
439     U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
440     \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert _{r_c} \right]
441     \label{generic2}
442     \end{equation}
443     where the sum describes a separate force-shifting that is applied to
444     each orientational contribution to the energy.
445    
446     The third term converges more rapidly than the first two terms as a
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 gezelter 4175 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
451 gezelter 4168 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 gezelter 4175 (Reference~\onlinecite{PaperI})
458 gezelter 4168
459    
460 mlamichh 4166 \subsection{Shifted potential (SP) }
461 gezelter 4168 A discontinuous truncation of the electrostatic potential at the
462     cutoff sphere introduces a severe artifact (oscillation in the
463     electrostatic energy) even for molecules with the higher-order
464     multipoles.\cite{PaperI} We have also formulated an extension of the
465     Wolf approach for point multipoles by simply projecting the image
466     multipole onto the surface of the cutoff sphere, and including the
467     interactions with the central multipole and the image. This
468     effectively shifts the total potential to zero at the cutoff radius,
469 mlamichh 4166 \begin{equation}
470 gezelter 4180 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 mlamichh 4166 \label{eq:SP}
473     \end{equation}
474 gezelter 4168 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 gezelter 4175 is not a simple shifting of the total potential at $r_c$. Each radial
479 gezelter 4168 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 mlamichh 4166
483 gezelter 4180 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 gezelter 4168
494 gezelter 4180 \subsection{The Self Term}
495 gezelter 4168 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 gezelter 4175 this series (Reference \onlinecite{PaperI}).
501 mlamichh 4162
502 gezelter 4168
503 gezelter 4170 \section{\label{sec:methodology}Methodology}
504 mlamichh 4166
505 gezelter 4170 To understand how the real-space multipole methods behave in computer
506     simulations, it is vital to test against established methods for
507     computing electrostatic interactions in periodic systems, and to
508     evaluate the size and sources of any errors that arise from the
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 gezelter 4174 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 gezelter 4175 disordered and ordered condensed-phase systems. The parameters used
516 gezelter 4180 in the test cases are given in table~\ref{tab:pars}.
517 gezelter 4174
518 gezelter 4175 \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 gezelter 4180 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
535 gezelter 4175 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
536 gezelter 4180 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
537 gezelter 4175 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 mlamichh 4166
557 gezelter 4175 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 gezelter 4180 and have been compared with the values obtained from the multipolar
562     Ewald sum. In Monte Carlo (MC) simulations, the energy differences
563 gezelter 4175 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
566     energies are not exact. For each of the multipolar systems listed
567     above, we have compared the change in electrostatic potential energy
568     ($\Delta E$) between 250 statistically-independent configurations. In
569     molecular dynamics (MD) simulations, the forces and torques govern the
570     behavior of the simulation, so we also compute the electrostatic
571     contributions to the forces and torques.
572    
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     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     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 gezelter 4170 \subsection{Model systems}
609 gezelter 4180 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 gezelter 4170
626 gezelter 4180 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 gezelter 4170
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
643     different configurations. We took the Ewald-computed energy
644     difference between two conformations to be the correct behavior. An
645     ideal performance by one of the new methods would reproduce these
646     energy differences exactly. The configurational energies being used
647     here contain only contributions from electrostatic interactions.
648     Lennard-Jones interactions were omitted from the comparison as they
649     should be identical for all methods.
650    
651     Since none of the real-space methods provide exact energy differences,
652 gezelter 4180 we used least square regressions analysis for the six different
653 gezelter 4170 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
656     regressions indicate perfect agreement between the real-space method
657     and the multipolar Ewald sum.
658    
659     Molecular systems were run long enough to explore independent
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 gezelter 4180 also been compared using least squares regression analysis. In the
664 gezelter 4170 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
667     different configurations for each system resulting in 3,072,000 data
668     points for comparison of forces and torques.
669    
670 mlamichh 4166 \subsection{Analysis of vector quantities}
671 gezelter 4170 Getting the magnitudes of the force and torque vectors correct is only
672     part of the issue for carrying out accurate molecular dynamics
673     simulations. Because the real space methods reweight the different
674     orientational contributions to the energies, it is also important to
675     understand how the methods impact the \textit{directionality} of the
676     force and torque vectors. Fisher developed a probablity density
677     function to analyse directional data sets,
678 mlamichh 4162 \begin{equation}
679 gezelter 4170 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
680 mlamichh 4162 \label{eq:pdf}
681     \end{equation}
682 gezelter 4170 where $\kappa$ measures directional dispersion of the data around the
683     mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
684     estimated as a reciprocal of the circular variance.\cite{Allen91} To
685     quantify the directional error, forces obtained from the Ewald sum
686     were taken as the mean (or correct) direction and the angle between
687     the forces obtained via the Ewald sum and the real-space methods were
688     evaluated,
689 mlamichh 4162 \begin{equation}
690 gezelter 4170 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
691     \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
692     \end{equation}
693     The total angular displacement of the vectors was calculated as,
694     \begin{equation}
695     R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
696 mlamichh 4162 \label{eq:displacement}
697     \end{equation}
698 gezelter 4170 where $N$ is number of force vectors. The circular variance is
699     defined as
700     \begin{equation}
701     \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
702     \end{equation}
703     The circular variance takes on values between from 0 to 1, with 0
704     indicating a perfect directional match between the Ewald force vectors
705     and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
706     correspond to higher values of $\kappa$, which indicates tighter
707     clustering of the real-space force vectors around the Ewald forces.
708 mlamichh 4162
709 gezelter 4170 A similar analysis was carried out for the electrostatic contribution
710     to the molecular torques as well as forces.
711    
712 mlamichh 4166 \subsection{Energy conservation}
713 gezelter 4170 To test conservation the energy for the methods, the mixed molecular
714     system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
715     ions was run for 1 ns in the microcanonical ensemble at an average
716     temperature of 300K. Each of the different electrostatic methods
717     (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
718     damping values. The molecular system was started with same initial
719     positions and velocities for all cutoff methods. The energy drift
720     ($\delta E_1$) and standard deviation of the energy about the slope
721     ($\delta E_0$) were evaluated from the total energy of the system as a
722     function of time. Although both measures are valuable at
723     investigating new methods for molecular dynamics, a useful interaction
724     model must allow for long simulation times with minimal energy drift.
725 mlamichh 4114
726 mlamichh 4166 \section{\label{sec:result}RESULTS}
727     \subsection{Configurational energy differences}
728     %The magnitude of the fluctuation in the total electrostatic energy per molecule for a dipolar crystal is very high as shown in (Fig … paper I).\cite{PaperI}As soon as, the net dipole moment within a cutoff radius is neutralized in the SP method, the magnitude of the fluctuation in the total electrostatic energy per molecule reduced significantly and rapidly converged to the correct energy constant (Refer figure … Paper I).\cite{PaperI} The GSF potential energy also converged to the correct energy constant for the cutoff radius rc = 6a for the undamped case. The potential energy from the TSF method converges towards the correct value for a very large cutoff radius. The speed of convergence for the all the cutoff methods can be increased by using damping function as shown in figure … Paper I\cite{PaperI}. For the quadrupolar crystals, the fluctuation in the total electrostatic energy for the hard cutoff method is small and short ranged as compared to the dipolar crystals (figure … in the paper I).\cite{PaperI} Similar to the dipolar crystals, the net quadrupolar neutralization of the cutoff sphere in the SP method reduces oscillation rapidly and converge electrostatic energy to the correct energy constant.
729     %The oscillation in the the electrostatic energy for the hard cutoff method is even true for the dipolar liquids as shown in Figure ~\ref{fig:rcutConvergence_dipolarLiquid}. As we placed image on the surface of the cutoff sphere in SP method, the oscillation in the energy is reduced. The fluctuation in the energy in liquid is much smaller as compared to the crystal (This result is similar to the results observed by \textit{Wolf et al.} in the case of ionic crystal and Mgo melt). The large magnitude in the fluctuation of the electrostatic energy in the crystal is because of large range of multipole ordering in the crystal. When the energy is evaluated by the direct truncation, it breaks up large number of multipolar ordering leaving behind net multipole moment. But in the case of liquid, there is only local ordering of the multipoles and their ordering disappears in the long range. Therefore, the direct truncation results a small oscillation in the electrostatic energy (which is smaller than deviation SP energy from the Ewald Figure ~\ref{fig:rcutConvergence_dipolarLiquid}) in the case of liquid. Although, the oscillation in the energy is very small for the case of liquid, this affects the change in potential energy ($\triangle E$), which is observed when $\triangle E$ evaluated from the SP method compared with Ewald as shown in figure 4a and 4b.
730     %\begin{figure}[h!]
731     % \centering
732     % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
733     % \caption{The energy per molecule plotted against cutoff radius, rc for i) Hard ii) SP iii) GSF, and iv) TSF method. The hard cutoff method shows fluctuation in the electorstatic energy and it disappers in all other methods. }
734     % \label{fig:rcutConvergence_dipolarLiquid}
735     % \end{figure}
736     %In MC simulations, the electrostatic differences between the molecules are important parameter for sampling. We have compared $\triangle E$ from the different methods (Hard, SP, GSF, and TSF) with the Ewald using linear regression analysis. The correlation coefficient ($R^2$) of the regression line measures the deviation of the evaluated quantities from the mean slope. We know that Ewald method evaluates accurate value of the electrostatic energy. Hence, if the proposed methods can quantify the electrostatic energy as good as Ewald then the correlation coefficient is 1.The correlation coefficient is 0 for the completely random result for any physical quantity measured by the proposed method. The slope is a measure of the accuracy of the average of a physical quantity obtained from the proposed methods. If the slope is 1 then we can conclude that the average of the physical quantity measured by the method is as good as Ewald. The deviation of the slope from 1 state that the method used in quantifying physical quantities is statistically biased as compared to Ewald.
737     %\begin{figure}
738     % \centering
739     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
740     % \label{fig:barGraph1}
741     % \end{figure}
742     % \begin{figure}
743     % \centering
744     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
745     % \caption{}
746 mlamichh 4162
747 gezelter 4167 % \label{fig:barGraph2}
748     % \end{figure}
749 gezelter 4174 %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 mlamichh 4114 \begin{figure}
782 gezelter 4174 \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 gezelter 4175 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 gezelter 4174 orientations of the multipoles (particularly in the ordered crystals)
805 gezelter 4175 mean that these energy jumps can go in either direction, producing a
806     significant amount of random scatter, but no systematic error.
807 gezelter 4174
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 gezelter 4175 effect, particularly for the crystalline systems.
815 gezelter 4174
816     Both the SP and GSF methods appear to reproduce the Ewald results with
817     excellent fidelity, particularly for moderate damping ($\alpha =
818 gezelter 4175 0.1-0.2$\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 gezelter 4174
823 mlamichh 4114 \subsection{Magnitude of the force and torque vectors}
824 gezelter 4174
825 gezelter 4175 The comparisons of the magnitudes of the forces and torques for the
826     data accumulated from all six systems are shown in Figures
827 gezelter 4174 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
828     correlation and slope for the forces agree well with the Ewald sum
829 gezelter 4175 even for the hard cutoffs.
830 gezelter 4174
831 gezelter 4175 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 gezelter 4174
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 gezelter 4175 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 gezelter 4180 perturbations are minimal, particularly for moderate damping and
849 gezelter 4174 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 mlamichh 4114 \begin{figure}
855 gezelter 4174 \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 mlamichh 4114 \begin{figure}
869 gezelter 4174 \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 gezelter 4175 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 gezelter 4174
898 mlamichh 4114 \subsection{Directionality of the force and torque vectors}
899 mlamichh 4162
900 gezelter 4174 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 gezelter 4180 ($\mathrm{Var}(\theta)$) in figure
906 gezelter 4174 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
907 gezelter 4175 from the new real-space methods exhibit nearly-ideal Fisher probability
908 gezelter 4174 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 gezelter 4175 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
913 gezelter 4174 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
914 gezelter 4175 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 gezelter 4174 agreement with the Ewald force directions.
917    
918 gezelter 4175 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 gezelter 4174 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 mlamichh 4114 \begin{figure}
927 gezelter 4174 \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 gezelter 4171
939 gezelter 4175 \subsection{Energy conservation\label{sec:conservation}}
940 gezelter 4171
941 gezelter 4174 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 gezelter 4175 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
948 gezelter 4174 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 gezelter 4175 a cutoff radius of 12\AA. The value of the damping coefficient was
951 gezelter 4174 also varied from the undamped case ($\alpha = 0$) to a heavily damped
952 gezelter 4175 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 gezelter 4174
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 gezelter 4175 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 gezelter 4174
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 mlamichh 4114 \begin{figure}
974 gezelter 4171 \centering
975 gezelter 4180 \includegraphics[width=\textwidth]{newDrift_12.pdf}
976 mlamichh 4162 \label{fig:energyDrift}
977 gezelter 4174 \caption{Analysis of the energy conservation of the real-space
978 gezelter 4171 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
979 gezelter 4180 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 gezelter 4171 \end{figure}
986    
987 gezelter 4174
988 mlamichh 4114 \section{CONCLUSION}
989 gezelter 4175 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 gezelter 4170
999 gezelter 4175 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 gezelter 4170
1011 gezelter 4175 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 gezelter 4180 \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 gezelter 4167 %\bibliographystyle{aip}
1073 gezelter 4168 \newpage
1074 mlamichh 4114 \bibliography{references}
1075     \end{document}
1076    
1077     %
1078     % ****** End of file aipsamp.tex ******