ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4184
Committed: Sat Jun 14 23:35:39 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 62725 byte(s)
Log Message:
Changes

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 gezelter 4181 \usepackage{mathptmx}
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 gezelter 4184 against the multipolar Ewald sum as a reference method. The tests
74     were carried out in a variety of condensed-phase environments which
75     were designed to test all levels of the multipole-multipole
76     interactions. Comparisons of the energy differences between
77     configurations, molecular forces, and torques were used to analyze
78     how well the real-space models perform relative to the more
79     computationally expensive Ewald treatment. We have also investigated the
80     energy conservation properties of the new methods in molecular
81     dynamics simulations using all of these methods. The SP method shows
82     excellent agreement with configurational energy differences, forces,
83     and torques, and would be suitable for use in Monte Carlo
84     calculations. Of the two new shifted-force methods, the GSF
85     approach shows the best agreement with Ewald-derived energies,
86     forces, and torques and exhibits energy conservation properties that
87     make it an excellent choice for efficiently computing electrostatic
88     interactions in molecular dynamics simulations.
89 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 4184 %\keywords{Electrostatics, Multipoles, Real-space}
94 gezelter 4167
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 gezelter 4184 periodicity in the Ewald method can also be problematic for molecular
126     interfaces.\cite{Fennell:2006lq} Modified Ewald methods that were
127     developed to handle two-dimensional (2D) electrostatic interactions in
128     interfacial systems have not seen similar particle-mesh
129     treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
130     Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
131     with system size.
132 gezelter 4167
133 mlamichh 4166 \subsection{Real-space methods}
134 gezelter 4168 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
135     method for calculating electrostatic interactions between point
136 gezelter 4167 charges. They argued that the effective Coulomb interaction in
137 gezelter 4184 condensed phase systems is actually short ranged.\cite{Wolf92,Wolf95}
138     For an ordered lattice (e.g., when computing the Madelung constant of
139     an ionic solid), the material can be considered as a set of ions
140 gezelter 4167 interacting with neutral dipolar or quadrupolar ``molecules'' giving
141     an effective distance dependence for the electrostatic interactions of
142 gezelter 4181 $r^{-5}$ (see figure \ref{fig:schematic}). For this reason, careful
143 gezelter 4184 application of Wolf's method can obtain accurate estimates of Madelung
144     constants using relatively short cutoff radii. Recently, Fukuda used
145     neutralization of the higher order moments for calculation of the
146     electrostatic interactions in point charge
147     systems.\cite{Fukuda:2013sf}
148 mlamichh 4114
149 gezelter 4181 \begin{figure}
150 gezelter 4167 \centering
151 gezelter 4181 \includegraphics[width=\linewidth]{schematic.pdf}
152     \caption{Top: Ionic systems exhibit local clustering of dissimilar
153     charges (in the smaller grey circle), so interactions are
154 gezelter 4184 effectively charge-multipole at longer distances. With hard
155     cutoffs, motion of individual charges in and out of the cutoff
156     sphere can break the effective multipolar ordering. Bottom:
157     dipolar crystals and fluids have a similar effective
158 gezelter 4181 \textit{quadrupolar} ordering (in the smaller grey circles), and
159     orientational averaging helps to reduce the effective range of the
160     interactions in the fluid. Placement of reversed image multipoles
161     on the surface of the cutoff sphere recovers the effective
162     higher-order multipole behavior.}
163     \label{fig:schematic}
164 gezelter 4167 \end{figure}
165 mlamichh 4114
166 gezelter 4167 The direct truncation of interactions at a cutoff radius creates
167 gezelter 4184 truncation defects. Wolf \textit{et al.} argued that
168 gezelter 4167 truncation errors are due to net charge remaining inside the cutoff
169     sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
170     placing an image charge on the surface of the cutoff sphere for every
171     real charge inside the cutoff. These charges are present for the
172     evaluation of both the pair interaction energy and the force, although
173     the force expression maintained a discontinuity at the cutoff sphere.
174     In the original Wolf formulation, the total energy for the charge and
175     image were not equal to the integral of their force expression, and as
176     a result, the total energy would not be conserved in molecular
177     dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
178     Fennel and Gezelter later proposed shifted force variants of the Wolf
179     method with commensurate force and energy expressions that do not
180 gezelter 4168 exhibit this problem.\cite{Fennell:2006lq} Related real-space
181     methods were also proposed by Chen \textit{et
182     al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
183     and by Wu and Brooks.\cite{Wu:044107}
184 gezelter 4167
185     Considering the interaction of one central ion in an ionic crystal
186     with a portion of the crystal at some distance, the effective Columbic
187 gezelter 4181 potential is found to decrease as $r^{-5}$. If one views the \ce{NaCl}
188     crystal as a simple cubic (SC) structure with an octupolar
189 gezelter 4167 \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
190     rapidly to the Madelung energy than the dipolar
191     approximation.\cite{Wolf92} To find the correct Madelung constant,
192     Lacman suggested that the NaCl structure could be constructed in a way
193     that the finite crystal terminates with complete \ce{(NaCl)4}
194 gezelter 4181 molecules.\cite{Lacman65} The central ion sees what is effectively a
195     set of octupoles at large distances. These facts suggest that the
196     Madelung constants are relatively short ranged for perfect ionic
197     crystals.\cite{Wolf:1999dn}
198 gezelter 4167
199     One can make a similar argument for crystals of point multipoles. The
200     Luttinger and Tisza treatment of energy constants for dipolar lattices
201     utilizes 24 basis vectors that contain dipoles at the eight corners of
202     a unit cube. Only three of these basis vectors, $X_1, Y_1,
203     \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
204     zero net dipole and retain contributions only from higher order
205     multipoles. The effective interaction between a dipole at the center
206     of a crystal and a group of eight dipoles farther away is
207     significantly shorter ranged than the $r^{-3}$ that one would expect
208     for raw dipole-dipole interactions. Only in crystals which retain a
209     bulk dipole moment (e.g. ferroelectrics) does the analogy with the
210     ionic crystal break down -- ferroelectric dipolar crystals can exist,
211     while ionic crystals with net charge in each unit cell would be
212     unstable.
213    
214     In ionic crystals, real-space truncation can break the effective
215 gezelter 4181 multipolar arrangements (see Fig. \ref{fig:schematic}), causing
216     significant swings in the electrostatic energy as individual ions move
217     back and forth across the boundary. This is why the image charges are
218 gezelter 4180 necessary for the Wolf sum to exhibit rapid convergence. Similarly,
219     the real-space truncation of point multipole interactions breaks
220     higher order multipole arrangements, and image multipoles are required
221     for real-space treatments of electrostatic energies.
222 gezelter 4167
223 gezelter 4181 The shorter effective range of electrostatic interactions is not
224     limited to perfect crystals, but can also apply in disordered fluids.
225     Even at elevated temperatures, there is, on average, local charge
226     balance in an ionic liquid, where each positive ion has surroundings
227     dominated by negaitve ions and vice versa. The reversed-charge images
228     on the cutoff sphere that are integral to the Wolf and DSF approaches
229     retain the effective multipolar interactions as the charges traverse
230     the cutoff boundary.
231    
232     In multipolar fluids (see Fig. \ref{fig:schematic}) there is
233     significant orientational averaging that additionally reduces the
234     effect of long-range multipolar interactions. The image multipoles
235     that are introduced in the TSF, GSF, and SP methods mimic this effect
236     and reduce the effective range of the multipolar interactions as
237     interacting molecules traverse each other's cutoff boundaries.
238    
239 gezelter 4167 % Because of this reason, although the nature of electrostatic
240     % interaction short ranged, the hard cutoff sphere creates very large
241     % fluctuation in the electrostatic energy for the perfect crystal. In
242     % addition, the charge neutralized potential proposed by Wolf et
243     % al. converged to correct Madelung constant but still holds oscillation
244     % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
245     % oscillation in the energy around its fully converged value can be due
246     % to the non-neutralized value of the higher order moments within the
247     % cutoff sphere.
248    
249     The forces and torques acting on atomic sites are the fundamental
250     factors driving dynamics in molecular simulations. Fennell and
251     Gezelter proposed the damped shifted force (DSF) energy kernel to
252     obtain consistent energies and forces on the atoms within the cutoff
253 gezelter 4168 sphere. Both the energy and the force go smoothly to zero as an atom
254     aproaches the cutoff radius. The comparisons of the accuracy these
255     quantities between the DSF kernel and SPME was surprisingly
256     good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
257     calculating electrostatic interactions in molecular systems with
258     relatively uniform charge
259     densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
260 gezelter 4167
261 gezelter 4168 \subsection{The damping function}
262 gezelter 4167 The damping function used in our research has been discussed in detail
263     in the first paper of this series.\cite{PaperI} The radial kernel
264 gezelter 4168 $1/r$ for the interactions between point charges can be replaced by
265     the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
266     accelerate the rate of convergence, where $\alpha$ is a damping
267     parameter with units of inverse distance. Altering the value of
268     $\alpha$ is equivalent to changing the width of Gaussian charge
269     distributions that replace each point charge -- Gaussian overlap
270     integrals yield complementary error functions when truncated at a
271     finite distance.
272 mlamichh 4114
273 gezelter 4168 By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
274     cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
275     agreement with SPME for the interaction energies, forces and torques
276     for charge-charge interactions.\cite{Fennell:2006lq}
277 gezelter 4167
278 gezelter 4168 \subsection{Point multipoles in molecular modeling}
279     Coarse-graining approaches which treat entire molecular subsystems as
280     a single rigid body are now widely used. A common feature of many
281     coarse-graining approaches is simplification of the electrostatic
282     interactions between bodies so that fewer site-site interactions are
283     required to compute configurational energies. Many coarse-grained
284     molecular structures would normally consist of equal positive and
285     negative charges, and rather than use multiple site-site interactions,
286     the interaction between higher order multipoles can also be used to
287     evaluate a single molecule-molecule
288     interaction.\cite{Ren06,Essex10,Essex11}
289 mlamichh 4166
290 gezelter 4168 Because electrons in a molecule are not localized at specific points,
291     the assignment of partial charges to atomic centers is a relatively
292     rough approximation. Atomic sites can also be assigned point
293     multipoles and polarizabilities to increase the accuracy of the
294     molecular model. Recently, water has been modeled with point
295 gezelter 4180 multipoles up to octupolar order using the soft sticky
296     dipole-quadrupole-octupole (SSDQO)
297     model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
298 gezelter 4168 multipoles up to quadrupolar order have also been coupled with point
299     polarizabilities in the high-quality AMOEBA and iAMOEBA water
300 gezelter 4180 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
301 gezelter 4168 using point multipole with the real space truncation without
302     accounting for multipolar neutrality will create energy conservation
303     issues in molecular dynamics (MD) simulations.
304 mlamichh 4166
305 gezelter 4168 In this paper we test a set of real-space methods that were developed
306     for point multipolar interactions. These methods extend the damped
307     shifted force (DSF) and Wolf methods originally developed for
308     charge-charge interactions and generalize them for higher order
309     multipoles. The detailed mathematical development of these methods has
310     been presented in the first paper in this series, while this work
311     covers the testing the energies, forces, torques, and energy
312     conservation properties of the methods in realistic simulation
313     environments. In all cases, the methods are compared with the
314     reference method, a full multipolar Ewald treatment.
315    
316    
317 mlamichh 4166 %\subsection{Conservation of total energy }
318 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.
319 mlamichh 4166
320 gezelter 4168 \section{\label{sec:method}Review of Methods}
321     Any real-space electrostatic method that is suitable for MD
322     simulations should have the electrostatic energy, forces and torques
323     between two sites go smoothly to zero as the distance between the
324     sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
325     this continuity at the cutoff is essential for energy conservation in
326     MD simulations. The mathematical details of the shifted potential
327     (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
328     methods have been discussed in detail in the previous paper in this
329     series.\cite{PaperI} Here we briefly review the new methods and
330     describe their essential features.
331 mlamichh 4166
332 gezelter 4168 \subsection{Taylor-shifted force (TSF)}
333 mlamichh 4114
334 gezelter 4168 The electrostatic potential energy between point multipoles can be
335     expressed as the product of two multipole operators and a Coulombic
336     kernel,
337 mlamichh 4114 \begin{equation}
338 gezelter 4168 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
339 mlamichh 4114 \end{equation}
340 gezelter 4180 where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
341     expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
342     a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
343 gezelter 4184 $\bf a$, etc.
344 mlamichh 4166
345 gezelter 4180 % Interactions between multipoles can be expressed as higher derivatives
346     % of the bare Coulomb potential, so one way of ensuring that the forces
347     % and torques vanish at the cutoff distance is to include a larger
348     % number of terms in the truncated Taylor expansion, e.g.,
349     % %
350     % \begin{equation}
351     % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
352     % \end{equation}
353     % %
354     % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
355     % Thus, for $f(r)=1/r$, we find
356     % %
357     % \begin{equation}
358     % f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
359     % \end{equation}
360     % This function is an approximate electrostatic potential that has
361     % vanishing second derivatives at the cutoff radius, making it suitable
362     % for shifting the forces and torques of charge-dipole interactions.
363 gezelter 4168
364 gezelter 4180 The TSF potential for any multipole-multipole interaction can be
365     written
366 gezelter 4168 \begin{equation}
367     U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
368     \label{generic}
369     \end{equation}
370 gezelter 4180 where $f_n(r)$ is a shifted kernel that is appropriate for the order
371 gezelter 4181 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
372     charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
373     and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
374     quadrupole-quadrupole. To ensure smooth convergence of the energy,
375     force, and torques, a Taylor expansion with $n$ terms must be
376     performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
377 gezelter 4168
378 gezelter 4180 % To carry out the same procedure for a damped electrostatic kernel, we
379     % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
380     % Many of the derivatives of the damped kernel are well known from
381     % Smith's early work on multipoles for the Ewald
382     % summation.\cite{Smith82,Smith98}
383 gezelter 4168
384 gezelter 4180 % Note that increasing the value of $n$ will add additional terms to the
385     % electrostatic potential, e.g., $f_2(r)$ includes orders up to
386     % $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
387     % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
388     % f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
389     % for computing multipole energies, forces, and torques, and smooth
390     % cutoffs of these quantities can be guaranteed as long as the number of
391     % terms in the Taylor series exceeds the derivative order required.
392 gezelter 4168
393     For multipole-multipole interactions, following this procedure results
394 gezelter 4180 in separate radial functions for each of the distinct orientational
395     contributions to the potential, and ensures that the forces and
396     torques from each of these contributions will vanish at the cutoff
397     radius. For example, the direct dipole dot product
398     ($\mathbf{D}_{\bf a}
399     \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
400 gezelter 4168 dot products:
401     \begin{equation}
402 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
403     \mathbf{D}_{\bf a} \cdot
404     \mathbf{D}_{\bf b} \right) v_{21}(r) +
405     \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
406     \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
407 gezelter 4168 \end{equation}
408    
409 gezelter 4180 For the Taylor shifted (TSF) method with the undamped kernel,
410     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
411     \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
412     - \frac{6}{r r_c^2}$. In these functions, one can easily see the
413     connection to unmodified electrostatics as well as the smooth
414     transition to zero in both these functions as $r\rightarrow r_c$. The
415     electrostatic forces and torques acting on the central multipole due
416 gezelter 4184 to another site within the cutoff sphere are derived from
417 gezelter 4168 Eq.~\ref{generic}, accounting for the appropriate number of
418     derivatives. Complete energy, force, and torque expressions are
419     presented in the first paper in this series (Reference
420 gezelter 4175 \onlinecite{PaperI}).
421 gezelter 4168
422     \subsection{Gradient-shifted force (GSF)}
423    
424 gezelter 4180 A second (and conceptually simpler) method involves shifting the
425     gradient of the raw Coulomb potential for each particular multipole
426 gezelter 4168 order. For example, the raw dipole-dipole potential energy may be
427     shifted smoothly by finding the gradient for two interacting dipoles
428     which have been projected onto the surface of the cutoff sphere
429     without changing their relative orientation,
430 gezelter 4181 \begin{equation}
431 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r) = U_{D_{\bf a}D_{\bf b}}(r) -
432     U_{D_{\bf a} D_{\bf b}}(r_c)
433     - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
434 gezelter 4184 \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
435 gezelter 4181 \end{equation}
436 gezelter 4180 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
437     a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
438     (although the signs are reversed for the dipole that has been
439     projected onto the cutoff sphere). In many ways, this simpler
440     approach is closer in spirit to the original shifted force method, in
441     that it projects a neutralizing multipole (and the resulting forces
442     from this multipole) onto a cutoff sphere. The resulting functional
443     forms for the potentials, forces, and torques turn out to be quite
444     similar in form to the Taylor-shifted approach, although the radial
445     contributions are significantly less perturbed by the gradient-shifted
446     approach than they are in the Taylor-shifted method.
447 gezelter 4168
448 gezelter 4180 For the gradient shifted (GSF) method with the undamped kernel,
449     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
450     $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
451     Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
452     because the Taylor expansion retains only one term, they are
453     significantly less perturbed than the TSF functions.
454    
455 gezelter 4168 In general, the gradient shifted potential between a central multipole
456     and any multipolar site inside the cutoff radius is given by,
457     \begin{equation}
458 gezelter 4184 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
459     U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
460     \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
461 gezelter 4168 \label{generic2}
462     \end{equation}
463     where the sum describes a separate force-shifting that is applied to
464 gezelter 4184 each orientational contribution to the energy. In this expression,
465     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
466     ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
467     represent the orientations the multipoles.
468 gezelter 4168
469     The third term converges more rapidly than the first two terms as a
470     function of radius, hence the contribution of the third term is very
471     small for large cutoff radii. The force and torque derived from
472 gezelter 4184 Eq. \ref{generic2} are consistent with the energy expression and
473 gezelter 4175 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
474 gezelter 4168 can be considered generalizations of the original DSF method for
475     higher order multipole interactions. GSF and TSF are also identical up
476     to the charge-dipole interaction but generate different expressions in
477     the energy, force and torque for higher order multipole-multipole
478     interactions. Complete energy, force, and torque expressions for the
479     GSF potential are presented in the first paper in this series
480 gezelter 4184 (Reference~\onlinecite{PaperI}).
481 gezelter 4168
482    
483 mlamichh 4166 \subsection{Shifted potential (SP) }
484 gezelter 4168 A discontinuous truncation of the electrostatic potential at the
485     cutoff sphere introduces a severe artifact (oscillation in the
486     electrostatic energy) even for molecules with the higher-order
487     multipoles.\cite{PaperI} We have also formulated an extension of the
488     Wolf approach for point multipoles by simply projecting the image
489     multipole onto the surface of the cutoff sphere, and including the
490     interactions with the central multipole and the image. This
491     effectively shifts the total potential to zero at the cutoff radius,
492 mlamichh 4166 \begin{equation}
493 gezelter 4180 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
494 gezelter 4184 U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
495 mlamichh 4166 \label{eq:SP}
496     \end{equation}
497 gezelter 4168 where the sum describes separate potential shifting that is done for
498     each orientational contribution to the energy (e.g. the direct dipole
499     product contribution is shifted {\it separately} from the
500     dipole-distance terms in dipole-dipole interactions). Note that this
501 gezelter 4175 is not a simple shifting of the total potential at $r_c$. Each radial
502 gezelter 4168 contribution is shifted separately. One consequence of this is that
503     multipoles that reorient after leaving the cutoff sphere can re-enter
504     the cutoff sphere without perturbing the total energy.
505 mlamichh 4166
506 gezelter 4180 For the shifted potential (SP) method with the undamped kernel,
507     $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
508     \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
509     central multipole and other multipolar sites goes smoothly to zero as
510     $r \rightarrow r_c$. However, the force and torque obtained from the
511     shifted potential (SP) are discontinuous at $r_c$. MD simulations
512     will still experience energy drift while operating under the SP
513     potential, but it may be suitable for Monte Carlo approaches where the
514     configurational energy differences are the primary quantity of
515     interest.
516 gezelter 4168
517 gezelter 4180 \subsection{The Self Term}
518 gezelter 4168 In the TSF, GSF, and SP methods, a self-interaction is retained for
519     the central multipole interacting with its own image on the surface of
520     the cutoff sphere. This self interaction is nearly identical with the
521     self-terms that arise in the Ewald sum for multipoles. Complete
522     expressions for the self terms are presented in the first paper in
523 gezelter 4175 this series (Reference \onlinecite{PaperI}).
524 mlamichh 4162
525 gezelter 4168
526 gezelter 4170 \section{\label{sec:methodology}Methodology}
527 mlamichh 4166
528 gezelter 4170 To understand how the real-space multipole methods behave in computer
529     simulations, it is vital to test against established methods for
530     computing electrostatic interactions in periodic systems, and to
531     evaluate the size and sources of any errors that arise from the
532     real-space cutoffs. In the first paper of this series, we compared
533     the dipolar and quadrupolar energy expressions against analytic
534     expressions for ordered dipolar and quadrupolar
535 gezelter 4174 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
536     used the multipolar Ewald sum as a reference method for comparing
537     energies, forces, and torques for molecular models that mimic
538 gezelter 4175 disordered and ordered condensed-phase systems. The parameters used
539 gezelter 4180 in the test cases are given in table~\ref{tab:pars}.
540 gezelter 4174
541 gezelter 4175 \begin{table}
542     \label{tab:pars}
543     \caption{The parameters used in the systems used to evaluate the new
544     real-space methods. The most comprehensive test was a liquid
545     composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
546     ions). This test excercises all orders of the multipolar
547     interactions developed in the first paper.}
548     \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
549     & \multicolumn{2}{c|}{LJ parameters} &
550     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
551     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
552     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
553     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
554     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
555     \AA\textsuperscript{2})} \\ \hline
556     Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
557 gezelter 4180 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
558 gezelter 4175 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
559 gezelter 4180 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
560 gezelter 4175 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
561     \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
562     \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
563     \end{tabularx}
564     \end{table}
565     The systems consist of pure multipolar solids (both dipole and
566     quadrupole), pure multipolar liquids (both dipole and quadrupole), a
567     fluid composed of sites containing both dipoles and quadrupoles
568     simultaneously, and a final test case that includes ions with point
569     charges in addition to the multipolar fluid. The solid-phase
570     parameters were chosen so that the systems can explore some
571     orientational freedom for the multipolar sites, while maintaining
572     relatively strict translational order. The SSDQ model used here is
573     not a particularly accurate water model, but it does test
574     dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
575     interactions at roughly the same magnitudes. The last test case, SSDQ
576     water with dissolved ions, exercises \textit{all} levels of the
577     multipole-multipole interactions we have derived so far and represents
578     the most complete test of the new methods.
579 mlamichh 4166
580 gezelter 4175 In the following section, we present results for the total
581     electrostatic energy, as well as the electrostatic contributions to
582     the force and torque on each molecule. These quantities have been
583     computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
584 gezelter 4180 and have been compared with the values obtained from the multipolar
585     Ewald sum. In Monte Carlo (MC) simulations, the energy differences
586 gezelter 4175 between two configurations is the primary quantity that governs how
587     the simulation proceeds. These differences are the most imporant
588     indicators of the reliability of a method even if the absolute
589     energies are not exact. For each of the multipolar systems listed
590     above, we have compared the change in electrostatic potential energy
591     ($\Delta E$) between 250 statistically-independent configurations. In
592     molecular dynamics (MD) simulations, the forces and torques govern the
593     behavior of the simulation, so we also compute the electrostatic
594     contributions to the forces and torques.
595    
596     \subsection{Implementation}
597     The real-space methods developed in the first paper in this series
598     have been implemented in our group's open source molecular simulation
599     program, OpenMD,\cite{openmd} which was used for all calculations in
600     this work. The complementary error function can be a relatively slow
601     function on some processors, so all of the radial functions are
602     precomputed on a fine grid and are spline-interpolated to provide
603     values when required.
604    
605     Using the same simulation code, we compare to a multipolar Ewald sum
606     with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
607     the Ewald sum is a re-implementation of the algorithm originally
608     proposed by Smith that does not use the particle mesh or smoothing
609     approximations.\cite{Smith82,Smith98} In all cases, the quantities
610     being compared are the electrostatic contributions to energies, force,
611     and torques. All other contributions to these quantities (i.e. from
612     Lennard-Jones interactions) are removed prior to the comparisons.
613    
614     The convergence parameter ($\alpha$) also plays a role in the balance
615     of the real-space and reciprocal-space portions of the Ewald
616     calculation. Typical molecular mechanics packages set this to a value
617     that depends on the cutoff radius and a tolerance (typically less than
618     $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
619     associated with increasing accuracy at the expense of computational
620     time spent on the reciprocal-space portion of the
621     summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
622     10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
623     Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
624    
625     The real-space models have self-interactions that provide
626     contributions to the energies only. Although the self interaction is
627     a rapid calculation, we note that in systems with fluctuating charges
628     or point polarizabilities, the self-term is not static and must be
629     recomputed at each time step.
630    
631 gezelter 4170 \subsection{Model systems}
632 gezelter 4180 To sample independent configurations of the multipolar crystals, body
633     centered cubic (bcc) crystals, which exhibit the minimum energy
634     structures for point dipoles, were generated using 3,456 molecules.
635     The multipoles were translationally locked in their respective crystal
636     sites for equilibration at a relatively low temperature (50K) so that
637     dipoles or quadrupoles could freely explore all accessible
638     orientations. The translational constraints were then removed, the
639     systems were re-equilibrated, and the crystals were simulated for an
640     additional 10 ps in the microcanonical (NVE) ensemble with an average
641     temperature of 50 K. The balance between moments of inertia and
642     particle mass were chosen to allow orientational sampling without
643     significant translational motion. Configurations were sampled at
644     equal time intervals in order to compare configurational energy
645     differences. The crystals were simulated far from the melting point
646     in order to avoid translational deformation away of the ideal lattice
647     geometry.
648 gezelter 4170
649 gezelter 4180 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
650     simulations, each system was created with 2,048 randomly-oriented
651     molecules. These were equilibrated at a temperature of 300K for 1 ns.
652     Each system was then simulated for 1 ns in the microcanonical (NVE)
653     ensemble. We collected 250 different configurations at equal time
654     intervals. For the liquid system that included ionic species, we
655     converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
656     \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
657     run under the same conditions for 1 ns. A total of 250 configurations
658     were collected. In the following comparisons of energies, forces, and
659     torques, the Lennard-Jones potentials were turned off and only the
660     purely electrostatic quantities were compared with the same values
661     obtained via the Ewald sum.
662 gezelter 4170
663     \subsection{Accuracy of Energy Differences, Forces and Torques}
664     The pairwise summation techniques (outlined above) were evaluated for
665     use in MC simulations by studying the energy differences between
666     different configurations. We took the Ewald-computed energy
667     difference between two conformations to be the correct behavior. An
668     ideal performance by one of the new methods would reproduce these
669     energy differences exactly. The configurational energies being used
670     here contain only contributions from electrostatic interactions.
671     Lennard-Jones interactions were omitted from the comparison as they
672     should be identical for all methods.
673    
674     Since none of the real-space methods provide exact energy differences,
675 gezelter 4180 we used least square regressions analysis for the six different
676 gezelter 4170 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
677     with the multipolar Ewald reference method. Unitary results for both
678     the correlation (slope) and correlation coefficient for these
679     regressions indicate perfect agreement between the real-space method
680     and the multipolar Ewald sum.
681    
682     Molecular systems were run long enough to explore independent
683     configurations and 250 configurations were recorded for comparison.
684     Each system provided 31,125 energy differences for a total of 186,750
685     data points. Similarly, the magnitudes of the forces and torques have
686 gezelter 4180 also been compared using least squares regression analysis. In the
687 gezelter 4170 forces and torques comparison, the magnitudes of the forces acting in
688     each molecule for each configuration were evaluated. For example, our
689     dipolar liquid simulation contains 2048 molecules and there are 250
690     different configurations for each system resulting in 3,072,000 data
691     points for comparison of forces and torques.
692    
693 mlamichh 4166 \subsection{Analysis of vector quantities}
694 gezelter 4170 Getting the magnitudes of the force and torque vectors correct is only
695     part of the issue for carrying out accurate molecular dynamics
696     simulations. Because the real space methods reweight the different
697     orientational contributions to the energies, it is also important to
698     understand how the methods impact the \textit{directionality} of the
699     force and torque vectors. Fisher developed a probablity density
700     function to analyse directional data sets,
701 mlamichh 4162 \begin{equation}
702 gezelter 4170 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
703 mlamichh 4162 \label{eq:pdf}
704     \end{equation}
705 gezelter 4170 where $\kappa$ measures directional dispersion of the data around the
706     mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
707     estimated as a reciprocal of the circular variance.\cite{Allen91} To
708     quantify the directional error, forces obtained from the Ewald sum
709     were taken as the mean (or correct) direction and the angle between
710     the forces obtained via the Ewald sum and the real-space methods were
711     evaluated,
712 mlamichh 4162 \begin{equation}
713 gezelter 4170 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
714     \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
715     \end{equation}
716     The total angular displacement of the vectors was calculated as,
717     \begin{equation}
718     R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
719 mlamichh 4162 \label{eq:displacement}
720     \end{equation}
721 gezelter 4170 where $N$ is number of force vectors. The circular variance is
722     defined as
723     \begin{equation}
724     \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
725     \end{equation}
726     The circular variance takes on values between from 0 to 1, with 0
727     indicating a perfect directional match between the Ewald force vectors
728     and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
729     correspond to higher values of $\kappa$, which indicates tighter
730     clustering of the real-space force vectors around the Ewald forces.
731 mlamichh 4162
732 gezelter 4170 A similar analysis was carried out for the electrostatic contribution
733     to the molecular torques as well as forces.
734    
735 mlamichh 4166 \subsection{Energy conservation}
736 gezelter 4170 To test conservation the energy for the methods, the mixed molecular
737     system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
738     ions was run for 1 ns in the microcanonical ensemble at an average
739     temperature of 300K. Each of the different electrostatic methods
740     (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
741     damping values. The molecular system was started with same initial
742     positions and velocities for all cutoff methods. The energy drift
743     ($\delta E_1$) and standard deviation of the energy about the slope
744     ($\delta E_0$) were evaluated from the total energy of the system as a
745     function of time. Although both measures are valuable at
746     investigating new methods for molecular dynamics, a useful interaction
747     model must allow for long simulation times with minimal energy drift.
748 mlamichh 4114
749 mlamichh 4166 \section{\label{sec:result}RESULTS}
750     \subsection{Configurational energy differences}
751     %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.
752     %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.
753     %\begin{figure}[h!]
754     % \centering
755     % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
756     % \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. }
757     % \label{fig:rcutConvergence_dipolarLiquid}
758     % \end{figure}
759     %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.
760     %\begin{figure}
761     % \centering
762     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
763     % \label{fig:barGraph1}
764     % \end{figure}
765     % \begin{figure}
766     % \centering
767     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
768     % \caption{}
769 mlamichh 4162
770 gezelter 4167 % \label{fig:barGraph2}
771     % \end{figure}
772 gezelter 4174 %The correlation coefficient ($R^2$) and slope of the linear
773     %regression plots for the energy differences for all six different
774     %molecular systems is shown in figure 4a and 4b.The plot shows that
775     %the correlation coefficient improves for the SP cutoff method as
776     %compared to the undamped hard cutoff method in the case of SSDQC,
777     %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
778     %crystal and liquid, the correlation coefficient is almost unchanged
779     %and close to 1. The correlation coefficient is smallest (0.696276
780     %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
781     %charge-charge and charge-multipole interactions. Since the
782     %charge-charge and charge-multipole interaction is long ranged, there
783     %is huge deviation of correlation coefficient from 1. Similarly, the
784     %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
785     %compared to interactions in the other multipolar systems, thus the
786     %correlation coefficient very close to 1 even for hard cutoff
787     %method. The idea of placing image multipole on the surface of the
788     %cutoff sphere improves the correlation coefficient and makes it close
789     %to 1 for all types of multipolar systems. Similarly the slope is
790     %hugely deviated from the correct value for the lower order
791     %multipole-multipole interaction and slightly deviated for higher
792     %order multipole – multipole interaction. The SP method improves both
793     %correlation coefficient ($R^2$) and slope significantly in SSDQC and
794     %dipolar systems. The Slope is found to be deviated more in dipolar
795     %crystal as compared to liquid which is associated with the large
796     %fluctuation in the electrostatic energy in crystal. The GSF also
797     %produced better values of correlation coefficient and slope with the
798     %proper selection of the damping alpha (Interested reader can consult
799     %accompanying supporting material). The TSF method gives good value of
800     %correlation coefficient for the dipolar crystal, dipolar liquid,
801     %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
802     %regression slopes are significantly deviated.
803    
804 mlamichh 4114 \begin{figure}
805 gezelter 4174 \centering
806     \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
807     \caption{Statistical analysis of the quality of configurational
808     energy differences for the real-space electrostatic methods
809     compared with the reference Ewald sum. Results with a value equal
810     to 1 (dashed line) indicate $\Delta E$ values indistinguishable
811     from those obtained using the multipolar Ewald sum. Different
812     values of the cutoff radius are indicated with different symbols
813     (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
814     triangles).}
815     \label{fig:slopeCorr_energy}
816     \end{figure}
817    
818     The combined correlation coefficient and slope for all six systems is
819     shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
820 gezelter 4175 reproduce the Ewald configurational energy differences with remarkable
821     fidelity. Undamped hard cutoffs introduce a significant amount of
822     random scatter in the energy differences which is apparent in the
823     reduced value of the correlation coefficient for this method. This
824     can be easily understood as configurations which exhibit small
825     traversals of a few dipoles or quadrupoles out of the cutoff sphere
826     will see large energy jumps when hard cutoffs are used. The
827 gezelter 4174 orientations of the multipoles (particularly in the ordered crystals)
828 gezelter 4175 mean that these energy jumps can go in either direction, producing a
829     significant amount of random scatter, but no systematic error.
830 gezelter 4174
831     The TSF method produces energy differences that are highly correlated
832     with the Ewald results, but it also introduces a significant
833     systematic bias in the values of the energies, particularly for
834     smaller cutoff values. The TSF method alters the distance dependence
835     of different orientational contributions to the energy in a
836     non-uniform way, so the size of the cutoff sphere can have a large
837 gezelter 4175 effect, particularly for the crystalline systems.
838 gezelter 4174
839     Both the SP and GSF methods appear to reproduce the Ewald results with
840     excellent fidelity, particularly for moderate damping ($\alpha =
841 gezelter 4175 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
842     12$\AA). With the exception of the undamped hard cutoff, and the TSF
843     method with short cutoffs, all of the methods would be appropriate for
844     use in Monte Carlo simulations.
845 gezelter 4174
846 mlamichh 4114 \subsection{Magnitude of the force and torque vectors}
847 gezelter 4174
848 gezelter 4175 The comparisons of the magnitudes of the forces and torques for the
849     data accumulated from all six systems are shown in Figures
850 gezelter 4174 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
851     correlation and slope for the forces agree well with the Ewald sum
852 gezelter 4175 even for the hard cutoffs.
853 gezelter 4174
854 gezelter 4175 For systems of molecules with only multipolar interactions, the pair
855     energy contributions are quite short ranged. Moreover, the force
856     decays more rapidly than the electrostatic energy, hence the hard
857     cutoff method can also produce reasonable agreement for this quantity.
858     Although the pure cutoff gives reasonably good electrostatic forces
859     for pairs of molecules included within each other's cutoff spheres,
860     the discontinuity in the force at the cutoff radius can potentially
861     cause energy conservation problems as molecules enter and leave the
862     cutoff spheres. This is discussed in detail in section
863     \ref{sec:conservation}.
864 gezelter 4174
865     The two shifted-force methods (GSF and TSF) exhibit a small amount of
866     systematic variation and scatter compared with the Ewald forces. The
867     shifted-force models intentionally perturb the forces between pairs of
868 gezelter 4175 molecules inside each other's cutoff spheres in order to correct the
869     energy conservation issues, and this perturbation is evident in the
870     statistics accumulated for the molecular forces. The GSF
871 gezelter 4180 perturbations are minimal, particularly for moderate damping and
872 gezelter 4174 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
873     reasonable agreement in the correlation coefficient but again the
874     systematic error in the forces is concerning if replication of Ewald
875     forces is desired.
876    
877 mlamichh 4114 \begin{figure}
878 gezelter 4174 \centering
879     \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
880     \caption{Statistical analysis of the quality of the force vector
881     magnitudes for the real-space electrostatic methods compared with
882     the reference Ewald sum. Results with a value equal to 1 (dashed
883     line) indicate force magnitude values indistinguishable from those
884     obtained using the multipolar Ewald sum. Different values of the
885     cutoff radius are indicated with different symbols (9\AA\ =
886     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
887     \label{fig:slopeCorr_force}
888     \end{figure}
889    
890    
891 mlamichh 4114 \begin{figure}
892 gezelter 4174 \centering
893     \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
894     \caption{Statistical analysis of the quality of the torque vector
895     magnitudes for the real-space electrostatic methods compared with
896     the reference Ewald sum. Results with a value equal to 1 (dashed
897     line) indicate force magnitude values indistinguishable from those
898     obtained using the multipolar Ewald sum. Different values of the
899     cutoff radius are indicated with different symbols (9\AA\ =
900     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
901     \label{fig:slopeCorr_torque}
902     \end{figure}
903    
904     The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
905     significantly influenced by the choice of real-space method. The
906     torque expressions have the same distance dependence as the energies,
907     which are naturally longer-ranged expressions than the inter-site
908     forces. Torques are also quite sensitive to orientations of
909     neighboring molecules, even those that are near the cutoff distance.
910    
911     The results shows that the torque from the hard cutoff method
912     reproduces the torques in quite good agreement with the Ewald sum.
913 gezelter 4175 The other real-space methods can cause some deviations, but excellent
914     agreement with the Ewald sum torques is recovered at moderate values
915     of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
916     radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
917     in the slope when compared with the Ewald torques even for larger
918     cutoff radii. It appears that the severity of the perturbations in
919     the TSF method are most in evidence for the torques.
920 gezelter 4174
921 mlamichh 4114 \subsection{Directionality of the force and torque vectors}
922 mlamichh 4162
923 gezelter 4174 The accurate evaluation of force and torque directions is just as
924     important for molecular dynamics simulations as the magnitudes of
925     these quantities. Force and torque vectors for all six systems were
926     analyzed using Fisher statistics, and the quality of the vector
927     directionality is shown in terms of circular variance
928 gezelter 4180 ($\mathrm{Var}(\theta)$) in figure
929 gezelter 4174 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
930 gezelter 4175 from the new real-space methods exhibit nearly-ideal Fisher probability
931 gezelter 4174 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
932     exhibit the best vectorial agreement with the Ewald sum. The force and
933     torque vectors from GSF method also show good agreement with the Ewald
934     method, which can also be systematically improved by using moderate
935 gezelter 4175 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
936 gezelter 4174 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
937 gezelter 4175 to a distribution with 95\% of force vectors within $6.37^\circ$ of
938     the corresponding Ewald forces. The TSF method produces the poorest
939 gezelter 4174 agreement with the Ewald force directions.
940    
941 gezelter 4175 Torques are again more perturbed than the forces by the new real-space
942     methods, but even here the variance is reasonably small. For the same
943 gezelter 4174 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
944     the circular variance was 0.01415, corresponds to a distribution which
945     has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
946     results. Again, the direction of the force and torque vectors can be
947     systematically improved by varying $\alpha$ and $r_c$.
948    
949 mlamichh 4114 \begin{figure}
950 gezelter 4174 \centering
951     \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
952     \caption{The circular variance of the direction of the force and
953     torque vectors obtained from the real-space methods around the
954     reference Ewald vectors. A variance equal to 0 (dashed line)
955     indicates direction of the force or torque vectors are
956     indistinguishable from those obtained from the Ewald sum. Here
957     different symbols represent different values of the cutoff radius
958     (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
959     \label{fig:slopeCorr_circularVariance}
960     \end{figure}
961 gezelter 4171
962 gezelter 4175 \subsection{Energy conservation\label{sec:conservation}}
963 gezelter 4171
964 gezelter 4174 We have tested the conservation of energy one can expect to see with
965     the new real-space methods using the SSDQ water model with a small
966     fraction of solvated ions. This is a test system which exercises all
967     orders of multipole-multipole interactions derived in the first paper
968     in this series and provides the most comprehensive test of the new
969     methods. A liquid-phase system was created with 2000 water molecules
970 gezelter 4175 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
971 gezelter 4174 temperature of 300K. After equilibration, this liquid-phase system
972     was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
973 gezelter 4175 a cutoff radius of 12\AA. The value of the damping coefficient was
974 gezelter 4174 also varied from the undamped case ($\alpha = 0$) to a heavily damped
975 gezelter 4175 case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods. A
976     sample was also run using the multipolar Ewald sum with the same
977     real-space cutoff.
978 gezelter 4174
979     In figure~\ref{fig:energyDrift} we show the both the linear drift in
980     energy over time, $\delta E_1$, and the standard deviation of energy
981     fluctuations around this drift $\delta E_0$. Both of the
982     shifted-force methods (GSF and TSF) provide excellent energy
983 gezelter 4181 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
984 gezelter 4174 while the hard cutoff is essentially unusable for molecular dynamics.
985     SP provides some benefit over the hard cutoff because the energetic
986     jumps that happen as particles leave and enter the cutoff sphere are
987 gezelter 4175 somewhat reduced, but like the Wolf method for charges, the SP method
988     would not be as useful for molecular dynamics as either of the
989     shifted-force methods.
990 gezelter 4174
991     We note that for all tested values of the cutoff radius, the new
992     real-space methods can provide better energy conservation behavior
993     than the multipolar Ewald sum, even when utilizing a relatively large
994     $k$-space cutoff values.
995    
996 mlamichh 4114 \begin{figure}
997 gezelter 4171 \centering
998 gezelter 4180 \includegraphics[width=\textwidth]{newDrift_12.pdf}
999 mlamichh 4162 \label{fig:energyDrift}
1000 gezelter 4174 \caption{Analysis of the energy conservation of the real-space
1001 gezelter 4171 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
1002 gezelter 4180 energy over time (in kcal / mol / particle / ns) and $\delta
1003     \mathrm{E}_0$ is the standard deviation of energy fluctuations
1004     around this drift (in kcal / mol / particle). All simulations were
1005     of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
1006     300 K starting from the same initial configuration. All runs
1007     utilized the same real-space cutoff, $r_c = 12$\AA.}
1008 gezelter 4171 \end{figure}
1009    
1010 gezelter 4174
1011 mlamichh 4114 \section{CONCLUSION}
1012 gezelter 4175 In the first paper in this series, we generalized the
1013     charge-neutralized electrostatic energy originally developed by Wolf
1014     \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1015     up to quadrupolar order. The SP method is essentially a
1016     multipole-capable version of the Wolf model. The SP method for
1017     multipoles provides excellent agreement with Ewald-derived energies,
1018     forces and torques, and is suitable for Monte Carlo simulations,
1019     although the forces and torques retain discontinuities at the cutoff
1020     distance that prevents its use in molecular dynamics.
1021 gezelter 4170
1022 gezelter 4175 We also developed two natural extensions of the damped shifted-force
1023     (DSF) model originally proposed by Fennel and
1024     Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1025     smooth truncation of energies, forces, and torques at the real-space
1026     cutoff, and both converge to DSF electrostatics for point-charge
1027     interactions. The TSF model is based on a high-order truncated Taylor
1028     expansion which can be relatively perturbative inside the cutoff
1029     sphere. The GSF model takes the gradient from an images of the
1030     interacting multipole that has been projected onto the cutoff sphere
1031     to derive shifted force and torque expressions, and is a significantly
1032     more gentle approach.
1033 gezelter 4170
1034 gezelter 4175 Of the two newly-developed shifted force models, the GSF method
1035     produced quantitative agreement with Ewald energy, force, and torques.
1036     It also performs well in conserving energy in MD simulations. The
1037     Taylor-shifted (TSF) model provides smooth dynamics, but these take
1038     place on a potential energy surface that is significantly perturbed
1039     from Ewald-based electrostatics.
1040    
1041     % The direct truncation of any electrostatic potential energy without
1042     % multipole neutralization creates large fluctuations in molecular
1043     % simulations. This fluctuation in the energy is very large for the case
1044     % of crystal because of long range of multipole ordering (Refer paper
1045     % I).\cite{PaperI} This is also significant in the case of the liquid
1046     % because of the local multipole ordering in the molecules. If the net
1047     % multipole within cutoff radius neutralized within cutoff sphere by
1048     % placing image multiples on the surface of the sphere, this fluctuation
1049     % in the energy reduced significantly. Also, the multipole
1050     % neutralization in the generalized SP method showed very good agreement
1051     % with the Ewald as compared to direct truncation for the evaluation of
1052     % the $\triangle E$ between the configurations. In MD simulations, the
1053     % energy conservation is very important. The conservation of the total
1054     % energy can be ensured by i) enforcing the smooth truncation of the
1055     % energy, force and torque in the cutoff radius and ii) making the
1056     % energy, force and torque consistent with each other. The GSF and TSF
1057     % methods ensure the consistency and smooth truncation of the energy,
1058     % force and torque at the cutoff radius, as a result show very good
1059     % total energy conservation. But the TSF method does not show good
1060     % agreement in the absolute value of the electrostatic energy, force and
1061     % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1062     % energy and torque accurately and also conserved energy.
1063    
1064     The only cases we have found where the new GSF and SP real-space
1065     methods can be problematic are those which retain a bulk dipole moment
1066     at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1067     materials, uniform weighting of the orientational contributions can be
1068     important for converging the total energy. In these cases, the
1069     damping function which causes the non-uniform weighting can be
1070     replaced by the bare electrostatic kernel, and the energies return to
1071     the expected converged values.
1072    
1073     Based on the results of this work, the GSF method is a suitable and
1074     efficient replacement for the Ewald sum for evaluating electrostatic
1075     interactions in MD simulations. Both methods retain excellent
1076     fidelity to the Ewald energies, forces and torques. Additionally, the
1077     energy drift and fluctuations from the GSF electrostatics are better
1078     than a multipolar Ewald sum for finite-sized reciprocal spaces.
1079     Because they use real-space cutoffs with moderate cutoff radii, the
1080     GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1081     increases. Additionally, they can be made extremely efficient using
1082     spline interpolations of the radial functions. They require no
1083     Fourier transforms or $k$-space sums, and guarantee the smooth
1084     handling of energies, forces, and torques as multipoles cross the
1085     real-space cutoff boundary.
1086    
1087 gezelter 4180 \begin{acknowledgments}
1088     JDG acknowledges helpful discussions with Christopher
1089     Fennell. Support for this project was provided by the National
1090     Science Foundation under grant CHE-1362211. Computational time was
1091     provided by the Center for Research Computing (CRC) at the
1092     University of Notre Dame.
1093     \end{acknowledgments}
1094    
1095 gezelter 4167 %\bibliographystyle{aip}
1096 gezelter 4168 \newpage
1097 mlamichh 4114 \bibliography{references}
1098     \end{document}
1099    
1100     %
1101     % ****** End of file aipsamp.tex ******