ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4188
Committed: Sun Jun 15 16:27:42 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 62193 byte(s)
Log Message:
Fix

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 4186 \title{Real space alternatives to the Ewald Sum. II. Comparison of Methods}
51 mlamichh 4114
52     \author{Madan Lamichhane}
53 gezelter 4186 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
54 mlamichh 4114
55     \author{Kathie E. Newman}
56 gezelter 4186 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
57 mlamichh 4114
58     \author{J. Daniel Gezelter}%
59     \email{gezelter@nd.edu.}
60 gezelter 4186 \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
61     }
62 mlamichh 4114
63 gezelter 4186 \date{\today}
64 mlamichh 4114
65     \begin{abstract}
66 gezelter 4187 We report on tests of the shifted potential (SP), gradient shifted
67     force (GSF), and Taylor shifted force (TSF) real-space methods for
68     multipole interactions developed in the first paper in this series,
69     using the multipolar Ewald sum as a reference method. The tests were
70     carried out in a variety of condensed-phase environments designed to
71     test up to quadrupole-quadrupole interactions. Comparisons of the
72     energy differences between configurations, molecular forces, and
73     torques were used to analyze how well the real-space models perform
74     relative to the more computationally expensive Ewald treatment. We
75     have also investigated the energy conservation properties of the new
76     methods in molecular dynamics simulations. The SP method shows
77     excellent agreement with configurational energy differences, forces,
78     and torques, and would be suitable for use in Monte Carlo
79     calculations. Of the two new shifted-force methods, the GSF
80     approach shows the best agreement with Ewald-derived energies,
81     forces, and torques and also exhibits energy conservation properties
82     that make it an excellent choice for efficient computation of
83     electrostatic interactions in molecular dynamics simulations.
84 mlamichh 4114 \end{abstract}
85    
86 gezelter 4175 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
87 mlamichh 4114 % Classification Scheme.
88 gezelter 4184 %\keywords{Electrostatics, Multipoles, Real-space}
89 gezelter 4167
90 mlamichh 4114 \maketitle
91    
92 mlamichh 4166 \section{\label{sec:intro}Introduction}
93 gezelter 4167 Computing the interactions between electrostatic sites is one of the
94 gezelter 4185 most expensive aspects of molecular simulations. There have been
95     significant efforts to develop practical, efficient and convergent
96     methods for handling these interactions. Ewald's method is perhaps the
97     best known and most accurate method for evaluating energies, forces,
98     and torques in explicitly-periodic simulation cells. In this approach,
99     the conditionally convergent electrostatic energy is converted into
100     two absolutely convergent contributions, one which is carried out in
101     real space with a cutoff radius, and one in reciprocal
102 gezelter 4187 space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
103 mlamichh 4114
104 gezelter 4167 When carried out as originally formulated, the reciprocal-space
105     portion of the Ewald sum exhibits relatively poor computational
106 gezelter 4187 scaling, making it prohibitive for large systems. By utilizing a
107     particle mesh and three dimensional fast Fourier transforms (FFT), the
108     particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
109 gezelter 4186 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
110     methods can decrease the computational cost from $O(N^2)$ down to $O(N
111     \log
112 gezelter 4187 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
113 gezelter 4167
114 gezelter 4185 Because of the artificial periodicity required for the Ewald sum,
115 gezelter 4167 interfacial molecular systems such as membranes and liquid-vapor
116 gezelter 4187 interfaces require modifications to the method. Parry's extension of
117     the three dimensional Ewald sum is appropriate for slab
118     geometries.\cite{Parry:1975if} Modified Ewald methods that were
119     developed to handle two-dimensional (2-D) electrostatic
120     interactions,\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
121     but these methods were originally quite computationally
122     expensive.\cite{Spohr97,Yeh99} There have been several successful
123     efforts that reduced the computational cost of 2-D lattice
124     summations,\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04}
125     bringing them more in line with the scaling for the full 3-D
126 gezelter 4188 treatments. The inherent periodicity in the Ewald method can also
127 gezelter 4187 be problematic for interfacial molecular
128     systems.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
129 gezelter 4167
130 mlamichh 4166 \subsection{Real-space methods}
131 gezelter 4168 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
132     method for calculating electrostatic interactions between point
133 gezelter 4185 charges. They argued that the effective Coulomb interaction in most
134     condensed phase systems is effectively short
135     ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
136     computing the Madelung constant of an ionic solid), the material can
137     be considered as a set of ions interacting with neutral dipolar or
138     quadrupolar ``molecules'' giving an effective distance dependence for
139     the electrostatic interactions of $r^{-5}$ (see figure
140     \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
141     cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
142     electrostatic energy per ion converges more rapidly to the Madelung
143     energy than the dipolar approximation.\cite{Wolf92} To find the
144     correct Madelung constant, Lacman suggested that the NaCl structure
145     could be constructed in a way that the finite crystal terminates with
146     complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
147     what is effectively a set of octupoles at large distances. These facts
148     suggest that the Madelung constants are relatively short ranged for
149     perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
150 gezelter 4186 application of Wolf's method can provide accurate estimates of
151 gezelter 4185 Madelung constants using relatively short cutoff radii.
152    
153     Direct truncation of interactions at a cutoff radius creates numerical
154 gezelter 4186 errors. Wolf \textit{et al.} suggest that truncation errors are due
155 gezelter 4185 to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
156     neutralize this charge they proposed placing an image charge on the
157     surface of the cutoff sphere for every real charge inside the cutoff.
158     These charges are present for the evaluation of both the pair
159     interaction energy and the force, although the force expression
160 gezelter 4186 maintains a discontinuity at the cutoff sphere. In the original Wolf
161 gezelter 4185 formulation, the total energy for the charge and image were not equal
162 gezelter 4186 to the integral of the force expression, and as a result, the total
163 gezelter 4185 energy would not be conserved in molecular dynamics (MD)
164     simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
165     Gezelter later proposed shifted force variants of the Wolf method with
166     commensurate force and energy expressions that do not exhibit this
167 gezelter 4186 problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
168     were also proposed by Chen \textit{et
169 gezelter 4185 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
170 gezelter 4186 and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
171     used additional neutralization of higher order moments for systems of
172     point charges.\cite{Fukuda:2013sf}
173 mlamichh 4114
174 gezelter 4181 \begin{figure}
175 gezelter 4167 \centering
176 gezelter 4181 \includegraphics[width=\linewidth]{schematic.pdf}
177     \caption{Top: Ionic systems exhibit local clustering of dissimilar
178     charges (in the smaller grey circle), so interactions are
179 gezelter 4184 effectively charge-multipole at longer distances. With hard
180     cutoffs, motion of individual charges in and out of the cutoff
181     sphere can break the effective multipolar ordering. Bottom:
182     dipolar crystals and fluids have a similar effective
183 gezelter 4181 \textit{quadrupolar} ordering (in the smaller grey circles), and
184     orientational averaging helps to reduce the effective range of the
185     interactions in the fluid. Placement of reversed image multipoles
186     on the surface of the cutoff sphere recovers the effective
187     higher-order multipole behavior.}
188     \label{fig:schematic}
189 gezelter 4167 \end{figure}
190 mlamichh 4114
191 gezelter 4185 One can make a similar effective range argument for crystals of point
192     \textit{multipoles}. The Luttinger and Tisza treatment of energy
193     constants for dipolar lattices utilizes 24 basis vectors that contain
194 gezelter 4186 dipoles at the eight corners of a unit cube.\cite{LT} Only three of
195     these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
196 gezelter 4185 moments, while the rest have zero net dipole and retain contributions
197 gezelter 4186 only from higher order multipoles. The lowest-energy crystalline
198 gezelter 4185 structures are built out of basis vectors that have only residual
199     quadrupolar moments (e.g. the $Z_5$ array). In these low energy
200     structures, the effective interaction between a dipole at the center
201 gezelter 4167 of a crystal and a group of eight dipoles farther away is
202     significantly shorter ranged than the $r^{-3}$ that one would expect
203     for raw dipole-dipole interactions. Only in crystals which retain a
204     bulk dipole moment (e.g. ferroelectrics) does the analogy with the
205     ionic crystal break down -- ferroelectric dipolar crystals can exist,
206     while ionic crystals with net charge in each unit cell would be
207     unstable.
208    
209     In ionic crystals, real-space truncation can break the effective
210 gezelter 4181 multipolar arrangements (see Fig. \ref{fig:schematic}), causing
211     significant swings in the electrostatic energy as individual ions move
212     back and forth across the boundary. This is why the image charges are
213 gezelter 4180 necessary for the Wolf sum to exhibit rapid convergence. Similarly,
214     the real-space truncation of point multipole interactions breaks
215     higher order multipole arrangements, and image multipoles are required
216     for real-space treatments of electrostatic energies.
217 gezelter 4167
218 gezelter 4181 The shorter effective range of electrostatic interactions is not
219     limited to perfect crystals, but can also apply in disordered fluids.
220 gezelter 4186 Even at elevated temperatures, there is local charge balance in an
221     ionic liquid, where each positive ion has surroundings dominated by
222     negaitve ions and vice versa. The reversed-charge images on the
223     cutoff sphere that are integral to the Wolf and DSF approaches retain
224     the effective multipolar interactions as the charges traverse the
225     cutoff boundary.
226 gezelter 4181
227     In multipolar fluids (see Fig. \ref{fig:schematic}) there is
228     significant orientational averaging that additionally reduces the
229     effect of long-range multipolar interactions. The image multipoles
230     that are introduced in the TSF, GSF, and SP methods mimic this effect
231     and reduce the effective range of the multipolar interactions as
232     interacting molecules traverse each other's cutoff boundaries.
233    
234 gezelter 4167 % Because of this reason, although the nature of electrostatic
235     % interaction short ranged, the hard cutoff sphere creates very large
236     % fluctuation in the electrostatic energy for the perfect crystal. In
237     % addition, the charge neutralized potential proposed by Wolf et
238     % al. converged to correct Madelung constant but still holds oscillation
239     % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
240     % oscillation in the energy around its fully converged value can be due
241     % to the non-neutralized value of the higher order moments within the
242     % cutoff sphere.
243    
244 gezelter 4186 Forces and torques acting on atomic sites are fundamental in driving
245     dynamics in molecular simulations, and the damped shifted force (DSF)
246     energy kernel provides consistent energies and forces on charged atoms
247     within the cutoff sphere. Both the energy and the force go smoothly to
248     zero as an atom aproaches the cutoff radius. The comparisons of the
249     accuracy these quantities between the DSF kernel and SPME was
250     surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
251     has seen increasing use in molecular systems with relatively uniform
252     charge
253     densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
254 gezelter 4167
255 gezelter 4168 \subsection{The damping function}
256 gezelter 4185 The damping function has been discussed in detail in the first paper
257     of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
258     interactions between point charges can be replaced by the
259     complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
260     convergence, where $\alpha$ is a damping parameter with units of
261     inverse distance. Altering the value of $\alpha$ is equivalent to
262     changing the width of Gaussian charge distributions that replace each
263     point charge, as Coulomb integrals with Gaussian charge distributions
264     produce complementary error functions when truncated at a finite
265     distance.
266 mlamichh 4114
267 gezelter 4185 With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
268     produced very good agreement with SPME for interaction energies,
269     forces and torques for charge-charge
270     interactions.\cite{Fennell:2006lq}
271 gezelter 4167
272 gezelter 4168 \subsection{Point multipoles in molecular modeling}
273     Coarse-graining approaches which treat entire molecular subsystems as
274     a single rigid body are now widely used. A common feature of many
275     coarse-graining approaches is simplification of the electrostatic
276     interactions between bodies so that fewer site-site interactions are
277 gezelter 4185 required to compute configurational
278     energies.\cite{Ren06,Essex10,Essex11}
279 mlamichh 4166
280 gezelter 4186 Additionally, because electrons in a molecule are not localized at
281     specific points, the assignment of partial charges to atomic centers
282     is always an approximation. For increased accuracy, atomic sites can
283     also be assigned point multipoles and polarizabilities. Recently,
284     water has been modeled with point multipoles up to octupolar order
285     using the soft sticky dipole-quadrupole-octupole (SSDQO)
286 gezelter 4180 model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
287 gezelter 4168 multipoles up to quadrupolar order have also been coupled with point
288     polarizabilities in the high-quality AMOEBA and iAMOEBA water
289 gezelter 4185 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
290     truncating point multipoles without smoothing the forces and torques
291 gezelter 4186 can create energy conservation issues in molecular dynamics
292     simulations.
293 mlamichh 4166
294 gezelter 4168 In this paper we test a set of real-space methods that were developed
295     for point multipolar interactions. These methods extend the damped
296     shifted force (DSF) and Wolf methods originally developed for
297     charge-charge interactions and generalize them for higher order
298 gezelter 4186 multipoles. The detailed mathematical development of these methods
299     has been presented in the first paper in this series, while this work
300     covers the testing of energies, forces, torques, and energy
301 gezelter 4168 conservation properties of the methods in realistic simulation
302     environments. In all cases, the methods are compared with the
303 gezelter 4186 reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
304 gezelter 4168
305    
306 mlamichh 4166 %\subsection{Conservation of total energy }
307 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.
308 mlamichh 4166
309 gezelter 4168 \section{\label{sec:method}Review of Methods}
310     Any real-space electrostatic method that is suitable for MD
311     simulations should have the electrostatic energy, forces and torques
312     between two sites go smoothly to zero as the distance between the
313     sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
314     this continuity at the cutoff is essential for energy conservation in
315     MD simulations. The mathematical details of the shifted potential
316     (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
317     methods have been discussed in detail in the previous paper in this
318     series.\cite{PaperI} Here we briefly review the new methods and
319     describe their essential features.
320 mlamichh 4166
321 gezelter 4168 \subsection{Taylor-shifted force (TSF)}
322 mlamichh 4114
323 gezelter 4168 The electrostatic potential energy between point multipoles can be
324     expressed as the product of two multipole operators and a Coulombic
325     kernel,
326 mlamichh 4114 \begin{equation}
327 gezelter 4168 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
328 mlamichh 4114 \end{equation}
329 gezelter 4180 where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
330     expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
331     a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
332 gezelter 4184 $\bf a$, etc.
333 mlamichh 4166
334 gezelter 4180 % Interactions between multipoles can be expressed as higher derivatives
335     % of the bare Coulomb potential, so one way of ensuring that the forces
336     % and torques vanish at the cutoff distance is to include a larger
337     % number of terms in the truncated Taylor expansion, e.g.,
338     % %
339     % \begin{equation}
340     % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
341     % \end{equation}
342     % %
343     % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344     % Thus, for $f(r)=1/r$, we find
345     % %
346     % \begin{equation}
347     % 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} .
348     % \end{equation}
349     % This function is an approximate electrostatic potential that has
350     % vanishing second derivatives at the cutoff radius, making it suitable
351     % for shifting the forces and torques of charge-dipole interactions.
352 gezelter 4168
353 gezelter 4180 The TSF potential for any multipole-multipole interaction can be
354     written
355 gezelter 4168 \begin{equation}
356     U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
357     \label{generic}
358     \end{equation}
359 gezelter 4180 where $f_n(r)$ is a shifted kernel that is appropriate for the order
360 gezelter 4181 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
361     charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
362     and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
363     quadrupole-quadrupole. To ensure smooth convergence of the energy,
364     force, and torques, a Taylor expansion with $n$ terms must be
365     performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
366 gezelter 4168
367 gezelter 4180 % To carry out the same procedure for a damped electrostatic kernel, we
368     % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369     % Many of the derivatives of the damped kernel are well known from
370     % Smith's early work on multipoles for the Ewald
371     % summation.\cite{Smith82,Smith98}
372 gezelter 4168
373 gezelter 4180 % Note that increasing the value of $n$ will add additional terms to the
374     % electrostatic potential, e.g., $f_2(r)$ includes orders up to
375     % $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
376     % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377     % f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
378     % for computing multipole energies, forces, and torques, and smooth
379     % cutoffs of these quantities can be guaranteed as long as the number of
380     % terms in the Taylor series exceeds the derivative order required.
381 gezelter 4168
382     For multipole-multipole interactions, following this procedure results
383 gezelter 4180 in separate radial functions for each of the distinct orientational
384     contributions to the potential, and ensures that the forces and
385     torques from each of these contributions will vanish at the cutoff
386     radius. For example, the direct dipole dot product
387     ($\mathbf{D}_{\bf a}
388     \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
389 gezelter 4168 dot products:
390     \begin{equation}
391 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
392     \mathbf{D}_{\bf a} \cdot
393     \mathbf{D}_{\bf b} \right) v_{21}(r) +
394     \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
395     \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
396 gezelter 4168 \end{equation}
397    
398 gezelter 4180 For the Taylor shifted (TSF) method with the undamped kernel,
399     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
400     \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
401     - \frac{6}{r r_c^2}$. In these functions, one can easily see the
402     connection to unmodified electrostatics as well as the smooth
403     transition to zero in both these functions as $r\rightarrow r_c$. The
404     electrostatic forces and torques acting on the central multipole due
405 gezelter 4184 to another site within the cutoff sphere are derived from
406 gezelter 4168 Eq.~\ref{generic}, accounting for the appropriate number of
407     derivatives. Complete energy, force, and torque expressions are
408     presented in the first paper in this series (Reference
409 gezelter 4175 \onlinecite{PaperI}).
410 gezelter 4168
411     \subsection{Gradient-shifted force (GSF)}
412    
413 gezelter 4180 A second (and conceptually simpler) method involves shifting the
414     gradient of the raw Coulomb potential for each particular multipole
415 gezelter 4168 order. For example, the raw dipole-dipole potential energy may be
416     shifted smoothly by finding the gradient for two interacting dipoles
417     which have been projected onto the surface of the cutoff sphere
418     without changing their relative orientation,
419 gezelter 4181 \begin{equation}
420 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r) = U_{D_{\bf a}D_{\bf b}}(r) -
421     U_{D_{\bf a} D_{\bf b}}(r_c)
422     - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
423 gezelter 4184 \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
424 gezelter 4181 \end{equation}
425 gezelter 4180 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
426     a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
427     (although the signs are reversed for the dipole that has been
428     projected onto the cutoff sphere). In many ways, this simpler
429     approach is closer in spirit to the original shifted force method, in
430     that it projects a neutralizing multipole (and the resulting forces
431     from this multipole) onto a cutoff sphere. The resulting functional
432     forms for the potentials, forces, and torques turn out to be quite
433     similar in form to the Taylor-shifted approach, although the radial
434     contributions are significantly less perturbed by the gradient-shifted
435     approach than they are in the Taylor-shifted method.
436 gezelter 4168
437 gezelter 4180 For the gradient shifted (GSF) method with the undamped kernel,
438     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
439     $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
440     Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
441     because the Taylor expansion retains only one term, they are
442     significantly less perturbed than the TSF functions.
443    
444 gezelter 4168 In general, the gradient shifted potential between a central multipole
445     and any multipolar site inside the cutoff radius is given by,
446     \begin{equation}
447 gezelter 4184 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
448     U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
449     \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
450 gezelter 4168 \label{generic2}
451     \end{equation}
452     where the sum describes a separate force-shifting that is applied to
453 gezelter 4184 each orientational contribution to the energy. In this expression,
454     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
455     ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
456     represent the orientations the multipoles.
457 gezelter 4168
458     The third term converges more rapidly than the first two terms as a
459     function of radius, hence the contribution of the third term is very
460     small for large cutoff radii. The force and torque derived from
461 gezelter 4184 Eq. \ref{generic2} are consistent with the energy expression and
462 gezelter 4175 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
463 gezelter 4168 can be considered generalizations of the original DSF method for
464     higher order multipole interactions. GSF and TSF are also identical up
465     to the charge-dipole interaction but generate different expressions in
466     the energy, force and torque for higher order multipole-multipole
467     interactions. Complete energy, force, and torque expressions for the
468     GSF potential are presented in the first paper in this series
469 gezelter 4184 (Reference~\onlinecite{PaperI}).
470 gezelter 4168
471    
472 mlamichh 4166 \subsection{Shifted potential (SP) }
473 gezelter 4168 A discontinuous truncation of the electrostatic potential at the
474     cutoff sphere introduces a severe artifact (oscillation in the
475     electrostatic energy) even for molecules with the higher-order
476     multipoles.\cite{PaperI} We have also formulated an extension of the
477     Wolf approach for point multipoles by simply projecting the image
478     multipole onto the surface of the cutoff sphere, and including the
479     interactions with the central multipole and the image. This
480     effectively shifts the total potential to zero at the cutoff radius,
481 mlamichh 4166 \begin{equation}
482 gezelter 4180 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
483 gezelter 4184 U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
484 mlamichh 4166 \label{eq:SP}
485     \end{equation}
486 gezelter 4168 where the sum describes separate potential shifting that is done for
487     each orientational contribution to the energy (e.g. the direct dipole
488     product contribution is shifted {\it separately} from the
489     dipole-distance terms in dipole-dipole interactions). Note that this
490 gezelter 4175 is not a simple shifting of the total potential at $r_c$. Each radial
491 gezelter 4168 contribution is shifted separately. One consequence of this is that
492     multipoles that reorient after leaving the cutoff sphere can re-enter
493     the cutoff sphere without perturbing the total energy.
494 mlamichh 4166
495 gezelter 4180 For the shifted potential (SP) method with the undamped kernel,
496     $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
497     \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
498     central multipole and other multipolar sites goes smoothly to zero as
499     $r \rightarrow r_c$. However, the force and torque obtained from the
500     shifted potential (SP) are discontinuous at $r_c$. MD simulations
501     will still experience energy drift while operating under the SP
502     potential, but it may be suitable for Monte Carlo approaches where the
503     configurational energy differences are the primary quantity of
504     interest.
505 gezelter 4168
506 gezelter 4180 \subsection{The Self Term}
507 gezelter 4168 In the TSF, GSF, and SP methods, a self-interaction is retained for
508     the central multipole interacting with its own image on the surface of
509     the cutoff sphere. This self interaction is nearly identical with the
510     self-terms that arise in the Ewald sum for multipoles. Complete
511     expressions for the self terms are presented in the first paper in
512 gezelter 4175 this series (Reference \onlinecite{PaperI}).
513 mlamichh 4162
514 gezelter 4168
515 gezelter 4170 \section{\label{sec:methodology}Methodology}
516 mlamichh 4166
517 gezelter 4170 To understand how the real-space multipole methods behave in computer
518     simulations, it is vital to test against established methods for
519     computing electrostatic interactions in periodic systems, and to
520     evaluate the size and sources of any errors that arise from the
521     real-space cutoffs. In the first paper of this series, we compared
522     the dipolar and quadrupolar energy expressions against analytic
523     expressions for ordered dipolar and quadrupolar
524 gezelter 4174 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
525     used the multipolar Ewald sum as a reference method for comparing
526     energies, forces, and torques for molecular models that mimic
527 gezelter 4175 disordered and ordered condensed-phase systems. The parameters used
528 gezelter 4180 in the test cases are given in table~\ref{tab:pars}.
529 gezelter 4174
530 gezelter 4175 \begin{table}
531     \label{tab:pars}
532     \caption{The parameters used in the systems used to evaluate the new
533     real-space methods. The most comprehensive test was a liquid
534     composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
535     ions). This test excercises all orders of the multipolar
536     interactions developed in the first paper.}
537     \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
538     & \multicolumn{2}{c|}{LJ parameters} &
539     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
540     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
541     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
542     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
543     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
544     \AA\textsuperscript{2})} \\ \hline
545     Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
546 gezelter 4180 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
547 gezelter 4175 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
548 gezelter 4180 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
549 gezelter 4175 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
550     \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
551     \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
552     \end{tabularx}
553     \end{table}
554     The systems consist of pure multipolar solids (both dipole and
555     quadrupole), pure multipolar liquids (both dipole and quadrupole), a
556     fluid composed of sites containing both dipoles and quadrupoles
557     simultaneously, and a final test case that includes ions with point
558     charges in addition to the multipolar fluid. The solid-phase
559     parameters were chosen so that the systems can explore some
560     orientational freedom for the multipolar sites, while maintaining
561     relatively strict translational order. The SSDQ model used here is
562     not a particularly accurate water model, but it does test
563     dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
564     interactions at roughly the same magnitudes. The last test case, SSDQ
565     water with dissolved ions, exercises \textit{all} levels of the
566     multipole-multipole interactions we have derived so far and represents
567     the most complete test of the new methods.
568 mlamichh 4166
569 gezelter 4175 In the following section, we present results for the total
570     electrostatic energy, as well as the electrostatic contributions to
571     the force and torque on each molecule. These quantities have been
572     computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
573 gezelter 4180 and have been compared with the values obtained from the multipolar
574     Ewald sum. In Monte Carlo (MC) simulations, the energy differences
575 gezelter 4175 between two configurations is the primary quantity that governs how
576     the simulation proceeds. These differences are the most imporant
577     indicators of the reliability of a method even if the absolute
578     energies are not exact. For each of the multipolar systems listed
579     above, we have compared the change in electrostatic potential energy
580     ($\Delta E$) between 250 statistically-independent configurations. In
581     molecular dynamics (MD) simulations, the forces and torques govern the
582     behavior of the simulation, so we also compute the electrostatic
583     contributions to the forces and torques.
584    
585     \subsection{Implementation}
586     The real-space methods developed in the first paper in this series
587     have been implemented in our group's open source molecular simulation
588 gezelter 4187 program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
589 gezelter 4175 this work. The complementary error function can be a relatively slow
590     function on some processors, so all of the radial functions are
591     precomputed on a fine grid and are spline-interpolated to provide
592     values when required.
593    
594     Using the same simulation code, we compare to a multipolar Ewald sum
595     with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
596     the Ewald sum is a re-implementation of the algorithm originally
597     proposed by Smith that does not use the particle mesh or smoothing
598     approximations.\cite{Smith82,Smith98} In all cases, the quantities
599     being compared are the electrostatic contributions to energies, force,
600     and torques. All other contributions to these quantities (i.e. from
601     Lennard-Jones interactions) are removed prior to the comparisons.
602    
603     The convergence parameter ($\alpha$) also plays a role in the balance
604     of the real-space and reciprocal-space portions of the Ewald
605     calculation. Typical molecular mechanics packages set this to a value
606     that depends on the cutoff radius and a tolerance (typically less than
607     $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
608     associated with increasing accuracy at the expense of computational
609     time spent on the reciprocal-space portion of the
610     summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
611     10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
612     Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
613    
614     The real-space models have self-interactions that provide
615     contributions to the energies only. Although the self interaction is
616     a rapid calculation, we note that in systems with fluctuating charges
617     or point polarizabilities, the self-term is not static and must be
618     recomputed at each time step.
619    
620 gezelter 4170 \subsection{Model systems}
621 gezelter 4180 To sample independent configurations of the multipolar crystals, body
622     centered cubic (bcc) crystals, which exhibit the minimum energy
623     structures for point dipoles, were generated using 3,456 molecules.
624     The multipoles were translationally locked in their respective crystal
625     sites for equilibration at a relatively low temperature (50K) so that
626     dipoles or quadrupoles could freely explore all accessible
627     orientations. The translational constraints were then removed, the
628     systems were re-equilibrated, and the crystals were simulated for an
629     additional 10 ps in the microcanonical (NVE) ensemble with an average
630     temperature of 50 K. The balance between moments of inertia and
631     particle mass were chosen to allow orientational sampling without
632     significant translational motion. Configurations were sampled at
633     equal time intervals in order to compare configurational energy
634     differences. The crystals were simulated far from the melting point
635     in order to avoid translational deformation away of the ideal lattice
636     geometry.
637 gezelter 4170
638 gezelter 4180 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
639     simulations, each system was created with 2,048 randomly-oriented
640     molecules. These were equilibrated at a temperature of 300K for 1 ns.
641     Each system was then simulated for 1 ns in the microcanonical (NVE)
642     ensemble. We collected 250 different configurations at equal time
643     intervals. For the liquid system that included ionic species, we
644     converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
645     \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
646     run under the same conditions for 1 ns. A total of 250 configurations
647     were collected. In the following comparisons of energies, forces, and
648     torques, the Lennard-Jones potentials were turned off and only the
649     purely electrostatic quantities were compared with the same values
650     obtained via the Ewald sum.
651 gezelter 4170
652     \subsection{Accuracy of Energy Differences, Forces and Torques}
653     The pairwise summation techniques (outlined above) were evaluated for
654     use in MC simulations by studying the energy differences between
655     different configurations. We took the Ewald-computed energy
656     difference between two conformations to be the correct behavior. An
657     ideal performance by one of the new methods would reproduce these
658     energy differences exactly. The configurational energies being used
659     here contain only contributions from electrostatic interactions.
660     Lennard-Jones interactions were omitted from the comparison as they
661     should be identical for all methods.
662    
663     Since none of the real-space methods provide exact energy differences,
664 gezelter 4180 we used least square regressions analysis for the six different
665 gezelter 4170 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
666     with the multipolar Ewald reference method. Unitary results for both
667     the correlation (slope) and correlation coefficient for these
668     regressions indicate perfect agreement between the real-space method
669     and the multipolar Ewald sum.
670    
671     Molecular systems were run long enough to explore independent
672     configurations and 250 configurations were recorded for comparison.
673     Each system provided 31,125 energy differences for a total of 186,750
674     data points. Similarly, the magnitudes of the forces and torques have
675 gezelter 4180 also been compared using least squares regression analysis. In the
676 gezelter 4170 forces and torques comparison, the magnitudes of the forces acting in
677     each molecule for each configuration were evaluated. For example, our
678     dipolar liquid simulation contains 2048 molecules and there are 250
679     different configurations for each system resulting in 3,072,000 data
680     points for comparison of forces and torques.
681    
682 mlamichh 4166 \subsection{Analysis of vector quantities}
683 gezelter 4170 Getting the magnitudes of the force and torque vectors correct is only
684     part of the issue for carrying out accurate molecular dynamics
685     simulations. Because the real space methods reweight the different
686     orientational contributions to the energies, it is also important to
687     understand how the methods impact the \textit{directionality} of the
688     force and torque vectors. Fisher developed a probablity density
689     function to analyse directional data sets,
690 mlamichh 4162 \begin{equation}
691 gezelter 4170 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
692 mlamichh 4162 \label{eq:pdf}
693     \end{equation}
694 gezelter 4170 where $\kappa$ measures directional dispersion of the data around the
695     mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
696     estimated as a reciprocal of the circular variance.\cite{Allen91} To
697     quantify the directional error, forces obtained from the Ewald sum
698     were taken as the mean (or correct) direction and the angle between
699     the forces obtained via the Ewald sum and the real-space methods were
700     evaluated,
701 mlamichh 4162 \begin{equation}
702 gezelter 4170 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
703     \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
704     \end{equation}
705     The total angular displacement of the vectors was calculated as,
706     \begin{equation}
707     R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
708 mlamichh 4162 \label{eq:displacement}
709     \end{equation}
710 gezelter 4170 where $N$ is number of force vectors. The circular variance is
711     defined as
712     \begin{equation}
713     \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
714     \end{equation}
715     The circular variance takes on values between from 0 to 1, with 0
716     indicating a perfect directional match between the Ewald force vectors
717     and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
718     correspond to higher values of $\kappa$, which indicates tighter
719     clustering of the real-space force vectors around the Ewald forces.
720 mlamichh 4162
721 gezelter 4170 A similar analysis was carried out for the electrostatic contribution
722     to the molecular torques as well as forces.
723    
724 mlamichh 4166 \subsection{Energy conservation}
725 gezelter 4170 To test conservation the energy for the methods, the mixed molecular
726     system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
727     ions was run for 1 ns in the microcanonical ensemble at an average
728     temperature of 300K. Each of the different electrostatic methods
729     (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
730     damping values. The molecular system was started with same initial
731     positions and velocities for all cutoff methods. The energy drift
732     ($\delta E_1$) and standard deviation of the energy about the slope
733     ($\delta E_0$) were evaluated from the total energy of the system as a
734     function of time. Although both measures are valuable at
735     investigating new methods for molecular dynamics, a useful interaction
736     model must allow for long simulation times with minimal energy drift.
737 mlamichh 4114
738 mlamichh 4166 \section{\label{sec:result}RESULTS}
739     \subsection{Configurational energy differences}
740     %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.
741     %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.
742     %\begin{figure}[h!]
743     % \centering
744     % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
745     % \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. }
746     % \label{fig:rcutConvergence_dipolarLiquid}
747     % \end{figure}
748     %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.
749     %\begin{figure}
750     % \centering
751     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
752     % \label{fig:barGraph1}
753     % \end{figure}
754     % \begin{figure}
755     % \centering
756     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
757     % \caption{}
758 mlamichh 4162
759 gezelter 4167 % \label{fig:barGraph2}
760     % \end{figure}
761 gezelter 4174 %The correlation coefficient ($R^2$) and slope of the linear
762     %regression plots for the energy differences for all six different
763     %molecular systems is shown in figure 4a and 4b.The plot shows that
764     %the correlation coefficient improves for the SP cutoff method as
765     %compared to the undamped hard cutoff method in the case of SSDQC,
766     %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
767     %crystal and liquid, the correlation coefficient is almost unchanged
768     %and close to 1. The correlation coefficient is smallest (0.696276
769     %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
770     %charge-charge and charge-multipole interactions. Since the
771     %charge-charge and charge-multipole interaction is long ranged, there
772     %is huge deviation of correlation coefficient from 1. Similarly, the
773     %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
774     %compared to interactions in the other multipolar systems, thus the
775     %correlation coefficient very close to 1 even for hard cutoff
776     %method. The idea of placing image multipole on the surface of the
777     %cutoff sphere improves the correlation coefficient and makes it close
778     %to 1 for all types of multipolar systems. Similarly the slope is
779     %hugely deviated from the correct value for the lower order
780     %multipole-multipole interaction and slightly deviated for higher
781     %order multipole – multipole interaction. The SP method improves both
782     %correlation coefficient ($R^2$) and slope significantly in SSDQC and
783     %dipolar systems. The Slope is found to be deviated more in dipolar
784     %crystal as compared to liquid which is associated with the large
785     %fluctuation in the electrostatic energy in crystal. The GSF also
786     %produced better values of correlation coefficient and slope with the
787     %proper selection of the damping alpha (Interested reader can consult
788     %accompanying supporting material). The TSF method gives good value of
789     %correlation coefficient for the dipolar crystal, dipolar liquid,
790     %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
791     %regression slopes are significantly deviated.
792    
793 mlamichh 4114 \begin{figure}
794 gezelter 4174 \centering
795     \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
796     \caption{Statistical analysis of the quality of configurational
797     energy differences for the real-space electrostatic methods
798     compared with the reference Ewald sum. Results with a value equal
799     to 1 (dashed line) indicate $\Delta E$ values indistinguishable
800     from those obtained using the multipolar Ewald sum. Different
801     values of the cutoff radius are indicated with different symbols
802     (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
803     triangles).}
804     \label{fig:slopeCorr_energy}
805     \end{figure}
806    
807     The combined correlation coefficient and slope for all six systems is
808     shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
809 gezelter 4175 reproduce the Ewald configurational energy differences with remarkable
810     fidelity. Undamped hard cutoffs introduce a significant amount of
811     random scatter in the energy differences which is apparent in the
812     reduced value of the correlation coefficient for this method. This
813     can be easily understood as configurations which exhibit small
814     traversals of a few dipoles or quadrupoles out of the cutoff sphere
815     will see large energy jumps when hard cutoffs are used. The
816 gezelter 4174 orientations of the multipoles (particularly in the ordered crystals)
817 gezelter 4175 mean that these energy jumps can go in either direction, producing a
818     significant amount of random scatter, but no systematic error.
819 gezelter 4174
820     The TSF method produces energy differences that are highly correlated
821     with the Ewald results, but it also introduces a significant
822     systematic bias in the values of the energies, particularly for
823     smaller cutoff values. The TSF method alters the distance dependence
824     of different orientational contributions to the energy in a
825     non-uniform way, so the size of the cutoff sphere can have a large
826 gezelter 4175 effect, particularly for the crystalline systems.
827 gezelter 4174
828     Both the SP and GSF methods appear to reproduce the Ewald results with
829     excellent fidelity, particularly for moderate damping ($\alpha =
830 gezelter 4175 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
831     12$\AA). With the exception of the undamped hard cutoff, and the TSF
832     method with short cutoffs, all of the methods would be appropriate for
833     use in Monte Carlo simulations.
834 gezelter 4174
835 mlamichh 4114 \subsection{Magnitude of the force and torque vectors}
836 gezelter 4174
837 gezelter 4175 The comparisons of the magnitudes of the forces and torques for the
838     data accumulated from all six systems are shown in Figures
839 gezelter 4174 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
840     correlation and slope for the forces agree well with the Ewald sum
841 gezelter 4175 even for the hard cutoffs.
842 gezelter 4174
843 gezelter 4175 For systems of molecules with only multipolar interactions, the pair
844     energy contributions are quite short ranged. Moreover, the force
845     decays more rapidly than the electrostatic energy, hence the hard
846     cutoff method can also produce reasonable agreement for this quantity.
847     Although the pure cutoff gives reasonably good electrostatic forces
848     for pairs of molecules included within each other's cutoff spheres,
849     the discontinuity in the force at the cutoff radius can potentially
850     cause energy conservation problems as molecules enter and leave the
851     cutoff spheres. This is discussed in detail in section
852     \ref{sec:conservation}.
853 gezelter 4174
854     The two shifted-force methods (GSF and TSF) exhibit a small amount of
855     systematic variation and scatter compared with the Ewald forces. The
856     shifted-force models intentionally perturb the forces between pairs of
857 gezelter 4175 molecules inside each other's cutoff spheres in order to correct the
858     energy conservation issues, and this perturbation is evident in the
859     statistics accumulated for the molecular forces. The GSF
860 gezelter 4180 perturbations are minimal, particularly for moderate damping and
861 gezelter 4174 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
862     reasonable agreement in the correlation coefficient but again the
863     systematic error in the forces is concerning if replication of Ewald
864     forces is desired.
865    
866 mlamichh 4114 \begin{figure}
867 gezelter 4174 \centering
868     \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
869     \caption{Statistical analysis of the quality of the force vector
870     magnitudes for the real-space electrostatic methods compared with
871     the reference Ewald sum. Results with a value equal to 1 (dashed
872     line) indicate force magnitude values indistinguishable from those
873     obtained using the multipolar Ewald sum. Different values of the
874     cutoff radius are indicated with different symbols (9\AA\ =
875     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
876     \label{fig:slopeCorr_force}
877     \end{figure}
878    
879    
880 mlamichh 4114 \begin{figure}
881 gezelter 4174 \centering
882     \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
883     \caption{Statistical analysis of the quality of the torque vector
884     magnitudes for the real-space electrostatic methods compared with
885     the reference Ewald sum. Results with a value equal to 1 (dashed
886     line) indicate force magnitude values indistinguishable from those
887     obtained using the multipolar Ewald sum. Different values of the
888     cutoff radius are indicated with different symbols (9\AA\ =
889     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
890     \label{fig:slopeCorr_torque}
891     \end{figure}
892    
893     The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
894     significantly influenced by the choice of real-space method. The
895     torque expressions have the same distance dependence as the energies,
896     which are naturally longer-ranged expressions than the inter-site
897     forces. Torques are also quite sensitive to orientations of
898     neighboring molecules, even those that are near the cutoff distance.
899    
900     The results shows that the torque from the hard cutoff method
901     reproduces the torques in quite good agreement with the Ewald sum.
902 gezelter 4175 The other real-space methods can cause some deviations, but excellent
903     agreement with the Ewald sum torques is recovered at moderate values
904     of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
905     radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
906     in the slope when compared with the Ewald torques even for larger
907     cutoff radii. It appears that the severity of the perturbations in
908     the TSF method are most in evidence for the torques.
909 gezelter 4174
910 mlamichh 4114 \subsection{Directionality of the force and torque vectors}
911 mlamichh 4162
912 gezelter 4174 The accurate evaluation of force and torque directions is just as
913     important for molecular dynamics simulations as the magnitudes of
914     these quantities. Force and torque vectors for all six systems were
915     analyzed using Fisher statistics, and the quality of the vector
916     directionality is shown in terms of circular variance
917 gezelter 4180 ($\mathrm{Var}(\theta)$) in figure
918 gezelter 4174 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
919 gezelter 4175 from the new real-space methods exhibit nearly-ideal Fisher probability
920 gezelter 4174 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
921     exhibit the best vectorial agreement with the Ewald sum. The force and
922     torque vectors from GSF method also show good agreement with the Ewald
923     method, which can also be systematically improved by using moderate
924 gezelter 4175 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
925 gezelter 4174 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
926 gezelter 4175 to a distribution with 95\% of force vectors within $6.37^\circ$ of
927     the corresponding Ewald forces. The TSF method produces the poorest
928 gezelter 4174 agreement with the Ewald force directions.
929    
930 gezelter 4175 Torques are again more perturbed than the forces by the new real-space
931     methods, but even here the variance is reasonably small. For the same
932 gezelter 4174 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
933     the circular variance was 0.01415, corresponds to a distribution which
934     has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
935     results. Again, the direction of the force and torque vectors can be
936     systematically improved by varying $\alpha$ and $r_c$.
937    
938 mlamichh 4114 \begin{figure}
939 gezelter 4174 \centering
940     \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
941     \caption{The circular variance of the direction of the force and
942     torque vectors obtained from the real-space methods around the
943     reference Ewald vectors. A variance equal to 0 (dashed line)
944     indicates direction of the force or torque vectors are
945     indistinguishable from those obtained from the Ewald sum. Here
946     different symbols represent different values of the cutoff radius
947     (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
948     \label{fig:slopeCorr_circularVariance}
949     \end{figure}
950 gezelter 4171
951 gezelter 4175 \subsection{Energy conservation\label{sec:conservation}}
952 gezelter 4171
953 gezelter 4174 We have tested the conservation of energy one can expect to see with
954     the new real-space methods using the SSDQ water model with a small
955     fraction of solvated ions. This is a test system which exercises all
956     orders of multipole-multipole interactions derived in the first paper
957     in this series and provides the most comprehensive test of the new
958     methods. A liquid-phase system was created with 2000 water molecules
959 gezelter 4175 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
960 gezelter 4174 temperature of 300K. After equilibration, this liquid-phase system
961     was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
962 gezelter 4175 a cutoff radius of 12\AA. The value of the damping coefficient was
963 gezelter 4174 also varied from the undamped case ($\alpha = 0$) to a heavily damped
964 gezelter 4175 case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods. A
965     sample was also run using the multipolar Ewald sum with the same
966     real-space cutoff.
967 gezelter 4174
968     In figure~\ref{fig:energyDrift} we show the both the linear drift in
969     energy over time, $\delta E_1$, and the standard deviation of energy
970     fluctuations around this drift $\delta E_0$. Both of the
971     shifted-force methods (GSF and TSF) provide excellent energy
972 gezelter 4181 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
973 gezelter 4174 while the hard cutoff is essentially unusable for molecular dynamics.
974     SP provides some benefit over the hard cutoff because the energetic
975     jumps that happen as particles leave and enter the cutoff sphere are
976 gezelter 4175 somewhat reduced, but like the Wolf method for charges, the SP method
977     would not be as useful for molecular dynamics as either of the
978     shifted-force methods.
979 gezelter 4174
980     We note that for all tested values of the cutoff radius, the new
981     real-space methods can provide better energy conservation behavior
982     than the multipolar Ewald sum, even when utilizing a relatively large
983     $k$-space cutoff values.
984    
985 mlamichh 4114 \begin{figure}
986 gezelter 4171 \centering
987 gezelter 4180 \includegraphics[width=\textwidth]{newDrift_12.pdf}
988 mlamichh 4162 \label{fig:energyDrift}
989 gezelter 4174 \caption{Analysis of the energy conservation of the real-space
990 gezelter 4171 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
991 gezelter 4180 energy over time (in kcal / mol / particle / ns) and $\delta
992     \mathrm{E}_0$ is the standard deviation of energy fluctuations
993     around this drift (in kcal / mol / particle). All simulations were
994     of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
995     300 K starting from the same initial configuration. All runs
996     utilized the same real-space cutoff, $r_c = 12$\AA.}
997 gezelter 4171 \end{figure}
998    
999 gezelter 4174
1000 mlamichh 4114 \section{CONCLUSION}
1001 gezelter 4175 In the first paper in this series, we generalized the
1002     charge-neutralized electrostatic energy originally developed by Wolf
1003     \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1004     up to quadrupolar order. The SP method is essentially a
1005     multipole-capable version of the Wolf model. The SP method for
1006     multipoles provides excellent agreement with Ewald-derived energies,
1007     forces and torques, and is suitable for Monte Carlo simulations,
1008     although the forces and torques retain discontinuities at the cutoff
1009     distance that prevents its use in molecular dynamics.
1010 gezelter 4170
1011 gezelter 4175 We also developed two natural extensions of the damped shifted-force
1012     (DSF) model originally proposed by Fennel and
1013     Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1014     smooth truncation of energies, forces, and torques at the real-space
1015     cutoff, and both converge to DSF electrostatics for point-charge
1016     interactions. The TSF model is based on a high-order truncated Taylor
1017     expansion which can be relatively perturbative inside the cutoff
1018     sphere. The GSF model takes the gradient from an images of the
1019     interacting multipole that has been projected onto the cutoff sphere
1020     to derive shifted force and torque expressions, and is a significantly
1021     more gentle approach.
1022 gezelter 4170
1023 gezelter 4175 Of the two newly-developed shifted force models, the GSF method
1024     produced quantitative agreement with Ewald energy, force, and torques.
1025     It also performs well in conserving energy in MD simulations. The
1026     Taylor-shifted (TSF) model provides smooth dynamics, but these take
1027     place on a potential energy surface that is significantly perturbed
1028     from Ewald-based electrostatics.
1029    
1030     % The direct truncation of any electrostatic potential energy without
1031     % multipole neutralization creates large fluctuations in molecular
1032     % simulations. This fluctuation in the energy is very large for the case
1033     % of crystal because of long range of multipole ordering (Refer paper
1034     % I).\cite{PaperI} This is also significant in the case of the liquid
1035     % because of the local multipole ordering in the molecules. If the net
1036     % multipole within cutoff radius neutralized within cutoff sphere by
1037     % placing image multiples on the surface of the sphere, this fluctuation
1038     % in the energy reduced significantly. Also, the multipole
1039     % neutralization in the generalized SP method showed very good agreement
1040     % with the Ewald as compared to direct truncation for the evaluation of
1041     % the $\triangle E$ between the configurations. In MD simulations, the
1042     % energy conservation is very important. The conservation of the total
1043     % energy can be ensured by i) enforcing the smooth truncation of the
1044     % energy, force and torque in the cutoff radius and ii) making the
1045     % energy, force and torque consistent with each other. The GSF and TSF
1046     % methods ensure the consistency and smooth truncation of the energy,
1047     % force and torque at the cutoff radius, as a result show very good
1048     % total energy conservation. But the TSF method does not show good
1049     % agreement in the absolute value of the electrostatic energy, force and
1050     % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1051     % energy and torque accurately and also conserved energy.
1052    
1053     The only cases we have found where the new GSF and SP real-space
1054     methods can be problematic are those which retain a bulk dipole moment
1055     at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1056     materials, uniform weighting of the orientational contributions can be
1057     important for converging the total energy. In these cases, the
1058     damping function which causes the non-uniform weighting can be
1059     replaced by the bare electrostatic kernel, and the energies return to
1060     the expected converged values.
1061    
1062     Based on the results of this work, the GSF method is a suitable and
1063     efficient replacement for the Ewald sum for evaluating electrostatic
1064     interactions in MD simulations. Both methods retain excellent
1065     fidelity to the Ewald energies, forces and torques. Additionally, the
1066     energy drift and fluctuations from the GSF electrostatics are better
1067     than a multipolar Ewald sum for finite-sized reciprocal spaces.
1068     Because they use real-space cutoffs with moderate cutoff radii, the
1069     GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1070     increases. Additionally, they can be made extremely efficient using
1071     spline interpolations of the radial functions. They require no
1072     Fourier transforms or $k$-space sums, and guarantee the smooth
1073     handling of energies, forces, and torques as multipoles cross the
1074     real-space cutoff boundary.
1075    
1076 gezelter 4180 \begin{acknowledgments}
1077     JDG acknowledges helpful discussions with Christopher
1078     Fennell. Support for this project was provided by the National
1079     Science Foundation under grant CHE-1362211. Computational time was
1080     provided by the Center for Research Computing (CRC) at the
1081     University of Notre Dame.
1082     \end{acknowledgments}
1083    
1084 gezelter 4167 %\bibliographystyle{aip}
1085 gezelter 4168 \newpage
1086 mlamichh 4114 \bibliography{references}
1087     \end{document}
1088    
1089     %
1090     % ****** End of file aipsamp.tex ******