ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4203
Committed: Wed Aug 6 19:10:04 2014 UTC (9 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 67688 byte(s)
Log Message:
Edits

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 4198 \title{Real space electrostatics for multipoles. II. Comparisons with
51     the Ewald Sum}
52 mlamichh 4114
53     \author{Madan Lamichhane}
54 gezelter 4186 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
55 mlamichh 4114
56     \author{Kathie E. Newman}
57 gezelter 4186 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
58 mlamichh 4114
59     \author{J. Daniel Gezelter}%
60     \email{gezelter@nd.edu.}
61 gezelter 4186 \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
62     }
63 mlamichh 4114
64 gezelter 4186 \date{\today}
65 mlamichh 4114
66     \begin{abstract}
67 gezelter 4187 We report on tests of the shifted potential (SP), gradient shifted
68     force (GSF), and Taylor shifted force (TSF) real-space methods for
69     multipole interactions developed in the first paper in this series,
70     using the multipolar Ewald sum as a reference method. The tests were
71     carried out in a variety of condensed-phase environments designed to
72     test up to quadrupole-quadrupole interactions. Comparisons of the
73     energy differences between configurations, molecular forces, and
74     torques were used to analyze how well the real-space models perform
75     relative to the more computationally expensive Ewald treatment. We
76     have also investigated the energy conservation properties of the new
77     methods in molecular dynamics simulations. The SP method shows
78     excellent agreement with configurational energy differences, forces,
79     and torques, and would be suitable for use in Monte Carlo
80     calculations. Of the two new shifted-force methods, the GSF
81     approach shows the best agreement with Ewald-derived energies,
82     forces, and torques and also exhibits energy conservation properties
83     that make it an excellent choice for efficient computation of
84     electrostatic interactions in molecular dynamics simulations.
85 mlamichh 4114 \end{abstract}
86    
87 gezelter 4175 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
88 mlamichh 4114 % Classification Scheme.
89 gezelter 4184 %\keywords{Electrostatics, Multipoles, Real-space}
90 gezelter 4167
91 mlamichh 4114 \maketitle
92    
93 mlamichh 4166 \section{\label{sec:intro}Introduction}
94 gezelter 4167 Computing the interactions between electrostatic sites is one of the
95 gezelter 4185 most expensive aspects of molecular simulations. There have been
96     significant efforts to develop practical, efficient and convergent
97     methods for handling these interactions. Ewald's method is perhaps the
98     best known and most accurate method for evaluating energies, forces,
99     and torques in explicitly-periodic simulation cells. In this approach,
100     the conditionally convergent electrostatic energy is converted into
101     two absolutely convergent contributions, one which is carried out in
102     real space with a cutoff radius, and one in reciprocal
103 gezelter 4187 space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
104 mlamichh 4114
105 gezelter 4167 When carried out as originally formulated, the reciprocal-space
106     portion of the Ewald sum exhibits relatively poor computational
107 gezelter 4187 scaling, making it prohibitive for large systems. By utilizing a
108     particle mesh and three dimensional fast Fourier transforms (FFT), the
109     particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
110 gezelter 4186 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
111     methods can decrease the computational cost from $O(N^2)$ down to $O(N
112     \log
113 gezelter 4187 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
114 gezelter 4167
115 gezelter 4185 Because of the artificial periodicity required for the Ewald sum,
116 gezelter 4167 interfacial molecular systems such as membranes and liquid-vapor
117 gezelter 4187 interfaces require modifications to the method. Parry's extension of
118     the three dimensional Ewald sum is appropriate for slab
119     geometries.\cite{Parry:1975if} Modified Ewald methods that were
120     developed to handle two-dimensional (2-D) electrostatic
121 gezelter 4191 interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
122     These methods were originally quite computationally
123 gezelter 4187 expensive.\cite{Spohr97,Yeh99} There have been several successful
124 gezelter 4191 efforts that reduced the computational cost of 2-D lattice summations,
125 gezelter 4187 bringing them more in line with the scaling for the full 3-D
126 gezelter 4191 treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
127     inherent periodicity required by the Ewald method can also be
128     problematic in a number of protein/solvent and ionic solution
129     environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
130 gezelter 4167
131 mlamichh 4166 \subsection{Real-space methods}
132 gezelter 4168 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
133     method for calculating electrostatic interactions between point
134 gezelter 4185 charges. They argued that the effective Coulomb interaction in most
135     condensed phase systems is effectively short
136     ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
137     computing the Madelung constant of an ionic solid), the material can
138     be considered as a set of ions interacting with neutral dipolar or
139     quadrupolar ``molecules'' giving an effective distance dependence for
140     the electrostatic interactions of $r^{-5}$ (see figure
141     \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
142     cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
143     electrostatic energy per ion converges more rapidly to the Madelung
144     energy than the dipolar approximation.\cite{Wolf92} To find the
145     correct Madelung constant, Lacman suggested that the NaCl structure
146     could be constructed in a way that the finite crystal terminates with
147     complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
148     what is effectively a set of octupoles at large distances. These facts
149     suggest that the Madelung constants are relatively short ranged for
150     perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
151 gezelter 4186 application of Wolf's method can provide accurate estimates of
152 gezelter 4185 Madelung constants using relatively short cutoff radii.
153    
154     Direct truncation of interactions at a cutoff radius creates numerical
155 gezelter 4186 errors. Wolf \textit{et al.} suggest that truncation errors are due
156 gezelter 4185 to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
157     neutralize this charge they proposed placing an image charge on the
158     surface of the cutoff sphere for every real charge inside the cutoff.
159     These charges are present for the evaluation of both the pair
160     interaction energy and the force, although the force expression
161 gezelter 4186 maintains a discontinuity at the cutoff sphere. In the original Wolf
162 gezelter 4185 formulation, the total energy for the charge and image were not equal
163 gezelter 4186 to the integral of the force expression, and as a result, the total
164 gezelter 4185 energy would not be conserved in molecular dynamics (MD)
165     simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
166     Gezelter later proposed shifted force variants of the Wolf method with
167     commensurate force and energy expressions that do not exhibit this
168 gezelter 4186 problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
169     were also proposed by Chen \textit{et
170 gezelter 4185 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
171 gezelter 4186 and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
172     used additional neutralization of higher order moments for systems of
173     point charges.\cite{Fukuda:2013sf}
174 mlamichh 4114
175 gezelter 4181 \begin{figure}
176 gezelter 4167 \centering
177 gezelter 4190 \includegraphics[width=\linewidth]{schematic.eps}
178 gezelter 4181 \caption{Top: Ionic systems exhibit local clustering of dissimilar
179     charges (in the smaller grey circle), so interactions are
180 gezelter 4184 effectively charge-multipole at longer distances. With hard
181     cutoffs, motion of individual charges in and out of the cutoff
182     sphere can break the effective multipolar ordering. Bottom:
183     dipolar crystals and fluids have a similar effective
184 gezelter 4181 \textit{quadrupolar} ordering (in the smaller grey circles), and
185     orientational averaging helps to reduce the effective range of the
186     interactions in the fluid. Placement of reversed image multipoles
187     on the surface of the cutoff sphere recovers the effective
188     higher-order multipole behavior.}
189     \label{fig:schematic}
190 gezelter 4167 \end{figure}
191 mlamichh 4114
192 gezelter 4185 One can make a similar effective range argument for crystals of point
193     \textit{multipoles}. The Luttinger and Tisza treatment of energy
194     constants for dipolar lattices utilizes 24 basis vectors that contain
195 gezelter 4186 dipoles at the eight corners of a unit cube.\cite{LT} Only three of
196     these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
197 gezelter 4185 moments, while the rest have zero net dipole and retain contributions
198 gezelter 4186 only from higher order multipoles. The lowest-energy crystalline
199 gezelter 4185 structures are built out of basis vectors that have only residual
200     quadrupolar moments (e.g. the $Z_5$ array). In these low energy
201     structures, the effective interaction between a dipole at the center
202 gezelter 4167 of a crystal and a group of eight dipoles farther away is
203     significantly shorter ranged than the $r^{-3}$ that one would expect
204     for raw dipole-dipole interactions. Only in crystals which retain a
205     bulk dipole moment (e.g. ferroelectrics) does the analogy with the
206     ionic crystal break down -- ferroelectric dipolar crystals can exist,
207     while ionic crystals with net charge in each unit cell would be
208     unstable.
209    
210     In ionic crystals, real-space truncation can break the effective
211 gezelter 4181 multipolar arrangements (see Fig. \ref{fig:schematic}), causing
212     significant swings in the electrostatic energy as individual ions move
213     back and forth across the boundary. This is why the image charges are
214 gezelter 4180 necessary for the Wolf sum to exhibit rapid convergence. Similarly,
215     the real-space truncation of point multipole interactions breaks
216     higher order multipole arrangements, and image multipoles are required
217     for real-space treatments of electrostatic energies.
218 gezelter 4167
219 gezelter 4181 The shorter effective range of electrostatic interactions is not
220     limited to perfect crystals, but can also apply in disordered fluids.
221 gezelter 4186 Even at elevated temperatures, there is local charge balance in an
222     ionic liquid, where each positive ion has surroundings dominated by
223     negaitve ions and vice versa. The reversed-charge images on the
224     cutoff sphere that are integral to the Wolf and DSF approaches retain
225     the effective multipolar interactions as the charges traverse the
226     cutoff boundary.
227 gezelter 4181
228     In multipolar fluids (see Fig. \ref{fig:schematic}) there is
229     significant orientational averaging that additionally reduces the
230     effect of long-range multipolar interactions. The image multipoles
231     that are introduced in the TSF, GSF, and SP methods mimic this effect
232     and reduce the effective range of the multipolar interactions as
233     interacting molecules traverse each other's cutoff boundaries.
234    
235 gezelter 4167 % Because of this reason, although the nature of electrostatic
236     % interaction short ranged, the hard cutoff sphere creates very large
237     % fluctuation in the electrostatic energy for the perfect crystal. In
238     % addition, the charge neutralized potential proposed by Wolf et
239     % al. converged to correct Madelung constant but still holds oscillation
240     % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
241     % oscillation in the energy around its fully converged value can be due
242     % to the non-neutralized value of the higher order moments within the
243     % cutoff sphere.
244    
245 gezelter 4186 Forces and torques acting on atomic sites are fundamental in driving
246     dynamics in molecular simulations, and the damped shifted force (DSF)
247     energy kernel provides consistent energies and forces on charged atoms
248     within the cutoff sphere. Both the energy and the force go smoothly to
249     zero as an atom aproaches the cutoff radius. The comparisons of the
250     accuracy these quantities between the DSF kernel and SPME was
251     surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
252     has seen increasing use in molecular systems with relatively uniform
253     charge
254     densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
255 gezelter 4167
256 gezelter 4168 \subsection{The damping function}
257 gezelter 4185 The damping function has been discussed in detail in the first paper
258     of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
259     interactions between point charges can be replaced by the
260     complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
261     convergence, where $\alpha$ is a damping parameter with units of
262     inverse distance. Altering the value of $\alpha$ is equivalent to
263     changing the width of Gaussian charge distributions that replace each
264     point charge, as Coulomb integrals with Gaussian charge distributions
265     produce complementary error functions when truncated at a finite
266     distance.
267 mlamichh 4114
268 gezelter 4185 With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
269     produced very good agreement with SPME for interaction energies,
270     forces and torques for charge-charge
271     interactions.\cite{Fennell:2006lq}
272 gezelter 4167
273 gezelter 4168 \subsection{Point multipoles in molecular modeling}
274     Coarse-graining approaches which treat entire molecular subsystems as
275     a single rigid body are now widely used. A common feature of many
276     coarse-graining approaches is simplification of the electrostatic
277     interactions between bodies so that fewer site-site interactions are
278 gezelter 4185 required to compute configurational
279     energies.\cite{Ren06,Essex10,Essex11}
280 mlamichh 4166
281 gezelter 4186 Additionally, because electrons in a molecule are not localized at
282     specific points, the assignment of partial charges to atomic centers
283     is always an approximation. For increased accuracy, atomic sites can
284     also be assigned point multipoles and polarizabilities. Recently,
285     water has been modeled with point multipoles up to octupolar order
286     using the soft sticky dipole-quadrupole-octupole (SSDQO)
287 gezelter 4180 model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
288 gezelter 4168 multipoles up to quadrupolar order have also been coupled with point
289     polarizabilities in the high-quality AMOEBA and iAMOEBA water
290 gezelter 4185 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
291     truncating point multipoles without smoothing the forces and torques
292 gezelter 4186 can create energy conservation issues in molecular dynamics
293     simulations.
294 mlamichh 4166
295 gezelter 4168 In this paper we test a set of real-space methods that were developed
296     for point multipolar interactions. These methods extend the damped
297     shifted force (DSF) and Wolf methods originally developed for
298     charge-charge interactions and generalize them for higher order
299 gezelter 4186 multipoles. The detailed mathematical development of these methods
300     has been presented in the first paper in this series, while this work
301     covers the testing of energies, forces, torques, and energy
302 gezelter 4168 conservation properties of the methods in realistic simulation
303     environments. In all cases, the methods are compared with the
304 gezelter 4186 reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
305 gezelter 4168
306    
307 mlamichh 4166 %\subsection{Conservation of total energy }
308 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.
309 mlamichh 4166
310 gezelter 4168 \section{\label{sec:method}Review of Methods}
311     Any real-space electrostatic method that is suitable for MD
312     simulations should have the electrostatic energy, forces and torques
313     between two sites go smoothly to zero as the distance between the
314     sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
315     this continuity at the cutoff is essential for energy conservation in
316     MD simulations. The mathematical details of the shifted potential
317     (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
318     methods have been discussed in detail in the previous paper in this
319     series.\cite{PaperI} Here we briefly review the new methods and
320     describe their essential features.
321 mlamichh 4166
322 gezelter 4168 \subsection{Taylor-shifted force (TSF)}
323 mlamichh 4114
324 gezelter 4168 The electrostatic potential energy between point multipoles can be
325     expressed as the product of two multipole operators and a Coulombic
326     kernel,
327 mlamichh 4114 \begin{equation}
328 gezelter 4168 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
329 mlamichh 4114 \end{equation}
330 gezelter 4180 where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
331     expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
332     a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
333 gezelter 4184 $\bf a$, etc.
334 mlamichh 4166
335 gezelter 4180 % Interactions between multipoles can be expressed as higher derivatives
336     % of the bare Coulomb potential, so one way of ensuring that the forces
337     % and torques vanish at the cutoff distance is to include a larger
338     % number of terms in the truncated Taylor expansion, e.g.,
339     % %
340     % \begin{equation}
341     % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
342     % \end{equation}
343     % %
344     % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
345     % Thus, for $f(r)=1/r$, we find
346     % %
347     % \begin{equation}
348     % 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} .
349     % \end{equation}
350     % This function is an approximate electrostatic potential that has
351     % vanishing second derivatives at the cutoff radius, making it suitable
352     % for shifting the forces and torques of charge-dipole interactions.
353 gezelter 4168
354 gezelter 4180 The TSF potential for any multipole-multipole interaction can be
355     written
356 gezelter 4168 \begin{equation}
357     U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
358     \label{generic}
359     \end{equation}
360 gezelter 4180 where $f_n(r)$ is a shifted kernel that is appropriate for the order
361 gezelter 4181 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
362     charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
363     and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
364     quadrupole-quadrupole. To ensure smooth convergence of the energy,
365     force, and torques, a Taylor expansion with $n$ terms must be
366     performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
367 gezelter 4168
368 gezelter 4180 % To carry out the same procedure for a damped electrostatic kernel, we
369     % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
370     % Many of the derivatives of the damped kernel are well known from
371     % Smith's early work on multipoles for the Ewald
372     % summation.\cite{Smith82,Smith98}
373 gezelter 4168
374 gezelter 4180 % Note that increasing the value of $n$ will add additional terms to the
375     % electrostatic potential, e.g., $f_2(r)$ includes orders up to
376     % $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
377     % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
378     % f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
379     % for computing multipole energies, forces, and torques, and smooth
380     % cutoffs of these quantities can be guaranteed as long as the number of
381     % terms in the Taylor series exceeds the derivative order required.
382 gezelter 4168
383     For multipole-multipole interactions, following this procedure results
384 gezelter 4180 in separate radial functions for each of the distinct orientational
385     contributions to the potential, and ensures that the forces and
386     torques from each of these contributions will vanish at the cutoff
387     radius. For example, the direct dipole dot product
388     ($\mathbf{D}_{\bf a}
389     \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
390 gezelter 4168 dot products:
391     \begin{equation}
392 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
393     \mathbf{D}_{\bf a} \cdot
394     \mathbf{D}_{\bf b} \right) v_{21}(r) +
395     \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
396     \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
397 gezelter 4168 \end{equation}
398    
399 gezelter 4180 For the Taylor shifted (TSF) method with the undamped kernel,
400     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
401     \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
402     - \frac{6}{r r_c^2}$. In these functions, one can easily see the
403     connection to unmodified electrostatics as well as the smooth
404     transition to zero in both these functions as $r\rightarrow r_c$. The
405     electrostatic forces and torques acting on the central multipole due
406 gezelter 4184 to another site within the cutoff sphere are derived from
407 gezelter 4168 Eq.~\ref{generic}, accounting for the appropriate number of
408     derivatives. Complete energy, force, and torque expressions are
409     presented in the first paper in this series (Reference
410 gezelter 4175 \onlinecite{PaperI}).
411 gezelter 4168
412     \subsection{Gradient-shifted force (GSF)}
413    
414 gezelter 4180 A second (and conceptually simpler) method involves shifting the
415     gradient of the raw Coulomb potential for each particular multipole
416 gezelter 4168 order. For example, the raw dipole-dipole potential energy may be
417     shifted smoothly by finding the gradient for two interacting dipoles
418     which have been projected onto the surface of the cutoff sphere
419     without changing their relative orientation,
420 gezelter 4181 \begin{equation}
421 gezelter 4180 U_{D_{\bf a}D_{\bf b}}(r) = U_{D_{\bf a}D_{\bf b}}(r) -
422     U_{D_{\bf a} D_{\bf b}}(r_c)
423     - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
424 gezelter 4184 \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
425 gezelter 4181 \end{equation}
426 gezelter 4180 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
427     a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
428     (although the signs are reversed for the dipole that has been
429     projected onto the cutoff sphere). In many ways, this simpler
430     approach is closer in spirit to the original shifted force method, in
431     that it projects a neutralizing multipole (and the resulting forces
432     from this multipole) onto a cutoff sphere. The resulting functional
433     forms for the potentials, forces, and torques turn out to be quite
434     similar in form to the Taylor-shifted approach, although the radial
435     contributions are significantly less perturbed by the gradient-shifted
436     approach than they are in the Taylor-shifted method.
437 gezelter 4168
438 gezelter 4180 For the gradient shifted (GSF) method with the undamped kernel,
439     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
440     $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
441     Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
442     because the Taylor expansion retains only one term, they are
443     significantly less perturbed than the TSF functions.
444    
445 gezelter 4168 In general, the gradient shifted potential between a central multipole
446     and any multipolar site inside the cutoff radius is given by,
447     \begin{equation}
448 gezelter 4184 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
449     U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
450     \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
451 gezelter 4168 \label{generic2}
452     \end{equation}
453     where the sum describes a separate force-shifting that is applied to
454 gezelter 4184 each orientational contribution to the energy. In this expression,
455     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
456     ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
457     represent the orientations the multipoles.
458 gezelter 4168
459     The third term converges more rapidly than the first two terms as a
460     function of radius, hence the contribution of the third term is very
461     small for large cutoff radii. The force and torque derived from
462 gezelter 4184 Eq. \ref{generic2} are consistent with the energy expression and
463 gezelter 4175 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
464 gezelter 4168 can be considered generalizations of the original DSF method for
465     higher order multipole interactions. GSF and TSF are also identical up
466     to the charge-dipole interaction but generate different expressions in
467     the energy, force and torque for higher order multipole-multipole
468     interactions. Complete energy, force, and torque expressions for the
469     GSF potential are presented in the first paper in this series
470 gezelter 4184 (Reference~\onlinecite{PaperI}).
471 gezelter 4168
472    
473 mlamichh 4166 \subsection{Shifted potential (SP) }
474 gezelter 4168 A discontinuous truncation of the electrostatic potential at the
475     cutoff sphere introduces a severe artifact (oscillation in the
476     electrostatic energy) even for molecules with the higher-order
477     multipoles.\cite{PaperI} We have also formulated an extension of the
478     Wolf approach for point multipoles by simply projecting the image
479     multipole onto the surface of the cutoff sphere, and including the
480     interactions with the central multipole and the image. This
481     effectively shifts the total potential to zero at the cutoff radius,
482 mlamichh 4166 \begin{equation}
483 gezelter 4180 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
484 gezelter 4184 U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
485 mlamichh 4166 \label{eq:SP}
486     \end{equation}
487 gezelter 4168 where the sum describes separate potential shifting that is done for
488     each orientational contribution to the energy (e.g. the direct dipole
489     product contribution is shifted {\it separately} from the
490     dipole-distance terms in dipole-dipole interactions). Note that this
491 gezelter 4175 is not a simple shifting of the total potential at $r_c$. Each radial
492 gezelter 4168 contribution is shifted separately. One consequence of this is that
493     multipoles that reorient after leaving the cutoff sphere can re-enter
494     the cutoff sphere without perturbing the total energy.
495 mlamichh 4166
496 gezelter 4180 For the shifted potential (SP) method with the undamped kernel,
497     $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
498     \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
499     central multipole and other multipolar sites goes smoothly to zero as
500     $r \rightarrow r_c$. However, the force and torque obtained from the
501     shifted potential (SP) are discontinuous at $r_c$. MD simulations
502     will still experience energy drift while operating under the SP
503     potential, but it may be suitable for Monte Carlo approaches where the
504     configurational energy differences are the primary quantity of
505     interest.
506 gezelter 4168
507 gezelter 4180 \subsection{The Self Term}
508 gezelter 4168 In the TSF, GSF, and SP methods, a self-interaction is retained for
509     the central multipole interacting with its own image on the surface of
510     the cutoff sphere. This self interaction is nearly identical with the
511     self-terms that arise in the Ewald sum for multipoles. Complete
512     expressions for the self terms are presented in the first paper in
513 gezelter 4175 this series (Reference \onlinecite{PaperI}).
514 mlamichh 4162
515 gezelter 4168
516 gezelter 4170 \section{\label{sec:methodology}Methodology}
517 mlamichh 4166
518 gezelter 4170 To understand how the real-space multipole methods behave in computer
519     simulations, it is vital to test against established methods for
520     computing electrostatic interactions in periodic systems, and to
521     evaluate the size and sources of any errors that arise from the
522     real-space cutoffs. In the first paper of this series, we compared
523     the dipolar and quadrupolar energy expressions against analytic
524     expressions for ordered dipolar and quadrupolar
525 gezelter 4174 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
526     used the multipolar Ewald sum as a reference method for comparing
527     energies, forces, and torques for molecular models that mimic
528 gezelter 4175 disordered and ordered condensed-phase systems. The parameters used
529 gezelter 4180 in the test cases are given in table~\ref{tab:pars}.
530 gezelter 4174
531 gezelter 4175 \begin{table}
532     \label{tab:pars}
533     \caption{The parameters used in the systems used to evaluate the new
534     real-space methods. The most comprehensive test was a liquid
535     composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
536     ions). This test excercises all orders of the multipolar
537     interactions developed in the first paper.}
538     \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
539     & \multicolumn{2}{c|}{LJ parameters} &
540     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
541     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
542     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
543     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
544     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
545     \AA\textsuperscript{2})} \\ \hline
546     Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
547 gezelter 4180 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
548 gezelter 4175 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
549 gezelter 4180 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
550 gezelter 4175 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
551     \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
552     \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
553     \end{tabularx}
554     \end{table}
555     The systems consist of pure multipolar solids (both dipole and
556     quadrupole), pure multipolar liquids (both dipole and quadrupole), a
557     fluid composed of sites containing both dipoles and quadrupoles
558     simultaneously, and a final test case that includes ions with point
559     charges in addition to the multipolar fluid. The solid-phase
560     parameters were chosen so that the systems can explore some
561     orientational freedom for the multipolar sites, while maintaining
562     relatively strict translational order. The SSDQ model used here is
563     not a particularly accurate water model, but it does test
564     dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
565     interactions at roughly the same magnitudes. The last test case, SSDQ
566     water with dissolved ions, exercises \textit{all} levels of the
567     multipole-multipole interactions we have derived so far and represents
568     the most complete test of the new methods.
569 mlamichh 4166
570 gezelter 4175 In the following section, we present results for the total
571     electrostatic energy, as well as the electrostatic contributions to
572     the force and torque on each molecule. These quantities have been
573     computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
574 gezelter 4180 and have been compared with the values obtained from the multipolar
575     Ewald sum. In Monte Carlo (MC) simulations, the energy differences
576 gezelter 4175 between two configurations is the primary quantity that governs how
577 gezelter 4191 the simulation proceeds. These differences are the most important
578 gezelter 4175 indicators of the reliability of a method even if the absolute
579     energies are not exact. For each of the multipolar systems listed
580     above, we have compared the change in electrostatic potential energy
581     ($\Delta E$) between 250 statistically-independent configurations. In
582     molecular dynamics (MD) simulations, the forces and torques govern the
583     behavior of the simulation, so we also compute the electrostatic
584     contributions to the forces and torques.
585    
586     \subsection{Implementation}
587     The real-space methods developed in the first paper in this series
588     have been implemented in our group's open source molecular simulation
589 gezelter 4187 program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
590 gezelter 4175 this work. The complementary error function can be a relatively slow
591     function on some processors, so all of the radial functions are
592     precomputed on a fine grid and are spline-interpolated to provide
593     values when required.
594    
595     Using the same simulation code, we compare to a multipolar Ewald sum
596     with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
597     the Ewald sum is a re-implementation of the algorithm originally
598     proposed by Smith that does not use the particle mesh or smoothing
599     approximations.\cite{Smith82,Smith98} In all cases, the quantities
600     being compared are the electrostatic contributions to energies, force,
601     and torques. All other contributions to these quantities (i.e. from
602     Lennard-Jones interactions) are removed prior to the comparisons.
603    
604     The convergence parameter ($\alpha$) also plays a role in the balance
605     of the real-space and reciprocal-space portions of the Ewald
606     calculation. Typical molecular mechanics packages set this to a value
607     that depends on the cutoff radius and a tolerance (typically less than
608     $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
609     associated with increasing accuracy at the expense of computational
610     time spent on the reciprocal-space portion of the
611     summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
612     10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
613     Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
614    
615     The real-space models have self-interactions that provide
616     contributions to the energies only. Although the self interaction is
617     a rapid calculation, we note that in systems with fluctuating charges
618     or point polarizabilities, the self-term is not static and must be
619     recomputed at each time step.
620    
621 gezelter 4170 \subsection{Model systems}
622 gezelter 4180 To sample independent configurations of the multipolar crystals, body
623     centered cubic (bcc) crystals, which exhibit the minimum energy
624     structures for point dipoles, were generated using 3,456 molecules.
625     The multipoles were translationally locked in their respective crystal
626     sites for equilibration at a relatively low temperature (50K) so that
627     dipoles or quadrupoles could freely explore all accessible
628     orientations. The translational constraints were then removed, the
629     systems were re-equilibrated, and the crystals were simulated for an
630     additional 10 ps in the microcanonical (NVE) ensemble with an average
631     temperature of 50 K. The balance between moments of inertia and
632     particle mass were chosen to allow orientational sampling without
633     significant translational motion. Configurations were sampled at
634     equal time intervals in order to compare configurational energy
635     differences. The crystals were simulated far from the melting point
636     in order to avoid translational deformation away of the ideal lattice
637     geometry.
638 gezelter 4170
639 gezelter 4180 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
640     simulations, each system was created with 2,048 randomly-oriented
641     molecules. These were equilibrated at a temperature of 300K for 1 ns.
642     Each system was then simulated for 1 ns in the microcanonical (NVE)
643     ensemble. We collected 250 different configurations at equal time
644     intervals. For the liquid system that included ionic species, we
645     converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
646     \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
647     run under the same conditions for 1 ns. A total of 250 configurations
648     were collected. In the following comparisons of energies, forces, and
649     torques, the Lennard-Jones potentials were turned off and only the
650     purely electrostatic quantities were compared with the same values
651     obtained via the Ewald sum.
652 gezelter 4170
653     \subsection{Accuracy of Energy Differences, Forces and Torques}
654     The pairwise summation techniques (outlined above) were evaluated for
655     use in MC simulations by studying the energy differences between
656     different configurations. We took the Ewald-computed energy
657     difference between two conformations to be the correct behavior. An
658     ideal performance by one of the new methods would reproduce these
659     energy differences exactly. The configurational energies being used
660     here contain only contributions from electrostatic interactions.
661     Lennard-Jones interactions were omitted from the comparison as they
662     should be identical for all methods.
663    
664     Since none of the real-space methods provide exact energy differences,
665 gezelter 4180 we used least square regressions analysis for the six different
666 gezelter 4170 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
667     with the multipolar Ewald reference method. Unitary results for both
668     the correlation (slope) and correlation coefficient for these
669     regressions indicate perfect agreement between the real-space method
670     and the multipolar Ewald sum.
671    
672     Molecular systems were run long enough to explore independent
673     configurations and 250 configurations were recorded for comparison.
674     Each system provided 31,125 energy differences for a total of 186,750
675     data points. Similarly, the magnitudes of the forces and torques have
676 gezelter 4180 also been compared using least squares regression analysis. In the
677 gezelter 4170 forces and torques comparison, the magnitudes of the forces acting in
678     each molecule for each configuration were evaluated. For example, our
679     dipolar liquid simulation contains 2048 molecules and there are 250
680     different configurations for each system resulting in 3,072,000 data
681     points for comparison of forces and torques.
682    
683 mlamichh 4166 \subsection{Analysis of vector quantities}
684 gezelter 4170 Getting the magnitudes of the force and torque vectors correct is only
685     part of the issue for carrying out accurate molecular dynamics
686     simulations. Because the real space methods reweight the different
687     orientational contributions to the energies, it is also important to
688     understand how the methods impact the \textit{directionality} of the
689     force and torque vectors. Fisher developed a probablity density
690     function to analyse directional data sets,
691 mlamichh 4162 \begin{equation}
692 gezelter 4170 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
693 mlamichh 4162 \label{eq:pdf}
694     \end{equation}
695 gezelter 4170 where $\kappa$ measures directional dispersion of the data around the
696     mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
697     estimated as a reciprocal of the circular variance.\cite{Allen91} To
698     quantify the directional error, forces obtained from the Ewald sum
699     were taken as the mean (or correct) direction and the angle between
700     the forces obtained via the Ewald sum and the real-space methods were
701     evaluated,
702 mlamichh 4162 \begin{equation}
703 gezelter 4170 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
704     \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
705     \end{equation}
706     The total angular displacement of the vectors was calculated as,
707     \begin{equation}
708     R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
709 mlamichh 4162 \label{eq:displacement}
710     \end{equation}
711 gezelter 4170 where $N$ is number of force vectors. The circular variance is
712     defined as
713     \begin{equation}
714     \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
715     \end{equation}
716     The circular variance takes on values between from 0 to 1, with 0
717     indicating a perfect directional match between the Ewald force vectors
718     and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
719     correspond to higher values of $\kappa$, which indicates tighter
720     clustering of the real-space force vectors around the Ewald forces.
721 mlamichh 4162
722 gezelter 4170 A similar analysis was carried out for the electrostatic contribution
723     to the molecular torques as well as forces.
724    
725 mlamichh 4166 \subsection{Energy conservation}
726 gezelter 4170 To test conservation the energy for the methods, the mixed molecular
727     system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
728     ions was run for 1 ns in the microcanonical ensemble at an average
729     temperature of 300K. Each of the different electrostatic methods
730     (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
731     damping values. The molecular system was started with same initial
732     positions and velocities for all cutoff methods. The energy drift
733     ($\delta E_1$) and standard deviation of the energy about the slope
734     ($\delta E_0$) were evaluated from the total energy of the system as a
735     function of time. Although both measures are valuable at
736     investigating new methods for molecular dynamics, a useful interaction
737     model must allow for long simulation times with minimal energy drift.
738 mlamichh 4114
739 mlamichh 4166 \section{\label{sec:result}RESULTS}
740     \subsection{Configurational energy differences}
741     %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.
742     %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.
743     %\begin{figure}[h!]
744     % \centering
745     % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
746     % \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. }
747     % \label{fig:rcutConvergence_dipolarLiquid}
748     % \end{figure}
749     %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.
750     %\begin{figure}
751     % \centering
752     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
753     % \label{fig:barGraph1}
754     % \end{figure}
755     % \begin{figure}
756     % \centering
757     % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
758     % \caption{}
759 mlamichh 4162
760 gezelter 4167 % \label{fig:barGraph2}
761     % \end{figure}
762 gezelter 4174 %The correlation coefficient ($R^2$) and slope of the linear
763     %regression plots for the energy differences for all six different
764     %molecular systems is shown in figure 4a and 4b.The plot shows that
765     %the correlation coefficient improves for the SP cutoff method as
766     %compared to the undamped hard cutoff method in the case of SSDQC,
767     %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
768     %crystal and liquid, the correlation coefficient is almost unchanged
769     %and close to 1. The correlation coefficient is smallest (0.696276
770     %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
771     %charge-charge and charge-multipole interactions. Since the
772     %charge-charge and charge-multipole interaction is long ranged, there
773     %is huge deviation of correlation coefficient from 1. Similarly, the
774     %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
775     %compared to interactions in the other multipolar systems, thus the
776     %correlation coefficient very close to 1 even for hard cutoff
777     %method. The idea of placing image multipole on the surface of the
778     %cutoff sphere improves the correlation coefficient and makes it close
779     %to 1 for all types of multipolar systems. Similarly the slope is
780     %hugely deviated from the correct value for the lower order
781     %multipole-multipole interaction and slightly deviated for higher
782     %order multipole – multipole interaction. The SP method improves both
783     %correlation coefficient ($R^2$) and slope significantly in SSDQC and
784     %dipolar systems. The Slope is found to be deviated more in dipolar
785     %crystal as compared to liquid which is associated with the large
786     %fluctuation in the electrostatic energy in crystal. The GSF also
787     %produced better values of correlation coefficient and slope with the
788     %proper selection of the damping alpha (Interested reader can consult
789     %accompanying supporting material). The TSF method gives good value of
790     %correlation coefficient for the dipolar crystal, dipolar liquid,
791     %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
792     %regression slopes are significantly deviated.
793    
794 mlamichh 4114 \begin{figure}
795 gezelter 4174 \centering
796 gezelter 4191 \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
797 gezelter 4174 \caption{Statistical analysis of the quality of configurational
798     energy differences for the real-space electrostatic methods
799     compared with the reference Ewald sum. Results with a value equal
800     to 1 (dashed line) indicate $\Delta E$ values indistinguishable
801     from those obtained using the multipolar Ewald sum. Different
802     values of the cutoff radius are indicated with different symbols
803     (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
804     triangles).}
805     \label{fig:slopeCorr_energy}
806     \end{figure}
807    
808     The combined correlation coefficient and slope for all six systems is
809     shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
810 gezelter 4175 reproduce the Ewald configurational energy differences with remarkable
811     fidelity. Undamped hard cutoffs introduce a significant amount of
812     random scatter in the energy differences which is apparent in the
813     reduced value of the correlation coefficient for this method. This
814     can be easily understood as configurations which exhibit small
815     traversals of a few dipoles or quadrupoles out of the cutoff sphere
816     will see large energy jumps when hard cutoffs are used. The
817 gezelter 4174 orientations of the multipoles (particularly in the ordered crystals)
818 gezelter 4175 mean that these energy jumps can go in either direction, producing a
819     significant amount of random scatter, but no systematic error.
820 gezelter 4174
821     The TSF method produces energy differences that are highly correlated
822     with the Ewald results, but it also introduces a significant
823     systematic bias in the values of the energies, particularly for
824     smaller cutoff values. The TSF method alters the distance dependence
825     of different orientational contributions to the energy in a
826     non-uniform way, so the size of the cutoff sphere can have a large
827 gezelter 4175 effect, particularly for the crystalline systems.
828 gezelter 4174
829     Both the SP and GSF methods appear to reproduce the Ewald results with
830     excellent fidelity, particularly for moderate damping ($\alpha =
831 gezelter 4175 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
832     12$\AA). With the exception of the undamped hard cutoff, and the TSF
833     method with short cutoffs, all of the methods would be appropriate for
834     use in Monte Carlo simulations.
835 gezelter 4174
836 mlamichh 4114 \subsection{Magnitude of the force and torque vectors}
837 gezelter 4174
838 gezelter 4175 The comparisons of the magnitudes of the forces and torques for the
839     data accumulated from all six systems are shown in Figures
840 gezelter 4174 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
841     correlation and slope for the forces agree well with the Ewald sum
842 gezelter 4175 even for the hard cutoffs.
843 gezelter 4174
844 gezelter 4175 For systems of molecules with only multipolar interactions, the pair
845     energy contributions are quite short ranged. Moreover, the force
846     decays more rapidly than the electrostatic energy, hence the hard
847     cutoff method can also produce reasonable agreement for this quantity.
848     Although the pure cutoff gives reasonably good electrostatic forces
849     for pairs of molecules included within each other's cutoff spheres,
850     the discontinuity in the force at the cutoff radius can potentially
851     cause energy conservation problems as molecules enter and leave the
852     cutoff spheres. This is discussed in detail in section
853     \ref{sec:conservation}.
854 gezelter 4174
855     The two shifted-force methods (GSF and TSF) exhibit a small amount of
856     systematic variation and scatter compared with the Ewald forces. The
857     shifted-force models intentionally perturb the forces between pairs of
858 gezelter 4175 molecules inside each other's cutoff spheres in order to correct the
859     energy conservation issues, and this perturbation is evident in the
860     statistics accumulated for the molecular forces. The GSF
861 gezelter 4180 perturbations are minimal, particularly for moderate damping and
862 gezelter 4174 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
863     reasonable agreement in the correlation coefficient but again the
864     systematic error in the forces is concerning if replication of Ewald
865     forces is desired.
866    
867 mlamichh 4114 \begin{figure}
868 gezelter 4174 \centering
869 gezelter 4191 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
870 gezelter 4174 \caption{Statistical analysis of the quality of the force vector
871     magnitudes for the real-space electrostatic methods compared with
872     the reference Ewald sum. Results with a value equal to 1 (dashed
873     line) indicate force magnitude values indistinguishable from those
874     obtained using the multipolar Ewald sum. Different values of the
875     cutoff radius are indicated with different symbols (9\AA\ =
876     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
877     \label{fig:slopeCorr_force}
878     \end{figure}
879    
880    
881 mlamichh 4114 \begin{figure}
882 gezelter 4174 \centering
883 gezelter 4191 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
884 gezelter 4174 \caption{Statistical analysis of the quality of the torque vector
885     magnitudes for the real-space electrostatic methods compared with
886     the reference Ewald sum. Results with a value equal to 1 (dashed
887     line) indicate force magnitude values indistinguishable from those
888     obtained using the multipolar Ewald sum. Different values of the
889     cutoff radius are indicated with different symbols (9\AA\ =
890     circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
891     \label{fig:slopeCorr_torque}
892     \end{figure}
893    
894     The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
895     significantly influenced by the choice of real-space method. The
896     torque expressions have the same distance dependence as the energies,
897     which are naturally longer-ranged expressions than the inter-site
898     forces. Torques are also quite sensitive to orientations of
899     neighboring molecules, even those that are near the cutoff distance.
900    
901     The results shows that the torque from the hard cutoff method
902     reproduces the torques in quite good agreement with the Ewald sum.
903 gezelter 4175 The other real-space methods can cause some deviations, but excellent
904     agreement with the Ewald sum torques is recovered at moderate values
905     of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
906     radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
907     in the slope when compared with the Ewald torques even for larger
908     cutoff radii. It appears that the severity of the perturbations in
909     the TSF method are most in evidence for the torques.
910 gezelter 4174
911 mlamichh 4114 \subsection{Directionality of the force and torque vectors}
912 mlamichh 4162
913 gezelter 4174 The accurate evaluation of force and torque directions is just as
914     important for molecular dynamics simulations as the magnitudes of
915     these quantities. Force and torque vectors for all six systems were
916     analyzed using Fisher statistics, and the quality of the vector
917     directionality is shown in terms of circular variance
918 gezelter 4180 ($\mathrm{Var}(\theta)$) in figure
919 gezelter 4174 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
920 gezelter 4175 from the new real-space methods exhibit nearly-ideal Fisher probability
921 gezelter 4174 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
922     exhibit the best vectorial agreement with the Ewald sum. The force and
923     torque vectors from GSF method also show good agreement with the Ewald
924     method, which can also be systematically improved by using moderate
925 gezelter 4175 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
926 gezelter 4174 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
927 gezelter 4175 to a distribution with 95\% of force vectors within $6.37^\circ$ of
928     the corresponding Ewald forces. The TSF method produces the poorest
929 gezelter 4174 agreement with the Ewald force directions.
930    
931 gezelter 4175 Torques are again more perturbed than the forces by the new real-space
932     methods, but even here the variance is reasonably small. For the same
933 gezelter 4174 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
934     the circular variance was 0.01415, corresponds to a distribution which
935     has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
936     results. Again, the direction of the force and torque vectors can be
937     systematically improved by varying $\alpha$ and $r_c$.
938    
939 mlamichh 4114 \begin{figure}
940 gezelter 4174 \centering
941 gezelter 4191 \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
942 gezelter 4174 \caption{The circular variance of the direction of the force and
943     torque vectors obtained from the real-space methods around the
944     reference Ewald vectors. A variance equal to 0 (dashed line)
945     indicates direction of the force or torque vectors are
946     indistinguishable from those obtained from the Ewald sum. Here
947     different symbols represent different values of the cutoff radius
948     (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
949     \label{fig:slopeCorr_circularVariance}
950     \end{figure}
951 gezelter 4171
952 gezelter 4175 \subsection{Energy conservation\label{sec:conservation}}
953 gezelter 4171
954 gezelter 4174 We have tested the conservation of energy one can expect to see with
955     the new real-space methods using the SSDQ water model with a small
956     fraction of solvated ions. This is a test system which exercises all
957     orders of multipole-multipole interactions derived in the first paper
958     in this series and provides the most comprehensive test of the new
959     methods. A liquid-phase system was created with 2000 water molecules
960 gezelter 4175 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
961 gezelter 4174 temperature of 300K. After equilibration, this liquid-phase system
962     was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
963 gezelter 4175 a cutoff radius of 12\AA. The value of the damping coefficient was
964 gezelter 4174 also varied from the undamped case ($\alpha = 0$) to a heavily damped
965 gezelter 4175 case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods. A
966     sample was also run using the multipolar Ewald sum with the same
967     real-space cutoff.
968 gezelter 4174
969     In figure~\ref{fig:energyDrift} we show the both the linear drift in
970     energy over time, $\delta E_1$, and the standard deviation of energy
971     fluctuations around this drift $\delta E_0$. Both of the
972     shifted-force methods (GSF and TSF) provide excellent energy
973 gezelter 4181 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
974 gezelter 4174 while the hard cutoff is essentially unusable for molecular dynamics.
975     SP provides some benefit over the hard cutoff because the energetic
976     jumps that happen as particles leave and enter the cutoff sphere are
977 gezelter 4175 somewhat reduced, but like the Wolf method for charges, the SP method
978     would not be as useful for molecular dynamics as either of the
979     shifted-force methods.
980 gezelter 4174
981     We note that for all tested values of the cutoff radius, the new
982     real-space methods can provide better energy conservation behavior
983     than the multipolar Ewald sum, even when utilizing a relatively large
984     $k$-space cutoff values.
985    
986 mlamichh 4114 \begin{figure}
987 gezelter 4171 \centering
988 gezelter 4190 \includegraphics[width=\textwidth]{newDrift_12.eps}
989 mlamichh 4162 \label{fig:energyDrift}
990 gezelter 4174 \caption{Analysis of the energy conservation of the real-space
991 gezelter 4171 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
992 gezelter 4180 energy over time (in kcal / mol / particle / ns) and $\delta
993     \mathrm{E}_0$ is the standard deviation of energy fluctuations
994     around this drift (in kcal / mol / particle). All simulations were
995     of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
996     300 K starting from the same initial configuration. All runs
997     utilized the same real-space cutoff, $r_c = 12$\AA.}
998 gezelter 4171 \end{figure}
999    
1000 gezelter 4203 \subsection{Reproduction of Structural Features\label{sec:structure}}
1001     One of the best tests of modified interaction potentials is the
1002     fidelity with which they can reproduce structural features in a
1003     liquid. One commonly-utilized measure of structural ordering is the
1004     pair distribution function, $g(r)$, which measures local density
1005     deviations in relation to the bulk density. In the electrostatic
1006     approaches studied here, the short-range repulsion from the
1007     Lennard-Jones potential is identical for the various electrostatic
1008     methods, and since short range repulsion determines much of the local
1009     liquid ordering, one would not expect to see any differences in
1010     $g(r)$. Indeed, the pair distributions are essentially identical for
1011     all of the electrostatic methods studied (for each of the different
1012     systems under investigation). Interested readers may consult the
1013     supplementary information for plots of these pair distribution
1014     functions.
1015 gezelter 4174
1016 gezelter 4203 A direct measure of the structural features that is a more
1017     enlightening test of the modified electrostatic methods is the average
1018     value of the electrostatic energy $\langle U_\mathrm{elect} \rangle$
1019     which is obtained by sampling the liquid-state configurations
1020     experienced by a liquid evolving entirely under the influence of the
1021     methods being investigated. In figure \ref{fig:Uelect} we show how
1022     $\langle U_\mathrm{elect} \rangle$ for varies with the damping parameter,
1023     $\alpha$, for each of the methods.
1024    
1025     \begin{figure}
1026     \centering
1027     \includegraphics[width=\textwidth]{averagePotentialEnergy_r9_12.eps}
1028     \label{fig:Uelect}
1029     \caption{The average electrostatic potential energy,
1030     $\langle U_\mathrm{elect} \rangle$ for the SSDQ water with ions as a function
1031     of the damping parameter, $\alpha$, for each of the real-space
1032     electrostatic methods. Top panel: simulations run with a real-space
1033     cutoff, $r_c = 9$\AA. Bottom panel: the same quantity, but with a
1034     larger cutoff, $r_c = 12$\AA.}
1035     \end{figure}
1036    
1037     It is clear that moderate damping is important for converging the mean
1038     potential energy values, particularly for the two shifted force
1039     methods (GSF and TSF). A value of $\alpha \approx 0.18$ \AA$^{-1}$ is
1040     sufficient to converge the SP and GSF energies with a cutoff of 12
1041     \AA, while shorter cutoffs require more dramatic damping ($\alpha
1042     \approx 0.36$ \AA$^{-1}$ for $r_c = 9$ \AA). It is also clear from
1043     fig. \ref{fig:Uelect} that it is possible to overdamp the real-space
1044     electrostatic methods, causing the estimate of the energy to drop
1045     below the Ewald results.
1046    
1047     These ``optimal'' values of the damping coefficient are slightly
1048     larger than what were observed for DSF electrostatics for purely
1049     point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1050     $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1051     multipole systems.
1052    
1053     \subsection{Reproduction of Dynamic Properties\label{sec:structure}}
1054     To test the fidelity of the electrostatic methods at reproducing
1055     dynamics in a multipolar liquid, it is also useful to look at
1056     transport properties, particularly the diffusion constant,
1057     \begin{equation}
1058     D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1059     \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1060     \label{eq:diff}
1061     \end{equation}
1062     which measures long-time behavior and is sensitive to the forces on
1063     the multipoles. For the soft dipolar fluid, and the SSDQ liquid
1064     systems, the self-diffusion constants (D) were calculated from linear
1065     fits to the long-time portion of the mean square displacement
1066     ($\langle r^{2}(t) \rangle$).\cite{Allen87}
1067    
1068     In addition to translational diffusion, orientational relaxation times
1069     were calculated for comparisons with the Ewald simulations and with
1070     experiments. These values were determined from the same 1~ns $NVE$
1071     trajectories used for translational diffusion by calculating the
1072     orientational time correlation function,
1073     \begin{equation}
1074     C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\gamma(t)
1075     \cdot\hat{\mathbf{u}}_i^\gamma(0)\right]\right\rangle,
1076     \label{eq:OrientCorr}
1077     \end{equation}
1078     where $P_l$ is the Legendre polynomial of order $l$ and
1079     $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
1080     axis $\gamma$. The body-fixed reference frame used for our
1081     orientational correlation functions has the $z$-axis running along the
1082     dipoles, and for the SSDQ water model, the $y$-axis connects the two
1083     implied hydrogen atoms.
1084    
1085     From the orientation autocorrelation functions, we can obtain time
1086     constants for rotational relaxation either by fitting an exponential
1087     function or by integrating the entire correlation function. These
1088     decay times are directly comparable to water orientational relaxation
1089     times from nuclear magnetic resonance (NMR). The relaxation constant
1090     obtained from $C_2^y(t)$ is normally of experimental interest because
1091     it describes the relaxation of the principle axis connecting the
1092     hydrogen atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular
1093     portion of the dipole-dipole relaxation from a proton NMR signal and
1094     should provide an estimate of the NMR relaxation time
1095     constant.\cite{Impey82}
1096    
1097     Results for the diffusion constants and orientational relaxation times
1098     are shown in figure \ref{fig:dynamics}. From this data, it is apparent
1099     that the values for both $D$ and $\tau_2$ using the Ewald sum are
1100     reproduced with high fidelity by the GSF method.
1101    
1102     The $\tau_2$ results in \ref{fig:dynamics} show a much greater
1103     difference between the real-space and the Ewald results.
1104    
1105    
1106 mlamichh 4114 \section{CONCLUSION}
1107 gezelter 4175 In the first paper in this series, we generalized the
1108     charge-neutralized electrostatic energy originally developed by Wolf
1109     \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1110     up to quadrupolar order. The SP method is essentially a
1111     multipole-capable version of the Wolf model. The SP method for
1112     multipoles provides excellent agreement with Ewald-derived energies,
1113     forces and torques, and is suitable for Monte Carlo simulations,
1114     although the forces and torques retain discontinuities at the cutoff
1115     distance that prevents its use in molecular dynamics.
1116 gezelter 4170
1117 gezelter 4175 We also developed two natural extensions of the damped shifted-force
1118     (DSF) model originally proposed by Fennel and
1119     Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1120     smooth truncation of energies, forces, and torques at the real-space
1121     cutoff, and both converge to DSF electrostatics for point-charge
1122     interactions. The TSF model is based on a high-order truncated Taylor
1123     expansion which can be relatively perturbative inside the cutoff
1124     sphere. The GSF model takes the gradient from an images of the
1125     interacting multipole that has been projected onto the cutoff sphere
1126     to derive shifted force and torque expressions, and is a significantly
1127     more gentle approach.
1128 gezelter 4170
1129 gezelter 4175 Of the two newly-developed shifted force models, the GSF method
1130     produced quantitative agreement with Ewald energy, force, and torques.
1131     It also performs well in conserving energy in MD simulations. The
1132     Taylor-shifted (TSF) model provides smooth dynamics, but these take
1133     place on a potential energy surface that is significantly perturbed
1134     from Ewald-based electrostatics.
1135    
1136     % The direct truncation of any electrostatic potential energy without
1137     % multipole neutralization creates large fluctuations in molecular
1138     % simulations. This fluctuation in the energy is very large for the case
1139     % of crystal because of long range of multipole ordering (Refer paper
1140     % I).\cite{PaperI} This is also significant in the case of the liquid
1141     % because of the local multipole ordering in the molecules. If the net
1142     % multipole within cutoff radius neutralized within cutoff sphere by
1143     % placing image multiples on the surface of the sphere, this fluctuation
1144     % in the energy reduced significantly. Also, the multipole
1145     % neutralization in the generalized SP method showed very good agreement
1146     % with the Ewald as compared to direct truncation for the evaluation of
1147     % the $\triangle E$ between the configurations. In MD simulations, the
1148     % energy conservation is very important. The conservation of the total
1149     % energy can be ensured by i) enforcing the smooth truncation of the
1150     % energy, force and torque in the cutoff radius and ii) making the
1151     % energy, force and torque consistent with each other. The GSF and TSF
1152     % methods ensure the consistency and smooth truncation of the energy,
1153     % force and torque at the cutoff radius, as a result show very good
1154     % total energy conservation. But the TSF method does not show good
1155     % agreement in the absolute value of the electrostatic energy, force and
1156     % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1157     % energy and torque accurately and also conserved energy.
1158    
1159     The only cases we have found where the new GSF and SP real-space
1160     methods can be problematic are those which retain a bulk dipole moment
1161     at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1162     materials, uniform weighting of the orientational contributions can be
1163     important for converging the total energy. In these cases, the
1164     damping function which causes the non-uniform weighting can be
1165     replaced by the bare electrostatic kernel, and the energies return to
1166     the expected converged values.
1167    
1168     Based on the results of this work, the GSF method is a suitable and
1169     efficient replacement for the Ewald sum for evaluating electrostatic
1170     interactions in MD simulations. Both methods retain excellent
1171     fidelity to the Ewald energies, forces and torques. Additionally, the
1172     energy drift and fluctuations from the GSF electrostatics are better
1173     than a multipolar Ewald sum for finite-sized reciprocal spaces.
1174     Because they use real-space cutoffs with moderate cutoff radii, the
1175     GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1176     increases. Additionally, they can be made extremely efficient using
1177     spline interpolations of the radial functions. They require no
1178     Fourier transforms or $k$-space sums, and guarantee the smooth
1179     handling of energies, forces, and torques as multipoles cross the
1180     real-space cutoff boundary.
1181    
1182 gezelter 4180 \begin{acknowledgments}
1183     JDG acknowledges helpful discussions with Christopher
1184     Fennell. Support for this project was provided by the National
1185     Science Foundation under grant CHE-1362211. Computational time was
1186     provided by the Center for Research Computing (CRC) at the
1187     University of Notre Dame.
1188     \end{acknowledgments}
1189    
1190 gezelter 4167 %\bibliographystyle{aip}
1191 gezelter 4168 \newpage
1192 mlamichh 4114 \bibliography{references}
1193     \end{document}
1194    
1195     %
1196     % ****** End of file aipsamp.tex ******