ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4221
Committed: Sun Sep 14 13:40:01 2014 UTC (9 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 62089 byte(s)
Log Message:
FInal Version, I think.

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 gezelter 4208 have also investigated the energy conservation, structural, and
77     dynamical properties of the new methods in molecular dynamics
78     simulations. The SP method shows excellent agreement with
79     configurational energy differences, forces, and torques, and would
80     be suitable for use in Monte Carlo calculations. Of the two new
81     shifted-force methods, the GSF approach shows the best agreement
82     with Ewald-derived energies, forces, and torques and also exhibits
83     energy conservation properties that make it an excellent choice for
84     efficient computation of electrostatic interactions in molecular
85     dynamics simulations. Both SP and GSF are able to reproduce
86 gezelter 4212 structural and dynamical properties in the liquid models with
87 gezelter 4208 excellent fidelity.
88 mlamichh 4114 \end{abstract}
89    
90 gezelter 4175 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
91 mlamichh 4114 % Classification Scheme.
92 gezelter 4184 %\keywords{Electrostatics, Multipoles, Real-space}
93 gezelter 4167
94 mlamichh 4114 \maketitle
95    
96 mlamichh 4166 \section{\label{sec:intro}Introduction}
97 gezelter 4167 Computing the interactions between electrostatic sites is one of the
98 gezelter 4185 most expensive aspects of molecular simulations. There have been
99     significant efforts to develop practical, efficient and convergent
100     methods for handling these interactions. Ewald's method is perhaps the
101     best known and most accurate method for evaluating energies, forces,
102     and torques in explicitly-periodic simulation cells. In this approach,
103     the conditionally convergent electrostatic energy is converted into
104     two absolutely convergent contributions, one which is carried out in
105     real space with a cutoff radius, and one in reciprocal
106 gezelter 4187 space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
107 mlamichh 4114
108 gezelter 4167 When carried out as originally formulated, the reciprocal-space
109     portion of the Ewald sum exhibits relatively poor computational
110 gezelter 4187 scaling, making it prohibitive for large systems. By utilizing a
111     particle mesh and three dimensional fast Fourier transforms (FFT), the
112     particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
113 gezelter 4186 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
114     methods can decrease the computational cost from $O(N^2)$ down to $O(N
115     \log
116 gezelter 4187 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
117 gezelter 4167
118 gezelter 4185 Because of the artificial periodicity required for the Ewald sum,
119 gezelter 4167 interfacial molecular systems such as membranes and liquid-vapor
120 gezelter 4187 interfaces require modifications to the method. Parry's extension of
121     the three dimensional Ewald sum is appropriate for slab
122     geometries.\cite{Parry:1975if} Modified Ewald methods that were
123     developed to handle two-dimensional (2-D) electrostatic
124 gezelter 4191 interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
125     These methods were originally quite computationally
126 gezelter 4187 expensive.\cite{Spohr97,Yeh99} There have been several successful
127 gezelter 4191 efforts that reduced the computational cost of 2-D lattice summations,
128 gezelter 4187 bringing them more in line with the scaling for the full 3-D
129 gezelter 4191 treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
130     inherent periodicity required by the Ewald method can also be
131     problematic in a number of protein/solvent and ionic solution
132     environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
133 gezelter 4167
134 mlamichh 4166 \subsection{Real-space methods}
135 gezelter 4168 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
136     method for calculating electrostatic interactions between point
137 gezelter 4185 charges. They argued that the effective Coulomb interaction in most
138     condensed phase systems is effectively short
139     ranged.\cite{Wolf92,Wolf95} For an ordered lattice (e.g., when
140     computing the Madelung constant of an ionic solid), the material can
141     be considered as a set of ions interacting with neutral dipolar or
142     quadrupolar ``molecules'' giving an effective distance dependence for
143     the electrostatic interactions of $r^{-5}$ (see figure
144     \ref{fig:schematic}). If one views the \ce{NaCl} crystal as a simple
145     cubic (SC) structure with an octupolar \ce{(NaCl)4} basis, the
146     electrostatic energy per ion converges more rapidly to the Madelung
147     energy than the dipolar approximation.\cite{Wolf92} To find the
148     correct Madelung constant, Lacman suggested that the NaCl structure
149     could be constructed in a way that the finite crystal terminates with
150     complete \ce{(NaCl)4} molecules.\cite{Lacman65} The central ion sees
151     what is effectively a set of octupoles at large distances. These facts
152     suggest that the Madelung constants are relatively short ranged for
153     perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
154 gezelter 4186 application of Wolf's method can provide accurate estimates of
155 gezelter 4185 Madelung constants using relatively short cutoff radii.
156    
157     Direct truncation of interactions at a cutoff radius creates numerical
158 gezelter 4186 errors. Wolf \textit{et al.} suggest that truncation errors are due
159 gezelter 4185 to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
160     neutralize this charge they proposed placing an image charge on the
161     surface of the cutoff sphere for every real charge inside the cutoff.
162     These charges are present for the evaluation of both the pair
163     interaction energy and the force, although the force expression
164 gezelter 4186 maintains a discontinuity at the cutoff sphere. In the original Wolf
165 gezelter 4185 formulation, the total energy for the charge and image were not equal
166 gezelter 4186 to the integral of the force expression, and as a result, the total
167 gezelter 4185 energy would not be conserved in molecular dynamics (MD)
168 gezelter 4214 simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennell and
169 gezelter 4185 Gezelter later proposed shifted force variants of the Wolf method with
170     commensurate force and energy expressions that do not exhibit this
171 gezelter 4186 problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
172     were also proposed by Chen \textit{et
173 gezelter 4185 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
174 gezelter 4212 and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfully
175 gezelter 4186 used additional neutralization of higher order moments for systems of
176     point charges.\cite{Fukuda:2013sf}
177 mlamichh 4114
178 gezelter 4181 \begin{figure}
179 gezelter 4167 \centering
180 gezelter 4190 \includegraphics[width=\linewidth]{schematic.eps}
181 gezelter 4181 \caption{Top: Ionic systems exhibit local clustering of dissimilar
182     charges (in the smaller grey circle), so interactions are
183 gezelter 4184 effectively charge-multipole at longer distances. With hard
184     cutoffs, motion of individual charges in and out of the cutoff
185     sphere can break the effective multipolar ordering. Bottom:
186     dipolar crystals and fluids have a similar effective
187 gezelter 4181 \textit{quadrupolar} ordering (in the smaller grey circles), and
188     orientational averaging helps to reduce the effective range of the
189     interactions in the fluid. Placement of reversed image multipoles
190     on the surface of the cutoff sphere recovers the effective
191 gezelter 4208 higher-order multipole behavior. \label{fig:schematic}}
192 gezelter 4167 \end{figure}
193 mlamichh 4114
194 gezelter 4185 One can make a similar effective range argument for crystals of point
195     \textit{multipoles}. The Luttinger and Tisza treatment of energy
196     constants for dipolar lattices utilizes 24 basis vectors that contain
197 gezelter 4186 dipoles at the eight corners of a unit cube.\cite{LT} Only three of
198     these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
199 gezelter 4185 moments, while the rest have zero net dipole and retain contributions
200 gezelter 4186 only from higher order multipoles. The lowest-energy crystalline
201 gezelter 4185 structures are built out of basis vectors that have only residual
202     quadrupolar moments (e.g. the $Z_5$ array). In these low energy
203     structures, the effective interaction between a dipole at the center
204 gezelter 4167 of a crystal and a group of eight dipoles farther away is
205     significantly shorter ranged than the $r^{-3}$ that one would expect
206     for raw dipole-dipole interactions. Only in crystals which retain a
207     bulk dipole moment (e.g. ferroelectrics) does the analogy with the
208     ionic crystal break down -- ferroelectric dipolar crystals can exist,
209     while ionic crystals with net charge in each unit cell would be
210     unstable.
211    
212     In ionic crystals, real-space truncation can break the effective
213 gezelter 4181 multipolar arrangements (see Fig. \ref{fig:schematic}), causing
214     significant swings in the electrostatic energy as individual ions move
215     back and forth across the boundary. This is why the image charges are
216 gezelter 4180 necessary for the Wolf sum to exhibit rapid convergence. Similarly,
217     the real-space truncation of point multipole interactions breaks
218     higher order multipole arrangements, and image multipoles are required
219     for real-space treatments of electrostatic energies.
220 gezelter 4167
221 gezelter 4181 The shorter effective range of electrostatic interactions is not
222     limited to perfect crystals, but can also apply in disordered fluids.
223 gezelter 4186 Even at elevated temperatures, there is local charge balance in an
224     ionic liquid, where each positive ion has surroundings dominated by
225 gezelter 4207 negative ions and vice versa. The reversed-charge images on the
226 gezelter 4214 cutoff sphere that are integral to the Wolf and damped shifted force
227     (DSF) approaches retain the effective multipolar interactions as the
228     charges traverse the cutoff boundary.
229 gezelter 4181
230     In multipolar fluids (see Fig. \ref{fig:schematic}) there is
231     significant orientational averaging that additionally reduces the
232     effect of long-range multipolar interactions. The image multipoles
233 gezelter 4214 that are introduced in the Taylor shifted force (TSF), gradient
234     shifted force (GSF), and shifted potential (SP) methods mimic this effect
235 gezelter 4181 and reduce the effective range of the multipolar interactions as
236     interacting molecules traverse each other's cutoff boundaries.
237    
238 gezelter 4186 Forces and torques acting on atomic sites are fundamental in driving
239 gezelter 4214 dynamics in molecular simulations, and the DSF energy kernel provides
240     consistent energies and forces on charged atoms within the cutoff
241     sphere. Both the energy and the force go smoothly to zero as an atom
242     approaches the cutoff radius. The comparisons of the accuracy these
243     quantities between the DSF kernel and SPME was surprisingly
244     good.\cite{Fennell:2006lq} As a result, the DSF method has seen
245     increasing use in molecular systems with relatively uniform charge
246 gezelter 4186 densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
247 gezelter 4167
248 gezelter 4168 \subsection{The damping function}
249 gezelter 4185 The damping function has been discussed in detail in the first paper
250     of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
251     interactions between point charges can be replaced by the
252     complementary error function $\mathrm{erfc}(\alpha r)/r$ to accelerate
253     convergence, where $\alpha$ is a damping parameter with units of
254     inverse distance. Altering the value of $\alpha$ is equivalent to
255     changing the width of Gaussian charge distributions that replace each
256     point charge, as Coulomb integrals with Gaussian charge distributions
257     produce complementary error functions when truncated at a finite
258     distance.
259 mlamichh 4114
260 gezelter 4214 With moderate damping coefficients, $\alpha \sim 0.2$ \AA$^{-1}$, the DSF method
261 gezelter 4185 produced very good agreement with SPME for interaction energies,
262     forces and torques for charge-charge
263     interactions.\cite{Fennell:2006lq}
264 gezelter 4167
265 gezelter 4168 \subsection{Point multipoles in molecular modeling}
266     Coarse-graining approaches which treat entire molecular subsystems as
267     a single rigid body are now widely used. A common feature of many
268     coarse-graining approaches is simplification of the electrostatic
269     interactions between bodies so that fewer site-site interactions are
270 gezelter 4185 required to compute configurational
271     energies.\cite{Ren06,Essex10,Essex11}
272 mlamichh 4166
273 gezelter 4186 Additionally, because electrons in a molecule are not localized at
274     specific points, the assignment of partial charges to atomic centers
275     is always an approximation. For increased accuracy, atomic sites can
276     also be assigned point multipoles and polarizabilities. Recently,
277     water has been modeled with point multipoles up to octupolar order
278     using the soft sticky dipole-quadrupole-octupole (SSDQO)
279 gezelter 4180 model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
280 gezelter 4168 multipoles up to quadrupolar order have also been coupled with point
281     polarizabilities in the high-quality AMOEBA and iAMOEBA water
282 gezelter 4185 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
283     truncating point multipoles without smoothing the forces and torques
284 gezelter 4186 can create energy conservation issues in molecular dynamics
285     simulations.
286 mlamichh 4166
287 gezelter 4168 In this paper we test a set of real-space methods that were developed
288     for point multipolar interactions. These methods extend the damped
289     shifted force (DSF) and Wolf methods originally developed for
290     charge-charge interactions and generalize them for higher order
291 gezelter 4186 multipoles. The detailed mathematical development of these methods
292     has been presented in the first paper in this series, while this work
293     covers the testing of energies, forces, torques, and energy
294 gezelter 4168 conservation properties of the methods in realistic simulation
295     environments. In all cases, the methods are compared with the
296 gezelter 4186 reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
297 gezelter 4168
298    
299     \section{\label{sec:method}Review of Methods}
300     Any real-space electrostatic method that is suitable for MD
301     simulations should have the electrostatic energy, forces and torques
302     between two sites go smoothly to zero as the distance between the
303 gezelter 4208 sites, $r_{ab}$ approaches the cutoff radius, $r_c$. Requiring
304 gezelter 4168 this continuity at the cutoff is essential for energy conservation in
305     MD simulations. The mathematical details of the shifted potential
306     (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
307     methods have been discussed in detail in the previous paper in this
308     series.\cite{PaperI} Here we briefly review the new methods and
309     describe their essential features.
310 mlamichh 4166
311 gezelter 4168 \subsection{Taylor-shifted force (TSF)}
312 mlamichh 4114
313 gezelter 4168 The electrostatic potential energy between point multipoles can be
314     expressed as the product of two multipole operators and a Coulombic
315     kernel,
316 mlamichh 4114 \begin{equation}
317 gezelter 4206 U_{ab}(r)= M_{a} M_{b} \frac{1}{r} \label{kernel}.
318 mlamichh 4114 \end{equation}
319 gezelter 4206 where the multipole operator for site $a$, $M_{a}$, is
320     expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
321     $a$, etc.
322 mlamichh 4166
323 gezelter 4180 The TSF potential for any multipole-multipole interaction can be
324     written
325 gezelter 4168 \begin{equation}
326     U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
327     \label{generic}
328     \end{equation}
329 gezelter 4180 where $f_n(r)$ is a shifted kernel that is appropriate for the order
330 gezelter 4181 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
331     charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
332     and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
333     quadrupole-quadrupole. To ensure smooth convergence of the energy,
334     force, and torques, a Taylor expansion with $n$ terms must be
335     performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
336 gezelter 4168
337     For multipole-multipole interactions, following this procedure results
338 gezelter 4180 in separate radial functions for each of the distinct orientational
339     contributions to the potential, and ensures that the forces and
340     torques from each of these contributions will vanish at the cutoff
341     radius. For example, the direct dipole dot product
342 gezelter 4206 ($\mathbf{D}_{a}
343     \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
344 gezelter 4168 dot products:
345     \begin{equation}
346 gezelter 4214 U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 gezelter 4206 \mathbf{D}_{a} \cdot
348     \mathbf{D}_{b} \right) v_{21}(r) +
349     \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
350     \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
351 gezelter 4168 \end{equation}
352    
353 gezelter 4180 For the Taylor shifted (TSF) method with the undamped kernel,
354     $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
355     \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
356     - \frac{6}{r r_c^2}$. In these functions, one can easily see the
357     connection to unmodified electrostatics as well as the smooth
358     transition to zero in both these functions as $r\rightarrow r_c$. The
359     electrostatic forces and torques acting on the central multipole due
360 gezelter 4184 to another site within the cutoff sphere are derived from
361 gezelter 4168 Eq.~\ref{generic}, accounting for the appropriate number of
362     derivatives. Complete energy, force, and torque expressions are
363     presented in the first paper in this series (Reference
364 gezelter 4175 \onlinecite{PaperI}).
365 gezelter 4168
366     \subsection{Gradient-shifted force (GSF)}
367    
368 gezelter 4180 A second (and conceptually simpler) method involves shifting the
369     gradient of the raw Coulomb potential for each particular multipole
370 gezelter 4168 order. For example, the raw dipole-dipole potential energy may be
371     shifted smoothly by finding the gradient for two interacting dipoles
372     which have been projected onto the surface of the cutoff sphere
373     without changing their relative orientation,
374 gezelter 4181 \begin{equation}
375 gezelter 4214 U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r) = U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r) -
376     U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c)
377 gezelter 4206 - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
378 gezelter 4214 \nabla U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c).
379 gezelter 4181 \end{equation}
380 gezelter 4206 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
381 gezelter 4180 (although the signs are reversed for the dipole that has been
382     projected onto the cutoff sphere). In many ways, this simpler
383     approach is closer in spirit to the original shifted force method, in
384     that it projects a neutralizing multipole (and the resulting forces
385     from this multipole) onto a cutoff sphere. The resulting functional
386     forms for the potentials, forces, and torques turn out to be quite
387     similar in form to the Taylor-shifted approach, although the radial
388     contributions are significantly less perturbed by the gradient-shifted
389     approach than they are in the Taylor-shifted method.
390 gezelter 4168
391 gezelter 4180 For the gradient shifted (GSF) method with the undamped kernel,
392 gezelter 4210 $v_{21}(r) = -\frac{1}{r^3} - \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
393 gezelter 4180 $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
394     Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
395     because the Taylor expansion retains only one term, they are
396     significantly less perturbed than the TSF functions.
397    
398 gezelter 4168 In general, the gradient shifted potential between a central multipole
399     and any multipolar site inside the cutoff radius is given by,
400     \begin{equation}
401 gezelter 4206 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
402     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
403     \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
404 gezelter 4168 \label{generic2}
405     \end{equation}
406     where the sum describes a separate force-shifting that is applied to
407 gezelter 4184 each orientational contribution to the energy. In this expression,
408     $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
409 gezelter 4206 ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
410 gezelter 4184 represent the orientations the multipoles.
411 gezelter 4168
412     The third term converges more rapidly than the first two terms as a
413     function of radius, hence the contribution of the third term is very
414     small for large cutoff radii. The force and torque derived from
415 gezelter 4184 Eq. \ref{generic2} are consistent with the energy expression and
416 gezelter 4175 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
417 gezelter 4168 can be considered generalizations of the original DSF method for
418     higher order multipole interactions. GSF and TSF are also identical up
419     to the charge-dipole interaction but generate different expressions in
420     the energy, force and torque for higher order multipole-multipole
421     interactions. Complete energy, force, and torque expressions for the
422     GSF potential are presented in the first paper in this series
423 gezelter 4184 (Reference~\onlinecite{PaperI}).
424 gezelter 4168
425    
426 mlamichh 4166 \subsection{Shifted potential (SP) }
427 gezelter 4168 A discontinuous truncation of the electrostatic potential at the
428     cutoff sphere introduces a severe artifact (oscillation in the
429     electrostatic energy) even for molecules with the higher-order
430     multipoles.\cite{PaperI} We have also formulated an extension of the
431     Wolf approach for point multipoles by simply projecting the image
432     multipole onto the surface of the cutoff sphere, and including the
433     interactions with the central multipole and the image. This
434     effectively shifts the total potential to zero at the cutoff radius,
435 mlamichh 4166 \begin{equation}
436 gezelter 4206 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
437     U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
438 mlamichh 4166 \label{eq:SP}
439     \end{equation}
440 gezelter 4168 where the sum describes separate potential shifting that is done for
441     each orientational contribution to the energy (e.g. the direct dipole
442     product contribution is shifted {\it separately} from the
443     dipole-distance terms in dipole-dipole interactions). Note that this
444 gezelter 4175 is not a simple shifting of the total potential at $r_c$. Each radial
445 gezelter 4168 contribution is shifted separately. One consequence of this is that
446     multipoles that reorient after leaving the cutoff sphere can re-enter
447     the cutoff sphere without perturbing the total energy.
448 mlamichh 4166
449 gezelter 4180 For the shifted potential (SP) method with the undamped kernel,
450     $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
451     \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
452     central multipole and other multipolar sites goes smoothly to zero as
453     $r \rightarrow r_c$. However, the force and torque obtained from the
454     shifted potential (SP) are discontinuous at $r_c$. MD simulations
455     will still experience energy drift while operating under the SP
456     potential, but it may be suitable for Monte Carlo approaches where the
457     configurational energy differences are the primary quantity of
458     interest.
459 gezelter 4168
460 gezelter 4180 \subsection{The Self Term}
461 gezelter 4168 In the TSF, GSF, and SP methods, a self-interaction is retained for
462     the central multipole interacting with its own image on the surface of
463     the cutoff sphere. This self interaction is nearly identical with the
464     self-terms that arise in the Ewald sum for multipoles. Complete
465     expressions for the self terms are presented in the first paper in
466 gezelter 4175 this series (Reference \onlinecite{PaperI}).
467 mlamichh 4162
468 gezelter 4168
469 gezelter 4170 \section{\label{sec:methodology}Methodology}
470 mlamichh 4166
471 gezelter 4170 To understand how the real-space multipole methods behave in computer
472     simulations, it is vital to test against established methods for
473     computing electrostatic interactions in periodic systems, and to
474     evaluate the size and sources of any errors that arise from the
475     real-space cutoffs. In the first paper of this series, we compared
476     the dipolar and quadrupolar energy expressions against analytic
477     expressions for ordered dipolar and quadrupolar
478 gezelter 4174 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
479     used the multipolar Ewald sum as a reference method for comparing
480     energies, forces, and torques for molecular models that mimic
481 gezelter 4175 disordered and ordered condensed-phase systems. The parameters used
482 gezelter 4207 in the test cases are given in table~\ref{tab:pars}.
483 gezelter 4174
484 gezelter 4175 \begin{table}
485 gezelter 4214 \caption{The parameters used in the systems used to evaluate the new
486     real-space methods. The most comprehensive test was a liquid
487     composed of 2000 soft DQ liquid molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
488     ions). This test exercises all orders of the multipolar
489     interactions developed in the first paper.\label{tab:pars}}
490 gezelter 4175 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
491     & \multicolumn{2}{c|}{LJ parameters} &
492     \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
493     Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
494     $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
495     $I_{zz}$ \\ \cline{6-8}\cline{10-12}
496     & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
497     \AA\textsuperscript{2})} \\ \hline
498     Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
499 gezelter 4180 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
500 gezelter 4175 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
501 gezelter 4180 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
502 gezelter 4214 Soft DQ liquid & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
503 gezelter 4175 \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
504     \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
505     \end{tabularx}
506     \end{table}
507     The systems consist of pure multipolar solids (both dipole and
508     quadrupole), pure multipolar liquids (both dipole and quadrupole), a
509     fluid composed of sites containing both dipoles and quadrupoles
510     simultaneously, and a final test case that includes ions with point
511     charges in addition to the multipolar fluid. The solid-phase
512     parameters were chosen so that the systems can explore some
513     orientational freedom for the multipolar sites, while maintaining
514 gezelter 4214 relatively strict translational order. The soft DQ liquid model used
515     here based loosely on the SSDQO water
516     model,\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} but is not itself a
517     particularly accurate water model. However, the soft DQ model does
518     test dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
519     interactions at roughly the same magnitudes. The last test case, a
520     soft DQ liquid with dissolved ions, exercises \textit{all} levels of
521     the multipole-multipole interactions we have derived so far and
522     represents the most complete test of the new methods.
523 mlamichh 4166
524 gezelter 4175 In the following section, we present results for the total
525     electrostatic energy, as well as the electrostatic contributions to
526     the force and torque on each molecule. These quantities have been
527     computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
528 gezelter 4180 and have been compared with the values obtained from the multipolar
529     Ewald sum. In Monte Carlo (MC) simulations, the energy differences
530 gezelter 4175 between two configurations is the primary quantity that governs how
531 gezelter 4191 the simulation proceeds. These differences are the most important
532 gezelter 4175 indicators of the reliability of a method even if the absolute
533     energies are not exact. For each of the multipolar systems listed
534     above, we have compared the change in electrostatic potential energy
535     ($\Delta E$) between 250 statistically-independent configurations. In
536     molecular dynamics (MD) simulations, the forces and torques govern the
537     behavior of the simulation, so we also compute the electrostatic
538     contributions to the forces and torques.
539    
540     \subsection{Implementation}
541     The real-space methods developed in the first paper in this series
542     have been implemented in our group's open source molecular simulation
543 gezelter 4187 program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
544 gezelter 4175 this work. The complementary error function can be a relatively slow
545     function on some processors, so all of the radial functions are
546     precomputed on a fine grid and are spline-interpolated to provide
547     values when required.
548    
549     Using the same simulation code, we compare to a multipolar Ewald sum
550     with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
551     the Ewald sum is a re-implementation of the algorithm originally
552     proposed by Smith that does not use the particle mesh or smoothing
553 gezelter 4207 approximations.\cite{Smith82,Smith98} This implementation was tested
554     extensively against the analytic energy constants for the multipolar
555     lattices that are discussed in reference \onlinecite{PaperI}. In all
556     cases discussed below, the quantities being compared are the
557     electrostatic contributions to energies, force, and torques. All
558     other contributions to these quantities (i.e. from Lennard-Jones
559     interactions) are removed prior to the comparisons.
560 gezelter 4175
561     The convergence parameter ($\alpha$) also plays a role in the balance
562     of the real-space and reciprocal-space portions of the Ewald
563     calculation. Typical molecular mechanics packages set this to a value
564     that depends on the cutoff radius and a tolerance (typically less than
565     $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
566     associated with increasing accuracy at the expense of computational
567     time spent on the reciprocal-space portion of the
568     summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
569     10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
570     Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
571    
572     The real-space models have self-interactions that provide
573     contributions to the energies only. Although the self interaction is
574     a rapid calculation, we note that in systems with fluctuating charges
575     or point polarizabilities, the self-term is not static and must be
576     recomputed at each time step.
577    
578 gezelter 4170 \subsection{Model systems}
579 gezelter 4180 To sample independent configurations of the multipolar crystals, body
580     centered cubic (bcc) crystals, which exhibit the minimum energy
581     structures for point dipoles, were generated using 3,456 molecules.
582     The multipoles were translationally locked in their respective crystal
583     sites for equilibration at a relatively low temperature (50K) so that
584     dipoles or quadrupoles could freely explore all accessible
585     orientations. The translational constraints were then removed, the
586     systems were re-equilibrated, and the crystals were simulated for an
587     additional 10 ps in the microcanonical (NVE) ensemble with an average
588     temperature of 50 K. The balance between moments of inertia and
589     particle mass were chosen to allow orientational sampling without
590     significant translational motion. Configurations were sampled at
591     equal time intervals in order to compare configurational energy
592     differences. The crystals were simulated far from the melting point
593     in order to avoid translational deformation away of the ideal lattice
594     geometry.
595 gezelter 4170
596 gezelter 4180 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
597     simulations, each system was created with 2,048 randomly-oriented
598     molecules. These were equilibrated at a temperature of 300K for 1 ns.
599     Each system was then simulated for 1 ns in the microcanonical (NVE)
600 gezelter 4214 ensemble with the Dullweber, Leimkuhler, and McLachlan (DLM)
601     symplectic splitting integrator using 1 fs
602     timesteps.\cite{Dullweber1997} We collected 250 different
603     configurations at equal time intervals. For the liquid system that
604     included ionic species, we converted 48 randomly-distributed molecules
605     into 24 \ce{Na+} and 24 \ce{Cl-} ions and re-equilibrated. After
606     equilibration, the system was run under the same conditions for 1
607     ns. A total of 250 configurations were collected. In the following
608     comparisons of energies, forces, and torques, the Lennard-Jones
609     potentials were turned off and only the purely electrostatic
610     quantities were compared with the same values obtained via the Ewald
611     sum.
612 gezelter 4170
613     \subsection{Accuracy of Energy Differences, Forces and Torques}
614     The pairwise summation techniques (outlined above) were evaluated for
615     use in MC simulations by studying the energy differences between
616     different configurations. We took the Ewald-computed energy
617     difference between two conformations to be the correct behavior. An
618     ideal performance by one of the new methods would reproduce these
619     energy differences exactly. The configurational energies being used
620     here contain only contributions from electrostatic interactions.
621     Lennard-Jones interactions were omitted from the comparison as they
622     should be identical for all methods.
623    
624     Since none of the real-space methods provide exact energy differences,
625 gezelter 4180 we used least square regressions analysis for the six different
626 gezelter 4170 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
627 gezelter 4214 with the multipolar Ewald reference method. A result of unity for
628     both the correlation (slope) and coefficient of determination ($R^2$)
629     for these regressions would indicate perfect agreement between the
630     real-space method and the multipolar Ewald sum.
631 gezelter 4170
632     Molecular systems were run long enough to explore independent
633     configurations and 250 configurations were recorded for comparison.
634     Each system provided 31,125 energy differences for a total of 186,750
635     data points. Similarly, the magnitudes of the forces and torques have
636 gezelter 4180 also been compared using least squares regression analysis. In the
637 gezelter 4170 forces and torques comparison, the magnitudes of the forces acting in
638     each molecule for each configuration were evaluated. For example, our
639     dipolar liquid simulation contains 2048 molecules and there are 250
640     different configurations for each system resulting in 3,072,000 data
641     points for comparison of forces and torques.
642    
643 mlamichh 4166 \subsection{Analysis of vector quantities}
644 gezelter 4170 Getting the magnitudes of the force and torque vectors correct is only
645     part of the issue for carrying out accurate molecular dynamics
646     simulations. Because the real space methods reweight the different
647     orientational contributions to the energies, it is also important to
648     understand how the methods impact the \textit{directionality} of the
649 gezelter 4212 force and torque vectors. Fisher developed a probability density
650 gezelter 4170 function to analyse directional data sets,
651 mlamichh 4162 \begin{equation}
652 gezelter 4170 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
653 mlamichh 4162 \label{eq:pdf}
654     \end{equation}
655 gezelter 4170 where $\kappa$ measures directional dispersion of the data around the
656     mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
657     estimated as a reciprocal of the circular variance.\cite{Allen91} To
658     quantify the directional error, forces obtained from the Ewald sum
659     were taken as the mean (or correct) direction and the angle between
660     the forces obtained via the Ewald sum and the real-space methods were
661     evaluated,
662 mlamichh 4162 \begin{equation}
663 gezelter 4214 \cos\theta_i = \frac{\mathbf{f}_i^\mathrm{~Ewald} \cdot
664     \mathbf{f}_i^\mathrm{~GSF}}{\left|\mathbf{f}_i^\mathrm{~Ewald}\right| \left|\mathbf{f}_i^\mathrm{~GSF}\right|}
665 gezelter 4170 \end{equation}
666     The total angular displacement of the vectors was calculated as,
667     \begin{equation}
668     R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
669 mlamichh 4162 \label{eq:displacement}
670     \end{equation}
671 gezelter 4170 where $N$ is number of force vectors. The circular variance is
672     defined as
673     \begin{equation}
674     \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
675     \end{equation}
676     The circular variance takes on values between from 0 to 1, with 0
677     indicating a perfect directional match between the Ewald force vectors
678     and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
679     correspond to higher values of $\kappa$, which indicates tighter
680     clustering of the real-space force vectors around the Ewald forces.
681 mlamichh 4162
682 gezelter 4170 A similar analysis was carried out for the electrostatic contribution
683     to the molecular torques as well as forces.
684    
685 mlamichh 4166 \subsection{Energy conservation}
686 gezelter 4170 To test conservation the energy for the methods, the mixed molecular
687 gezelter 4214 system of 2000 soft DQ liquid molecules with 24 \ce{Na+} and 24
688     \ce{Cl-} ions was run for 1 ns in the microcanonical ensemble at an
689     average temperature of 300K. Each of the different electrostatic
690     methods (Ewald, Hard, SP, GSF, and TSF) was tested for a range of
691     different damping values. The molecular system was started with same
692     initial positions and velocities for all cutoff methods. The energy
693     drift ($\delta E_1$) and standard deviation of the energy about the
694     slope ($\delta E_0$) were evaluated from the total energy of the
695     system as a function of time. Although both measures are valuable at
696 gezelter 4170 investigating new methods for molecular dynamics, a useful interaction
697     model must allow for long simulation times with minimal energy drift.
698 mlamichh 4114
699 mlamichh 4166 \section{\label{sec:result}RESULTS}
700     \subsection{Configurational energy differences}
701 gezelter 4174
702 mlamichh 4114 \begin{figure}
703 gezelter 4174 \centering
704 gezelter 4214 \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined.eps}
705 gezelter 4174 \caption{Statistical analysis of the quality of configurational
706     energy differences for the real-space electrostatic methods
707     compared with the reference Ewald sum. Results with a value equal
708     to 1 (dashed line) indicate $\Delta E$ values indistinguishable
709     from those obtained using the multipolar Ewald sum. Different
710     values of the cutoff radius are indicated with different symbols
711 gezelter 4214 (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = inverted
712 gezelter 4208 triangles).\label{fig:slopeCorr_energy}}
713 gezelter 4174 \end{figure}
714    
715 gezelter 4214 The combined coefficient of determination and slope for all six
716     systems is shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the
717     methods reproduce the Ewald configurational energy differences with
718     remarkable fidelity. Undamped hard cutoffs introduce a significant
719     amount of random scatter in the energy differences which is apparent
720     in the reduced value of $R^2$ for this method. This can be easily
721     understood as configurations which exhibit small traversals of a few
722     dipoles or quadrupoles out of the cutoff sphere will see large energy
723     jumps when hard cutoffs are used. The orientations of the multipoles
724     (particularly in the ordered crystals) mean that these energy jumps
725     can go in either direction, producing a significant amount of random
726     scatter, but no systematic error.
727 gezelter 4174
728     The TSF method produces energy differences that are highly correlated
729     with the Ewald results, but it also introduces a significant
730     systematic bias in the values of the energies, particularly for
731     smaller cutoff values. The TSF method alters the distance dependence
732     of different orientational contributions to the energy in a
733     non-uniform way, so the size of the cutoff sphere can have a large
734 gezelter 4175 effect, particularly for the crystalline systems.
735 gezelter 4174
736     Both the SP and GSF methods appear to reproduce the Ewald results with
737 gezelter 4214 excellent fidelity, particularly for moderate damping ($\alpha \approx
738     0.2$~\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
739     12$~\AA). With the exception of the undamped hard cutoff, and the TSF
740 gezelter 4175 method with short cutoffs, all of the methods would be appropriate for
741     use in Monte Carlo simulations.
742 gezelter 4174
743 mlamichh 4114 \subsection{Magnitude of the force and torque vectors}
744 gezelter 4174
745 gezelter 4175 The comparisons of the magnitudes of the forces and torques for the
746     data accumulated from all six systems are shown in Figures
747 gezelter 4174 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
748     correlation and slope for the forces agree well with the Ewald sum
749 gezelter 4175 even for the hard cutoffs.
750 gezelter 4174
751 gezelter 4175 For systems of molecules with only multipolar interactions, the pair
752     energy contributions are quite short ranged. Moreover, the force
753     decays more rapidly than the electrostatic energy, hence the hard
754     cutoff method can also produce reasonable agreement for this quantity.
755     Although the pure cutoff gives reasonably good electrostatic forces
756     for pairs of molecules included within each other's cutoff spheres,
757     the discontinuity in the force at the cutoff radius can potentially
758     cause energy conservation problems as molecules enter and leave the
759     cutoff spheres. This is discussed in detail in section
760     \ref{sec:conservation}.
761 gezelter 4174
762     The two shifted-force methods (GSF and TSF) exhibit a small amount of
763     systematic variation and scatter compared with the Ewald forces. The
764     shifted-force models intentionally perturb the forces between pairs of
765 gezelter 4175 molecules inside each other's cutoff spheres in order to correct the
766     energy conservation issues, and this perturbation is evident in the
767     statistics accumulated for the molecular forces. The GSF
768 gezelter 4180 perturbations are minimal, particularly for moderate damping and
769 gezelter 4214 commonly-used cutoff values ($r_c = 12$~\AA). The TSF method shows
770     reasonable agreement in $R^2$, but again the systematic error in the
771     forces is concerning if replication of Ewald forces is desired.
772 gezelter 4174
773 gezelter 4208 It is important to note that the forces and torques from the SP and
774     the Hard cutoffs are not identical. The SP method shifts each
775     orientational contribution separately (e.g. the dipole-dipole dot
776     product is shifted by a different function than the dipole-distance
777     products), while the hard cutoff contains no orientation-dependent
778     shifting. The forces and torques for these methods therefore diverge
779     for multipoles even though the forces for point charges are identical.
780    
781 mlamichh 4114 \begin{figure}
782 gezelter 4174 \centering
783 gezelter 4191 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
784 gezelter 4174 \caption{Statistical analysis of the quality of the force vector
785     magnitudes for the real-space electrostatic methods compared with
786     the reference Ewald sum. Results with a value equal to 1 (dashed
787     line) indicate force magnitude values indistinguishable from those
788     obtained using the multipolar Ewald sum. Different values of the
789 gezelter 4214 cutoff radius are indicated with different symbols (9~\AA\ =
790     circles, 12~\AA\ = squares, and 15~\AA\ = inverted
791 gezelter 4208 triangles).\label{fig:slopeCorr_force}}
792 gezelter 4174 \end{figure}
793    
794    
795 mlamichh 4114 \begin{figure}
796 gezelter 4174 \centering
797 gezelter 4191 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
798 gezelter 4174 \caption{Statistical analysis of the quality of the torque vector
799     magnitudes for the real-space electrostatic methods compared with
800     the reference Ewald sum. Results with a value equal to 1 (dashed
801     line) indicate force magnitude values indistinguishable from those
802     obtained using the multipolar Ewald sum. Different values of the
803 gezelter 4214 cutoff radius are indicated with different symbols (9~\AA\ =
804     circles, 12~\AA\ = squares, and 15~\AA\ = inverted
805 gezelter 4208 triangles).\label{fig:slopeCorr_torque}}
806 gezelter 4174 \end{figure}
807    
808     The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
809     significantly influenced by the choice of real-space method. The
810     torque expressions have the same distance dependence as the energies,
811     which are naturally longer-ranged expressions than the inter-site
812     forces. Torques are also quite sensitive to orientations of
813     neighboring molecules, even those that are near the cutoff distance.
814    
815     The results shows that the torque from the hard cutoff method
816     reproduces the torques in quite good agreement with the Ewald sum.
817 gezelter 4175 The other real-space methods can cause some deviations, but excellent
818     agreement with the Ewald sum torques is recovered at moderate values
819 gezelter 4214 of the damping coefficient ($\alpha \approx 0.2$~\AA$^{-1}$) and cutoff
820     radius ($r_c \ge 12$~\AA). The TSF method exhibits only fair agreement
821 gezelter 4175 in the slope when compared with the Ewald torques even for larger
822     cutoff radii. It appears that the severity of the perturbations in
823     the TSF method are most in evidence for the torques.
824 gezelter 4174
825 mlamichh 4114 \subsection{Directionality of the force and torque vectors}
826 mlamichh 4162
827 gezelter 4174 The accurate evaluation of force and torque directions is just as
828     important for molecular dynamics simulations as the magnitudes of
829     these quantities. Force and torque vectors for all six systems were
830     analyzed using Fisher statistics, and the quality of the vector
831     directionality is shown in terms of circular variance
832 gezelter 4214 ($\mathrm{Var}(\theta)$) in
833     Fig. \ref{fig:slopeCorr_circularVariance}. The force and torque
834     vectors from the new real-space methods exhibit nearly-ideal Fisher
835     probability distributions (Eq.~\ref{eq:pdf}). Both the hard and SP
836     cutoff methods exhibit the best vectorial agreement with the Ewald
837     sum. The force and torque vectors from GSF method also show good
838     agreement with the Ewald method, which can also be systematically
839     improved by using moderate damping and a reasonable cutoff radius. For
840     $\alpha = 0.2$~\AA$^{-1}$ and $r_c = 12$~\AA, we observe
841     $\mathrm{Var}(\theta) = 0.00206$, which corresponds to a distribution
842     with 95\% of force vectors within $6.37^\circ$ of the corresponding
843     Ewald forces. The TSF method produces the poorest agreement with the
844     Ewald force directions.
845 gezelter 4174
846 gezelter 4175 Torques are again more perturbed than the forces by the new real-space
847     methods, but even here the variance is reasonably small. For the same
848 gezelter 4214 method (GSF) with the same parameters ($\alpha = 0.2$~\AA$^{-1}$, $r_c
849     = 12$~\AA), the circular variance was 0.01415, corresponds to a
850     distribution which has 95\% of torque vectors are within $16.75^\circ$
851     of the Ewald results. Again, the direction of the force and torque
852     vectors can be systematically improved by varying $\alpha$ and $r_c$.
853 gezelter 4174
854 mlamichh 4114 \begin{figure}
855 gezelter 4174 \centering
856 gezelter 4191 \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
857 gezelter 4174 \caption{The circular variance of the direction of the force and
858     torque vectors obtained from the real-space methods around the
859     reference Ewald vectors. A variance equal to 0 (dashed line)
860     indicates direction of the force or torque vectors are
861     indistinguishable from those obtained from the Ewald sum. Here
862     different symbols represent different values of the cutoff radius
863 gezelter 4214 (9~\AA\ = circle, 12~\AA\ = square, 15~\AA\ = inverted triangle)\label{fig:slopeCorr_circularVariance}}
864 gezelter 4174 \end{figure}
865 gezelter 4171
866 gezelter 4175 \subsection{Energy conservation\label{sec:conservation}}
867 gezelter 4171
868 gezelter 4174 We have tested the conservation of energy one can expect to see with
869 gezelter 4214 the new real-space methods using the soft DQ liquid model with a small
870 gezelter 4174 fraction of solvated ions. This is a test system which exercises all
871     orders of multipole-multipole interactions derived in the first paper
872     in this series and provides the most comprehensive test of the new
873 gezelter 4214 methods. A liquid-phase system was created with 2000 liquid-phase
874     molecules and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
875 gezelter 4206 temperature of 300K. After equilibration in the canonical (NVT)
876 gezelter 4221 ensemble using a Nos\'e-Hoover thermostat, six
877     statistically-independent replicas of this liquid-phase system were
878     run in the microcanonical (NVE) ensemble under the Ewald, Hard, SP,
879     GSF, and TSF methods with a cutoff radius of 12~\AA. The value of the
880     damping coefficient was also varied from the undamped case ($\alpha =
881     0$) to a heavily damped case ($\alpha = 0.3$~\AA$^{-1}$) for all of
882     the real space methods. A sample was also run using the multipolar
883     Ewald sum with the same real-space cutoff.
884 gezelter 4174
885     In figure~\ref{fig:energyDrift} we show the both the linear drift in
886     energy over time, $\delta E_1$, and the standard deviation of energy
887     fluctuations around this drift $\delta E_0$. Both of the
888     shifted-force methods (GSF and TSF) provide excellent energy
889 gezelter 4181 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
890 gezelter 4174 while the hard cutoff is essentially unusable for molecular dynamics.
891     SP provides some benefit over the hard cutoff because the energetic
892     jumps that happen as particles leave and enter the cutoff sphere are
893 gezelter 4175 somewhat reduced, but like the Wolf method for charges, the SP method
894     would not be as useful for molecular dynamics as either of the
895     shifted-force methods.
896 gezelter 4174
897     We note that for all tested values of the cutoff radius, the new
898     real-space methods can provide better energy conservation behavior
899 gezelter 4208 than the multipolar Ewald sum, even when relatively large $k$-space
900     cutoff values are utilized.
901 gezelter 4174
902 mlamichh 4114 \begin{figure}
903 gezelter 4171 \centering
904 gezelter 4221 \includegraphics[width=\textwidth]{finalDrift.eps}
905 gezelter 4214 \caption{Energy conservation of the real-space methods for the soft
906 gezelter 4221 DQ liquid / ion system. $\delta \mathrm{E}_1$ is the linear drift
907 gezelter 4214 in energy over time (in kcal/mol/particle/ns) and $\delta
908 gezelter 4210 \mathrm{E}_0$ is the standard deviation of energy fluctuations
909     around this drift (in kcal/mol/particle). Points that appear in
910     the green region at the bottom exhibit better energy conservation
911     than would be obtained using common parameters for Ewald-based
912     electrostatics.\label{fig:energyDrift}}
913 gezelter 4171 \end{figure}
914    
915 gezelter 4206 \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
916     The most important test of the modified interaction potentials is the
917     fidelity with which they can reproduce structural features and
918     dynamical properties in a liquid. One commonly-utilized measure of
919     structural ordering is the pair distribution function, $g(r)$, which
920     measures local density deviations in relation to the bulk density. In
921     the electrostatic approaches studied here, the short-range repulsion
922     from the Lennard-Jones potential is identical for the various
923     electrostatic methods, and since short range repulsion determines much
924     of the local liquid ordering, one would not expect to see many
925     differences in $g(r)$. Indeed, the pair distributions are essentially
926     identical for all of the electrostatic methods studied (for each of
927 gezelter 4214 the different systems under investigation).
928 gezelter 4174
929 gezelter 4214 % An example of this agreement for the soft DQ liquid/ion system is
930     % shown in Fig. \ref{fig:gofr}.
931 gezelter 4203
932 gezelter 4214 % \begin{figure}
933     % \centering
934     % \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
935     % \caption{The pair distribution functions, $g(r)$, for the SSDQ
936     % water/ion system obtained using the different real-space methods are
937     % essentially identical with the result from the Ewald
938     % treatment.\label{fig:gofr}}
939     % \end{figure}
940    
941 gezelter 4212 There is a minor over-structuring of the first solvation shell when
942 gezelter 4210 using TSF or when overdamping with any of the real-space methods.
943     With moderate damping, GSF and SP produce pair distributions that are
944     identical (within numerical noise) to their Ewald counterparts. The
945 gezelter 4212 degree of over-structuring can be measured most easily using the
946 gezelter 4210 coordination number,
947     \begin{equation}
948     n_C = 4\pi\rho \int_{0}^{a}r^2\text{g}(r)dr,
949     \end{equation}
950     where $\rho$ is the number density of the site-site pair interactions,
951 gezelter 4211 and $a$ is the radial location of the minima following the first peak
952 gezelter 4214 in $g(r)$ ($a = 4.2$~\AA\ for the soft DQ liquid / ion system). The
953 gezelter 4210 coordination number is shown as a function of the damping coefficient
954 gezelter 4211 for all of the real space methods in Fig. \ref{fig:Props}.
955 gezelter 4203
956 gezelter 4210 A more demanding test of modified electrostatics is the average value
957     of the electrostatic energy $\langle U_\mathrm{elect} \rangle / N$
958     which is obtained by sampling the liquid-state configurations
959     experienced by a liquid evolving entirely under the influence of each
960 gezelter 4214 of the methods. In Fig. \ref{fig:Props} we demonstrate how $\langle
961 gezelter 4210 U_\mathrm{elect} \rangle / N$ varies with the damping parameter,
962 gezelter 4211 $\alpha$, for each of the methods.
963 gezelter 4206
964     As in the crystals studied in the first paper, damping is important
965     for converging the mean electrostatic energy values, particularly for
966     the two shifted force methods (GSF and TSF). A value of $\alpha
967 gezelter 4214 \approx 0.2$~\AA$^{-1}$ is sufficient to converge the SP and GSF
968 gezelter 4206 energies with a cutoff of 12 \AA, while shorter cutoffs require more
969 gezelter 4214 dramatic damping ($\alpha \approx 0.28$~\AA$^{-1}$ for $r_c = 9$~\AA).
970 gezelter 4206 Overdamping the real-space electrostatic methods occurs with $\alpha >
971 gezelter 4214 0.3$~\AA$^{-1}$, causing the estimate of the electrostatic energy to
972     drop below the Ewald results.
973 gezelter 4206
974 gezelter 4211 These ``optimal'' values of the damping coefficient for structural
975     features are similar to those observed for DSF electrostatics for
976     purely point-charge systems, and the range $\alpha= 0.175 \rightarrow
977 gezelter 4214 0.225$~\AA$^{-1}$ for $r_c = 12$~\AA\ appears to be an excellent
978 gezelter 4210 compromise for mixed charge/multipolar systems.
979 gezelter 4203
980     To test the fidelity of the electrostatic methods at reproducing
981 gezelter 4210 \textit{dynamics} in a multipolar liquid, it is also useful to look at
982 gezelter 4203 transport properties, particularly the diffusion constant,
983     \begin{equation}
984     D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
985     \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
986     \label{eq:diff}
987     \end{equation}
988     which measures long-time behavior and is sensitive to the forces on
989 gezelter 4211 the multipoles. The self-diffusion constants (D) were calculated from
990 gezelter 4210 linear fits to the long-time portion of the mean square displacement,
991 gezelter 4214 $\langle r^{2}(t) \rangle$.\cite{Allen87} In Fig. \ref{fig:Props} we
992 gezelter 4210 demonstrate how the diffusion constant depends on the choice of
993     real-space methods and the damping coefficient. Both the SP and GSF
994     methods can obtain excellent agreement with Ewald again using moderate
995     damping.
996 gezelter 4203
997     In addition to translational diffusion, orientational relaxation times
998     were calculated for comparisons with the Ewald simulations and with
999 gezelter 4211 experiments. These values were determined by calculating the
1000     orientational time correlation function,
1001 gezelter 4203 \begin{equation}
1002 gezelter 4206 C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1003     \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1004 gezelter 4203 \label{eq:OrientCorr}
1005     \end{equation}
1006 gezelter 4211 from the same 350 ps microcanonical trajectories that were used for
1007     translational diffusion. Here, $P_l$ is the Legendre polynomial of
1008     order $l$ and $\hat{\mathbf{A}}_\gamma$ is the unit vector for body
1009     axis $\gamma$. The reference frame used for our sample dipolar
1010 gezelter 4214 systems has the $z$-axis running along the dipoles, and for the soft
1011     DQ liquid model, the $y$-axis connects the two implied hydrogen-like
1012 gezelter 4211 positions. From the orientation autocorrelation functions, we can
1013     obtain time constants for rotational relaxation either by fitting to a
1014     multi-exponential model for the orientational relaxation, or by
1015     integrating the correlation functions.
1016 gezelter 4203
1017 gezelter 4214 In a good model for water, the orientational decay times would be
1018 gezelter 4211 comparable to water orientational relaxation times from nuclear
1019     magnetic resonance (NMR). The relaxation constant obtained from
1020     $C_2^y(t)$ is normally of experimental interest because it describes
1021     the relaxation of the principle axis connecting the hydrogen
1022     atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1023     of the dipole-dipole relaxation from a proton NMR signal and can
1024     provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1025 gezelter 4214 In Fig. \ref{fig:Props} we compare the $\tau_2^y$ and $\tau_2^z$
1026 gezelter 4211 values for the various real-space methods over a range of different
1027     damping coefficients. The rotational relaxation for the $z$ axis
1028     primarily probes the torques on the dipoles, while the relaxation for
1029     the $y$ axis is sensitive primarily to the quadrupolar torques.
1030 gezelter 4203
1031 gezelter 4210 \begin{figure}
1032 gezelter 4214 \includegraphics[width=\textwidth]{properties.eps}
1033 gezelter 4210 \caption{Comparison of the structural and dynamic properties for the
1034 gezelter 4214 combined multipolar liquid (soft DQ liquid + ions) for all of the
1035     real-space methods with $r_c = 12$~\AA. Electrostatic energies,
1036 gezelter 4210 $\langle U_\mathrm{elect} \rangle / N$ (in kcal / mol),
1037 gezelter 4211 coordination numbers, $n_C$, diffusion constants (in $10^{-5}
1038     \mathrm{cm}^2\mathrm{s}^{-1}$), and rotational correlation times
1039     (in ps) all show excellent agreement with Ewald results for
1040     damping coefficients in the range $\alpha= 0.175 \rightarrow
1041 gezelter 4214 0.225$~\AA$^{-1}$. \label{fig:Props}}
1042 gezelter 4210 \end{figure}
1043 gezelter 4203
1044 gezelter 4211 In Fig. \ref{fig:Props} it appears that values for $D$, $\tau_2^y$,
1045     and $\tau_2^z$ using the Ewald sum are reproduced with excellent
1046     fidelity by the GSF and SP methods. All of the real space methods can
1047     be \textit{overdamped}, which reduces the effective range of multipole
1048     interactions, causing structural and dynamical changes from the
1049     correct behavior. Because overdamping weakens orientational
1050     preferences between adjacent molecules, it manifests as too-rapid
1051     orientational decay coupled with faster diffusion and
1052     over-coordination of the liquid. Underdamping is less problematic for
1053     the SP and GSF methods, as their structural and dynamical properties
1054     still reproduce the Ewald results even in the completely undamped
1055     ($\alpha = 0$) case. An optimal range for the electrostatic damping
1056 gezelter 4214 parameter appears to be $\alpha= 0.175 \rightarrow 0.225$~\AA$^{-1}$
1057     for $r_c = 12$~\AA, which similar to the optimal range found for the
1058 gezelter 4211 damped shifted force potential for point charges.\cite{Fennell:2006lq}
1059 gezelter 4203
1060 mlamichh 4114 \section{CONCLUSION}
1061 gezelter 4175 In the first paper in this series, we generalized the
1062     charge-neutralized electrostatic energy originally developed by Wolf
1063     \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1064     up to quadrupolar order. The SP method is essentially a
1065     multipole-capable version of the Wolf model. The SP method for
1066     multipoles provides excellent agreement with Ewald-derived energies,
1067     forces and torques, and is suitable for Monte Carlo simulations,
1068     although the forces and torques retain discontinuities at the cutoff
1069     distance that prevents its use in molecular dynamics.
1070 gezelter 4170
1071 gezelter 4175 We also developed two natural extensions of the damped shifted-force
1072 gezelter 4208 (DSF) model originally proposed by Zahn {\it et al.} and extended by
1073 gezelter 4214 Fennell and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
1074 gezelter 4208 approaches provide smooth truncation of energies, forces, and torques
1075     at the real-space cutoff, and both converge to DSF electrostatics for
1076     point-charge interactions. The TSF model is based on a high-order
1077     truncated Taylor expansion which can be relatively perturbative inside
1078     the cutoff sphere. The GSF model takes the gradient from an images of
1079     the interacting multipole that has been projected onto the cutoff
1080     sphere to derive shifted force and torque expressions, and is a
1081     significantly more gentle approach.
1082 gezelter 4170
1083 gezelter 4211 The GSF method produces quantitative agreement with Ewald energies,
1084 gezelter 4212 forces, and torques. It also performs well in conserving energy in MD
1085 gezelter 4208 simulations. The Taylor-shifted (TSF) model provides smooth dynamics,
1086     but these take place on a potential energy surface that is
1087     significantly perturbed from Ewald-based electrostatics. Because it
1088     performs relatively poorly compared with GSF, it may seem odd that
1089     that the TSF model was included in this work. However, the functional
1090     forms derived for the SP and GSF methods depend on the separation of
1091     orientational contributions that were made visible by the Taylor
1092     series of the electrostatic kernel at the cutoff radius. The TSF
1093     method also has the unique property that a large number of derivatives
1094     can be made to vanish at the cutoff radius. This property has proven
1095 gezelter 4214 useful in past treatments of the corrections to the Clausius-Mossotti
1096     fluctuation formula for dielectric constants.\cite{Izvekov:2008wo}
1097 gezelter 4175
1098 gezelter 4208 Reproduction of both structural and dynamical features in the liquid
1099     systems is remarkably good for both the SP and GSF models. Pair
1100     distribution functions are essentially equivalent to the same
1101     functions produced using Ewald-based electrostatics, and with moderate
1102     damping, a structural feature that directly probes the electrostatic
1103     interaction (e.g. the mean electrostatic potential energy) can also be
1104     made quantitative. Dynamical features are sensitive probes of the
1105     forces and torques produced by these methods, and even though the
1106     smooth behavior of forces is produced by perturbing the overall
1107     potential, the diffusion constants and orientational correlation times
1108     are quite close to the Ewald-based results.
1109 gezelter 4175
1110     The only cases we have found where the new GSF and SP real-space
1111     methods can be problematic are those which retain a bulk dipole moment
1112     at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1113     materials, uniform weighting of the orientational contributions can be
1114     important for converging the total energy. In these cases, the
1115     damping function which causes the non-uniform weighting can be
1116     replaced by the bare electrostatic kernel, and the energies return to
1117     the expected converged values.
1118    
1119 gezelter 4208 Based on the results of this work, we can conclude that the GSF method
1120     is a suitable and efficient replacement for the Ewald sum for
1121     evaluating electrostatic interactions in modern MD simulations, and
1122 gezelter 4212 the SP method would be an excellent choice for Monte Carlo
1123 gezelter 4208 simulations where smooth forces and energy conservation are not
1124     important. Both the SP and GSF methods retain excellent fidelity to
1125     the Ewald energies, forces and torques. Additionally, the energy
1126     drift and fluctuations from the GSF electrostatics are significantly
1127 gezelter 4211 better than a multipolar Ewald sum for finite-sized reciprocal spaces,
1128     and physical properties are reproduced accurately.
1129 gezelter 4175
1130 gezelter 4208 As in all purely pairwise cutoff methods, the SP, GSF and TSF methods
1131     are expected to scale approximately {\it linearly} with system size,
1132     and are easily parallelizable. This should result in substantial
1133     reductions in the computational cost of performing large simulations.
1134     With the proper use of pre-computation and spline interpolation of the
1135     radial functions, the real-space methods are essentially the same cost
1136     as a simple real-space cutoff. They require no Fourier transforms or
1137     $k$-space sums, and guarantee the smooth handling of energies, forces,
1138     and torques as multipoles cross the real-space cutoff boundary.
1139    
1140     We are not suggesting that there is any flaw with the Ewald sum; in
1141     fact, it is the standard by which the SP, GSF, and TSF methods have
1142     been judged in this work. However, these results provide evidence
1143     that in the typical simulations performed today, the Ewald summation
1144     may no longer be required to obtain the level of accuracy most
1145     researchers have come to expect.
1146    
1147 gezelter 4180 \begin{acknowledgments}
1148     JDG acknowledges helpful discussions with Christopher
1149     Fennell. Support for this project was provided by the National
1150     Science Foundation under grant CHE-1362211. Computational time was
1151     provided by the Center for Research Computing (CRC) at the
1152     University of Notre Dame.
1153     \end{acknowledgments}
1154    
1155 gezelter 4167 %\bibliographystyle{aip}
1156 gezelter 4168 \newpage
1157 mlamichh 4114 \bibliography{references}
1158     \end{document}
1159    
1160     %
1161     % ****** End of file aipsamp.tex ******