ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4206
Committed: Thu Aug 7 20:53:27 2014 UTC (9 years, 10 months ago) by gezelter
Content type: application/x-tex
File size: 68624 byte(s)
Log Message:
latest

File Contents

# Content
1 % ****** 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 aip,jcp,
24 amsmath,amssymb,
25 preprint,
26 %reprint,%
27 %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 %\usepackage{bm}% bold math
34 %\usepackage[mathlines]{lineno}% Enable numbering of text and display math
35 %\linenumbers\relax % Commence numbering lines
36 \usepackage{amsmath}
37 \usepackage{times}
38 \usepackage{mathptmx}
39 \usepackage{tabularx}
40 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
41 \usepackage{url}
42 \usepackage[english]{babel}
43
44 \newcolumntype{Y}{>{\centering\arraybackslash}X}
45
46 \begin{document}
47
48 %\preprint{AIP/123-QED}
49
50 \title{Real space electrostatics for multipoles. II. Comparisons with
51 the Ewald Sum}
52
53 \author{Madan Lamichhane}
54 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
55
56 \author{Kathie E. Newman}
57 \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
58
59 \author{J. Daniel Gezelter}%
60 \email{gezelter@nd.edu.}
61 \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
62 }
63
64 \date{\today}
65
66 \begin{abstract}
67 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 \end{abstract}
86
87 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
88 % Classification Scheme.
89 %\keywords{Electrostatics, Multipoles, Real-space}
90
91 \maketitle
92
93 \section{\label{sec:intro}Introduction}
94 Computing the interactions between electrostatic sites is one of the
95 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 space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
104
105 When carried out as originally formulated, the reciprocal-space
106 portion of the Ewald sum exhibits relatively poor computational
107 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 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
111 methods can decrease the computational cost from $O(N^2)$ down to $O(N
112 \log
113 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
114
115 Because of the artificial periodicity required for the Ewald sum,
116 interfacial molecular systems such as membranes and liquid-vapor
117 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 interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
122 These methods were originally quite computationally
123 expensive.\cite{Spohr97,Yeh99} There have been several successful
124 efforts that reduced the computational cost of 2-D lattice summations,
125 bringing them more in line with the scaling for the full 3-D
126 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
131 \subsection{Real-space methods}
132 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
133 method for calculating electrostatic interactions between point
134 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 application of Wolf's method can provide accurate estimates of
152 Madelung constants using relatively short cutoff radii.
153
154 Direct truncation of interactions at a cutoff radius creates numerical
155 errors. Wolf \textit{et al.} suggest that truncation errors are due
156 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 maintains a discontinuity at the cutoff sphere. In the original Wolf
162 formulation, the total energy for the charge and image were not equal
163 to the integral of the force expression, and as a result, the total
164 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 problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
169 were also proposed by Chen \textit{et
170 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
171 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
175 \begin{figure}
176 \centering
177 \includegraphics[width=\linewidth]{schematic.eps}
178 \caption{Top: Ionic systems exhibit local clustering of dissimilar
179 charges (in the smaller grey circle), so interactions are
180 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 \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 \end{figure}
191
192 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 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 moments, while the rest have zero net dipole and retain contributions
198 only from higher order multipoles. The lowest-energy crystalline
199 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 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 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 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
219 The shorter effective range of electrostatic interactions is not
220 limited to perfect crystals, but can also apply in disordered fluids.
221 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
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 % 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 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
256 \subsection{The damping function}
257 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
268 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
273 \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 required to compute configurational
279 energies.\cite{Ren06,Essex10,Essex11}
280
281 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 model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
288 multipoles up to quadrupolar order have also been coupled with point
289 polarizabilities in the high-quality AMOEBA and iAMOEBA water
290 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
291 truncating point multipoles without smoothing the forces and torques
292 can create energy conservation issues in molecular dynamics
293 simulations.
294
295 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 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 conservation properties of the methods in realistic simulation
303 environments. In all cases, the methods are compared with the
304 reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
305
306
307 %\subsection{Conservation of total energy }
308 %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
310 \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
322 \subsection{Taylor-shifted force (TSF)}
323
324 The electrostatic potential energy between point multipoles can be
325 expressed as the product of two multipole operators and a Coulombic
326 kernel,
327 \begin{equation}
328 U_{ab}(r)= M_{a} M_{b} \frac{1}{r} \label{kernel}.
329 \end{equation}
330 where the multipole operator for site $a$, $M_{a}$, is
331 expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
332 $a$, etc.
333
334 % Interactions between multipoles can be expressed as higher derivatives
335 % of the bare Coulomb potential, so one way of ensuring that the forces
336 % and torques vanish at the cutoff distance is to include a larger
337 % number of terms in the truncated Taylor expansion, e.g.,
338 % %
339 % \begin{equation}
340 % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
341 % \end{equation}
342 % %
343 % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344 % Thus, for $f(r)=1/r$, we find
345 % %
346 % \begin{equation}
347 % f_1(r)=\frac{1}{r}- \frac{1}{r_c} + (r - r_c) \frac{1}{r_c^2} - \frac{(r-r_c)^2}{r_c^3} .
348 % \end{equation}
349 % This function is an approximate electrostatic potential that has
350 % vanishing second derivatives at the cutoff radius, making it suitable
351 % for shifting the forces and torques of charge-dipole interactions.
352
353 The TSF potential for any multipole-multipole interaction can be
354 written
355 \begin{equation}
356 U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
357 \label{generic}
358 \end{equation}
359 where $f_n(r)$ is a shifted kernel that is appropriate for the order
360 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
361 charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
362 and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
363 quadrupole-quadrupole. To ensure smooth convergence of the energy,
364 force, and torques, a Taylor expansion with $n$ terms must be
365 performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
366
367 % To carry out the same procedure for a damped electrostatic kernel, we
368 % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369 % Many of the derivatives of the damped kernel are well known from
370 % Smith's early work on multipoles for the Ewald
371 % summation.\cite{Smith82,Smith98}
372
373 % Note that increasing the value of $n$ will add additional terms to the
374 % electrostatic potential, e.g., $f_2(r)$ includes orders up to
375 % $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
376 % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377 % f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
378 % for computing multipole energies, forces, and torques, and smooth
379 % cutoffs of these quantities can be guaranteed as long as the number of
380 % terms in the Taylor series exceeds the derivative order required.
381
382 For multipole-multipole interactions, following this procedure results
383 in separate radial functions for each of the distinct orientational
384 contributions to the potential, and ensures that the forces and
385 torques from each of these contributions will vanish at the cutoff
386 radius. For example, the direct dipole dot product
387 ($\mathbf{D}_{a}
388 \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
389 dot products:
390 \begin{equation}
391 U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
392 \mathbf{D}_{a} \cdot
393 \mathbf{D}_{b} \right) v_{21}(r) +
394 \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
395 \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
396 \end{equation}
397
398 For the Taylor shifted (TSF) method with the undamped kernel,
399 $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
400 \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
401 - \frac{6}{r r_c^2}$. In these functions, one can easily see the
402 connection to unmodified electrostatics as well as the smooth
403 transition to zero in both these functions as $r\rightarrow r_c$. The
404 electrostatic forces and torques acting on the central multipole due
405 to another site within the cutoff sphere are derived from
406 Eq.~\ref{generic}, accounting for the appropriate number of
407 derivatives. Complete energy, force, and torque expressions are
408 presented in the first paper in this series (Reference
409 \onlinecite{PaperI}).
410
411 \subsection{Gradient-shifted force (GSF)}
412
413 A second (and conceptually simpler) method involves shifting the
414 gradient of the raw Coulomb potential for each particular multipole
415 order. For example, the raw dipole-dipole potential energy may be
416 shifted smoothly by finding the gradient for two interacting dipoles
417 which have been projected onto the surface of the cutoff sphere
418 without changing their relative orientation,
419 \begin{equation}
420 U_{D_{a}D_{b}}(r) = U_{D_{a}D_{b}}(r) -
421 U_{D_{a}D_{b}}(r_c)
422 - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
423 \nabla U_{D_{a}D_{b}}(r_c).
424 \end{equation}
425 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
426 (although the signs are reversed for the dipole that has been
427 projected onto the cutoff sphere). In many ways, this simpler
428 approach is closer in spirit to the original shifted force method, in
429 that it projects a neutralizing multipole (and the resulting forces
430 from this multipole) onto a cutoff sphere. The resulting functional
431 forms for the potentials, forces, and torques turn out to be quite
432 similar in form to the Taylor-shifted approach, although the radial
433 contributions are significantly less perturbed by the gradient-shifted
434 approach than they are in the Taylor-shifted method.
435
436 For the gradient shifted (GSF) method with the undamped kernel,
437 $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
438 $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
439 Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
440 because the Taylor expansion retains only one term, they are
441 significantly less perturbed than the TSF functions.
442
443 In general, the gradient shifted potential between a central multipole
444 and any multipolar site inside the cutoff radius is given by,
445 \begin{equation}
446 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
447 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
448 \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
449 \label{generic2}
450 \end{equation}
451 where the sum describes a separate force-shifting that is applied to
452 each orientational contribution to the energy. In this expression,
453 $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
454 ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
455 represent the orientations the multipoles.
456
457 The third term converges more rapidly than the first two terms as a
458 function of radius, hence the contribution of the third term is very
459 small for large cutoff radii. The force and torque derived from
460 Eq. \ref{generic2} are consistent with the energy expression and
461 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
462 can be considered generalizations of the original DSF method for
463 higher order multipole interactions. GSF and TSF are also identical up
464 to the charge-dipole interaction but generate different expressions in
465 the energy, force and torque for higher order multipole-multipole
466 interactions. Complete energy, force, and torque expressions for the
467 GSF potential are presented in the first paper in this series
468 (Reference~\onlinecite{PaperI}).
469
470
471 \subsection{Shifted potential (SP) }
472 A discontinuous truncation of the electrostatic potential at the
473 cutoff sphere introduces a severe artifact (oscillation in the
474 electrostatic energy) even for molecules with the higher-order
475 multipoles.\cite{PaperI} We have also formulated an extension of the
476 Wolf approach for point multipoles by simply projecting the image
477 multipole onto the surface of the cutoff sphere, and including the
478 interactions with the central multipole and the image. This
479 effectively shifts the total potential to zero at the cutoff radius,
480 \begin{equation}
481 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
482 U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
483 \label{eq:SP}
484 \end{equation}
485 where the sum describes separate potential shifting that is done for
486 each orientational contribution to the energy (e.g. the direct dipole
487 product contribution is shifted {\it separately} from the
488 dipole-distance terms in dipole-dipole interactions). Note that this
489 is not a simple shifting of the total potential at $r_c$. Each radial
490 contribution is shifted separately. One consequence of this is that
491 multipoles that reorient after leaving the cutoff sphere can re-enter
492 the cutoff sphere without perturbing the total energy.
493
494 For the shifted potential (SP) method with the undamped kernel,
495 $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
496 \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
497 central multipole and other multipolar sites goes smoothly to zero as
498 $r \rightarrow r_c$. However, the force and torque obtained from the
499 shifted potential (SP) are discontinuous at $r_c$. MD simulations
500 will still experience energy drift while operating under the SP
501 potential, but it may be suitable for Monte Carlo approaches where the
502 configurational energy differences are the primary quantity of
503 interest.
504
505 \subsection{The Self Term}
506 In the TSF, GSF, and SP methods, a self-interaction is retained for
507 the central multipole interacting with its own image on the surface of
508 the cutoff sphere. This self interaction is nearly identical with the
509 self-terms that arise in the Ewald sum for multipoles. Complete
510 expressions for the self terms are presented in the first paper in
511 this series (Reference \onlinecite{PaperI}).
512
513
514 \section{\label{sec:methodology}Methodology}
515
516 To understand how the real-space multipole methods behave in computer
517 simulations, it is vital to test against established methods for
518 computing electrostatic interactions in periodic systems, and to
519 evaluate the size and sources of any errors that arise from the
520 real-space cutoffs. In the first paper of this series, we compared
521 the dipolar and quadrupolar energy expressions against analytic
522 expressions for ordered dipolar and quadrupolar
523 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
524 used the multipolar Ewald sum as a reference method for comparing
525 energies, forces, and torques for molecular models that mimic
526 disordered and ordered condensed-phase systems. The parameters used
527 in the test cases are given in table~\ref{tab:pars}.
528
529 \begin{table}
530 \label{tab:pars}
531 \caption{The parameters used in the systems used to evaluate the new
532 real-space methods. The most comprehensive test was a liquid
533 composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
534 ions). This test excercises all orders of the multipolar
535 interactions developed in the first paper.}
536 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
537 & \multicolumn{2}{c|}{LJ parameters} &
538 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
539 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
540 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
541 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
542 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
543 \AA\textsuperscript{2})} \\ \hline
544 Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
545 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
546 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
547 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
548 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
549 \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
550 \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
551 \end{tabularx}
552 \end{table}
553 The systems consist of pure multipolar solids (both dipole and
554 quadrupole), pure multipolar liquids (both dipole and quadrupole), a
555 fluid composed of sites containing both dipoles and quadrupoles
556 simultaneously, and a final test case that includes ions with point
557 charges in addition to the multipolar fluid. The solid-phase
558 parameters were chosen so that the systems can explore some
559 orientational freedom for the multipolar sites, while maintaining
560 relatively strict translational order. The SSDQ model used here is
561 not a particularly accurate water model, but it does test
562 dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
563 interactions at roughly the same magnitudes. The last test case, SSDQ
564 water with dissolved ions, exercises \textit{all} levels of the
565 multipole-multipole interactions we have derived so far and represents
566 the most complete test of the new methods.
567
568 In the following section, we present results for the total
569 electrostatic energy, as well as the electrostatic contributions to
570 the force and torque on each molecule. These quantities have been
571 computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
572 and have been compared with the values obtained from the multipolar
573 Ewald sum. In Monte Carlo (MC) simulations, the energy differences
574 between two configurations is the primary quantity that governs how
575 the simulation proceeds. These differences are the most important
576 indicators of the reliability of a method even if the absolute
577 energies are not exact. For each of the multipolar systems listed
578 above, we have compared the change in electrostatic potential energy
579 ($\Delta E$) between 250 statistically-independent configurations. In
580 molecular dynamics (MD) simulations, the forces and torques govern the
581 behavior of the simulation, so we also compute the electrostatic
582 contributions to the forces and torques.
583
584 \subsection{Implementation}
585 The real-space methods developed in the first paper in this series
586 have been implemented in our group's open source molecular simulation
587 program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
588 this work. The complementary error function can be a relatively slow
589 function on some processors, so all of the radial functions are
590 precomputed on a fine grid and are spline-interpolated to provide
591 values when required.
592
593 Using the same simulation code, we compare to a multipolar Ewald sum
594 with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
595 the Ewald sum is a re-implementation of the algorithm originally
596 proposed by Smith that does not use the particle mesh or smoothing
597 approximations.\cite{Smith82,Smith98} In all cases, the quantities
598 being compared are the electrostatic contributions to energies, force,
599 and torques. All other contributions to these quantities (i.e. from
600 Lennard-Jones interactions) are removed prior to the comparisons.
601
602 The convergence parameter ($\alpha$) also plays a role in the balance
603 of the real-space and reciprocal-space portions of the Ewald
604 calculation. Typical molecular mechanics packages set this to a value
605 that depends on the cutoff radius and a tolerance (typically less than
606 $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
607 associated with increasing accuracy at the expense of computational
608 time spent on the reciprocal-space portion of the
609 summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
610 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
611 Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
612
613 The real-space models have self-interactions that provide
614 contributions to the energies only. Although the self interaction is
615 a rapid calculation, we note that in systems with fluctuating charges
616 or point polarizabilities, the self-term is not static and must be
617 recomputed at each time step.
618
619 \subsection{Model systems}
620 To sample independent configurations of the multipolar crystals, body
621 centered cubic (bcc) crystals, which exhibit the minimum energy
622 structures for point dipoles, were generated using 3,456 molecules.
623 The multipoles were translationally locked in their respective crystal
624 sites for equilibration at a relatively low temperature (50K) so that
625 dipoles or quadrupoles could freely explore all accessible
626 orientations. The translational constraints were then removed, the
627 systems were re-equilibrated, and the crystals were simulated for an
628 additional 10 ps in the microcanonical (NVE) ensemble with an average
629 temperature of 50 K. The balance between moments of inertia and
630 particle mass were chosen to allow orientational sampling without
631 significant translational motion. Configurations were sampled at
632 equal time intervals in order to compare configurational energy
633 differences. The crystals were simulated far from the melting point
634 in order to avoid translational deformation away of the ideal lattice
635 geometry.
636
637 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
638 simulations, each system was created with 2,048 randomly-oriented
639 molecules. These were equilibrated at a temperature of 300K for 1 ns.
640 Each system was then simulated for 1 ns in the microcanonical (NVE)
641 ensemble. We collected 250 different configurations at equal time
642 intervals. For the liquid system that included ionic species, we
643 converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
644 \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
645 run under the same conditions for 1 ns. A total of 250 configurations
646 were collected. In the following comparisons of energies, forces, and
647 torques, the Lennard-Jones potentials were turned off and only the
648 purely electrostatic quantities were compared with the same values
649 obtained via the Ewald sum.
650
651 \subsection{Accuracy of Energy Differences, Forces and Torques}
652 The pairwise summation techniques (outlined above) were evaluated for
653 use in MC simulations by studying the energy differences between
654 different configurations. We took the Ewald-computed energy
655 difference between two conformations to be the correct behavior. An
656 ideal performance by one of the new methods would reproduce these
657 energy differences exactly. The configurational energies being used
658 here contain only contributions from electrostatic interactions.
659 Lennard-Jones interactions were omitted from the comparison as they
660 should be identical for all methods.
661
662 Since none of the real-space methods provide exact energy differences,
663 we used least square regressions analysis for the six different
664 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
665 with the multipolar Ewald reference method. Unitary results for both
666 the correlation (slope) and correlation coefficient for these
667 regressions indicate perfect agreement between the real-space method
668 and the multipolar Ewald sum.
669
670 Molecular systems were run long enough to explore independent
671 configurations and 250 configurations were recorded for comparison.
672 Each system provided 31,125 energy differences for a total of 186,750
673 data points. Similarly, the magnitudes of the forces and torques have
674 also been compared using least squares regression analysis. In the
675 forces and torques comparison, the magnitudes of the forces acting in
676 each molecule for each configuration were evaluated. For example, our
677 dipolar liquid simulation contains 2048 molecules and there are 250
678 different configurations for each system resulting in 3,072,000 data
679 points for comparison of forces and torques.
680
681 \subsection{Analysis of vector quantities}
682 Getting the magnitudes of the force and torque vectors correct is only
683 part of the issue for carrying out accurate molecular dynamics
684 simulations. Because the real space methods reweight the different
685 orientational contributions to the energies, it is also important to
686 understand how the methods impact the \textit{directionality} of the
687 force and torque vectors. Fisher developed a probablity density
688 function to analyse directional data sets,
689 \begin{equation}
690 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
691 \label{eq:pdf}
692 \end{equation}
693 where $\kappa$ measures directional dispersion of the data around the
694 mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
695 estimated as a reciprocal of the circular variance.\cite{Allen91} To
696 quantify the directional error, forces obtained from the Ewald sum
697 were taken as the mean (or correct) direction and the angle between
698 the forces obtained via the Ewald sum and the real-space methods were
699 evaluated,
700 \begin{equation}
701 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
702 \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
703 \end{equation}
704 The total angular displacement of the vectors was calculated as,
705 \begin{equation}
706 R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
707 \label{eq:displacement}
708 \end{equation}
709 where $N$ is number of force vectors. The circular variance is
710 defined as
711 \begin{equation}
712 \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
713 \end{equation}
714 The circular variance takes on values between from 0 to 1, with 0
715 indicating a perfect directional match between the Ewald force vectors
716 and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
717 correspond to higher values of $\kappa$, which indicates tighter
718 clustering of the real-space force vectors around the Ewald forces.
719
720 A similar analysis was carried out for the electrostatic contribution
721 to the molecular torques as well as forces.
722
723 \subsection{Energy conservation}
724 To test conservation the energy for the methods, the mixed molecular
725 system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
726 ions was run for 1 ns in the microcanonical ensemble at an average
727 temperature of 300K. Each of the different electrostatic methods
728 (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
729 damping values. The molecular system was started with same initial
730 positions and velocities for all cutoff methods. The energy drift
731 ($\delta E_1$) and standard deviation of the energy about the slope
732 ($\delta E_0$) were evaluated from the total energy of the system as a
733 function of time. Although both measures are valuable at
734 investigating new methods for molecular dynamics, a useful interaction
735 model must allow for long simulation times with minimal energy drift.
736
737 \section{\label{sec:result}RESULTS}
738 \subsection{Configurational energy differences}
739 %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.
740 %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.
741 %\begin{figure}[h!]
742 % \centering
743 % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
744 % \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. }
745 % \label{fig:rcutConvergence_dipolarLiquid}
746 % \end{figure}
747 %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.
748 %\begin{figure}
749 % \centering
750 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
751 % \label{fig:barGraph1}
752 % \end{figure}
753 % \begin{figure}
754 % \centering
755 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
756 % \caption{}
757
758 % \label{fig:barGraph2}
759 % \end{figure}
760 %The correlation coefficient ($R^2$) and slope of the linear
761 %regression plots for the energy differences for all six different
762 %molecular systems is shown in figure 4a and 4b.The plot shows that
763 %the correlation coefficient improves for the SP cutoff method as
764 %compared to the undamped hard cutoff method in the case of SSDQC,
765 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
766 %crystal and liquid, the correlation coefficient is almost unchanged
767 %and close to 1. The correlation coefficient is smallest (0.696276
768 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
769 %charge-charge and charge-multipole interactions. Since the
770 %charge-charge and charge-multipole interaction is long ranged, there
771 %is huge deviation of correlation coefficient from 1. Similarly, the
772 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
773 %compared to interactions in the other multipolar systems, thus the
774 %correlation coefficient very close to 1 even for hard cutoff
775 %method. The idea of placing image multipole on the surface of the
776 %cutoff sphere improves the correlation coefficient and makes it close
777 %to 1 for all types of multipolar systems. Similarly the slope is
778 %hugely deviated from the correct value for the lower order
779 %multipole-multipole interaction and slightly deviated for higher
780 %order multipole – multipole interaction. The SP method improves both
781 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
782 %dipolar systems. The Slope is found to be deviated more in dipolar
783 %crystal as compared to liquid which is associated with the large
784 %fluctuation in the electrostatic energy in crystal. The GSF also
785 %produced better values of correlation coefficient and slope with the
786 %proper selection of the damping alpha (Interested reader can consult
787 %accompanying supporting material). The TSF method gives good value of
788 %correlation coefficient for the dipolar crystal, dipolar liquid,
789 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
790 %regression slopes are significantly deviated.
791
792 \begin{figure}
793 \centering
794 \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
795 \caption{Statistical analysis of the quality of configurational
796 energy differences for the real-space electrostatic methods
797 compared with the reference Ewald sum. Results with a value equal
798 to 1 (dashed line) indicate $\Delta E$ values indistinguishable
799 from those obtained using the multipolar Ewald sum. Different
800 values of the cutoff radius are indicated with different symbols
801 (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
802 triangles).}
803 \label{fig:slopeCorr_energy}
804 \end{figure}
805
806 The combined correlation coefficient and slope for all six systems is
807 shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
808 reproduce the Ewald configurational energy differences with remarkable
809 fidelity. Undamped hard cutoffs introduce a significant amount of
810 random scatter in the energy differences which is apparent in the
811 reduced value of the correlation coefficient for this method. This
812 can be easily understood as configurations which exhibit small
813 traversals of a few dipoles or quadrupoles out of the cutoff sphere
814 will see large energy jumps when hard cutoffs are used. The
815 orientations of the multipoles (particularly in the ordered crystals)
816 mean that these energy jumps can go in either direction, producing a
817 significant amount of random scatter, but no systematic error.
818
819 The TSF method produces energy differences that are highly correlated
820 with the Ewald results, but it also introduces a significant
821 systematic bias in the values of the energies, particularly for
822 smaller cutoff values. The TSF method alters the distance dependence
823 of different orientational contributions to the energy in a
824 non-uniform way, so the size of the cutoff sphere can have a large
825 effect, particularly for the crystalline systems.
826
827 Both the SP and GSF methods appear to reproduce the Ewald results with
828 excellent fidelity, particularly for moderate damping ($\alpha =
829 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
830 12$\AA). With the exception of the undamped hard cutoff, and the TSF
831 method with short cutoffs, all of the methods would be appropriate for
832 use in Monte Carlo simulations.
833
834 \subsection{Magnitude of the force and torque vectors}
835
836 The comparisons of the magnitudes of the forces and torques for the
837 data accumulated from all six systems are shown in Figures
838 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
839 correlation and slope for the forces agree well with the Ewald sum
840 even for the hard cutoffs.
841
842 For systems of molecules with only multipolar interactions, the pair
843 energy contributions are quite short ranged. Moreover, the force
844 decays more rapidly than the electrostatic energy, hence the hard
845 cutoff method can also produce reasonable agreement for this quantity.
846 Although the pure cutoff gives reasonably good electrostatic forces
847 for pairs of molecules included within each other's cutoff spheres,
848 the discontinuity in the force at the cutoff radius can potentially
849 cause energy conservation problems as molecules enter and leave the
850 cutoff spheres. This is discussed in detail in section
851 \ref{sec:conservation}.
852
853 The two shifted-force methods (GSF and TSF) exhibit a small amount of
854 systematic variation and scatter compared with the Ewald forces. The
855 shifted-force models intentionally perturb the forces between pairs of
856 molecules inside each other's cutoff spheres in order to correct the
857 energy conservation issues, and this perturbation is evident in the
858 statistics accumulated for the molecular forces. The GSF
859 perturbations are minimal, particularly for moderate damping and
860 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
861 reasonable agreement in the correlation coefficient but again the
862 systematic error in the forces is concerning if replication of Ewald
863 forces is desired.
864
865 \begin{figure}
866 \centering
867 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
868 \caption{Statistical analysis of the quality of the force vector
869 magnitudes for the real-space electrostatic methods compared with
870 the reference Ewald sum. Results with a value equal to 1 (dashed
871 line) indicate force magnitude values indistinguishable from those
872 obtained using the multipolar Ewald sum. Different values of the
873 cutoff radius are indicated with different symbols (9\AA\ =
874 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
875 \label{fig:slopeCorr_force}
876 \end{figure}
877
878
879 \begin{figure}
880 \centering
881 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
882 \caption{Statistical analysis of the quality of the torque vector
883 magnitudes for the real-space electrostatic methods compared with
884 the reference Ewald sum. Results with a value equal to 1 (dashed
885 line) indicate force magnitude values indistinguishable from those
886 obtained using the multipolar Ewald sum. Different values of the
887 cutoff radius are indicated with different symbols (9\AA\ =
888 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
889 \label{fig:slopeCorr_torque}
890 \end{figure}
891
892 The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
893 significantly influenced by the choice of real-space method. The
894 torque expressions have the same distance dependence as the energies,
895 which are naturally longer-ranged expressions than the inter-site
896 forces. Torques are also quite sensitive to orientations of
897 neighboring molecules, even those that are near the cutoff distance.
898
899 The results shows that the torque from the hard cutoff method
900 reproduces the torques in quite good agreement with the Ewald sum.
901 The other real-space methods can cause some deviations, but excellent
902 agreement with the Ewald sum torques is recovered at moderate values
903 of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
904 radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
905 in the slope when compared with the Ewald torques even for larger
906 cutoff radii. It appears that the severity of the perturbations in
907 the TSF method are most in evidence for the torques.
908
909 \subsection{Directionality of the force and torque vectors}
910
911 The accurate evaluation of force and torque directions is just as
912 important for molecular dynamics simulations as the magnitudes of
913 these quantities. Force and torque vectors for all six systems were
914 analyzed using Fisher statistics, and the quality of the vector
915 directionality is shown in terms of circular variance
916 ($\mathrm{Var}(\theta)$) in figure
917 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
918 from the new real-space methods exhibit nearly-ideal Fisher probability
919 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
920 exhibit the best vectorial agreement with the Ewald sum. The force and
921 torque vectors from GSF method also show good agreement with the Ewald
922 method, which can also be systematically improved by using moderate
923 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
924 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
925 to a distribution with 95\% of force vectors within $6.37^\circ$ of
926 the corresponding Ewald forces. The TSF method produces the poorest
927 agreement with the Ewald force directions.
928
929 Torques are again more perturbed than the forces by the new real-space
930 methods, but even here the variance is reasonably small. For the same
931 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
932 the circular variance was 0.01415, corresponds to a distribution which
933 has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
934 results. Again, the direction of the force and torque vectors can be
935 systematically improved by varying $\alpha$ and $r_c$.
936
937 \begin{figure}
938 \centering
939 \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
940 \caption{The circular variance of the direction of the force and
941 torque vectors obtained from the real-space methods around the
942 reference Ewald vectors. A variance equal to 0 (dashed line)
943 indicates direction of the force or torque vectors are
944 indistinguishable from those obtained from the Ewald sum. Here
945 different symbols represent different values of the cutoff radius
946 (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
947 \label{fig:slopeCorr_circularVariance}
948 \end{figure}
949
950 \subsection{Energy conservation\label{sec:conservation}}
951
952 We have tested the conservation of energy one can expect to see with
953 the new real-space methods using the SSDQ water model with a small
954 fraction of solvated ions. This is a test system which exercises all
955 orders of multipole-multipole interactions derived in the first paper
956 in this series and provides the most comprehensive test of the new
957 methods. A liquid-phase system was created with 2000 water molecules
958 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
959 temperature of 300K. After equilibration in the canonical (NVT)
960 ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
961 was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
962 Hard, SP, GSF, and TSF methods with a cutoff radius of 12\AA. The
963 value of the damping coefficient was also varied from the undamped
964 case ($\alpha = 0$) to a heavily damped case ($\alpha = 0.3$
965 \AA$^{-1}$) for all of the real space methods. A sample was also run
966 using the multipolar Ewald sum with the same real-space cutoff.
967
968 In figure~\ref{fig:energyDrift} we show the both the linear drift in
969 energy over time, $\delta E_1$, and the standard deviation of energy
970 fluctuations around this drift $\delta E_0$. Both of the
971 shifted-force methods (GSF and TSF) provide excellent energy
972 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
973 while the hard cutoff is essentially unusable for molecular dynamics.
974 SP provides some benefit over the hard cutoff because the energetic
975 jumps that happen as particles leave and enter the cutoff sphere are
976 somewhat reduced, but like the Wolf method for charges, the SP method
977 would not be as useful for molecular dynamics as either of the
978 shifted-force methods.
979
980 We note that for all tested values of the cutoff radius, the new
981 real-space methods can provide better energy conservation behavior
982 than the multipolar Ewald sum, even when utilizing a relatively large
983 $k$-space cutoff values.
984
985 \begin{figure}
986 \centering
987 \includegraphics[width=\textwidth]{newDrift_12.eps}
988 \label{fig:energyDrift}
989 \caption{Analysis of the energy conservation of the real-space
990 methods. $\delta \mathrm{E}_1$ is the linear drift in energy over
991 time (in kcal / mol / particle / ns) and $\delta \mathrm{E}_0$ is
992 the standard deviation of energy fluctuations around this drift (in
993 kcal / mol / particle). Points that appear below the dashed grey
994 (Ewald) lines exhibit better energy conservation than commonly-used
995 parameters for Ewald-based electrostatics. All simulations were of
996 a 2000-molecule simulation of SSDQ water with 48 ionic charges at
997 300 K starting from the same initial configuration. All runs
998 utilized the same real-space cutoff, $r_c = 12$\AA.}
999 \end{figure}
1000
1001 \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
1002 The most important test of the modified interaction potentials is the
1003 fidelity with which they can reproduce structural features and
1004 dynamical properties in a liquid. One commonly-utilized measure of
1005 structural ordering is the pair distribution function, $g(r)$, which
1006 measures local density deviations in relation to the bulk density. In
1007 the electrostatic approaches studied here, the short-range repulsion
1008 from the Lennard-Jones potential is identical for the various
1009 electrostatic methods, and since short range repulsion determines much
1010 of the local liquid ordering, one would not expect to see many
1011 differences in $g(r)$. Indeed, the pair distributions are essentially
1012 identical for all of the electrostatic methods studied (for each of
1013 the different systems under investigation). An example of this
1014 agreement for the SSDQ water/ion system is shown in
1015 Fig. \ref{fig:gofr}.
1016
1017 \begin{figure}
1018 \centering
1019 \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
1020 \label{fig:gofr}
1021 \caption{The pair distribution functions, $g(r)$, for the SSDQ
1022 water/ion system obtained using the different real-space methods are
1023 essentially identical with the result from the Ewald
1024 treatment.}
1025 \end{figure}
1026
1027 There is a very slight overstructuring of the first solvation shell
1028 when using when using TSF at lower values of the damping coefficient
1029 ($\alpha \le 0.1$) or when using undamped GSF. With moderate damping,
1030 GSF and SP produce pair distributions that are identical (within
1031 numerical noise) to their Ewald counterparts.
1032
1033 A structural property that is a more demanding test of modified
1034 electrostatics is the mean value of the electrostatic energy $\langle
1035 U_\mathrm{elect} \rangle / N$ which is obtained by sampling the
1036 liquid-state configurations experienced by a liquid evolving entirely
1037 under the influence of each of the methods. In table \ref{tab:Props}
1038 we demonstrate how $\langle U_\mathrm{elect} \rangle / N$ varies with
1039 the damping parameter, $\alpha$, for each of the methods.
1040
1041 As in the crystals studied in the first paper, damping is important
1042 for converging the mean electrostatic energy values, particularly for
1043 the two shifted force methods (GSF and TSF). A value of $\alpha
1044 \approx 0.2$ \AA$^{-1}$ is sufficient to converge the SP and GSF
1045 energies with a cutoff of 12 \AA, while shorter cutoffs require more
1046 dramatic damping ($\alpha \approx 0.3$ \AA$^{-1}$ for $r_c = 9$ \AA).
1047 Overdamping the real-space electrostatic methods occurs with $\alpha >
1048 0.4$, causing the estimate of the energy to drop below the Ewald
1049 results.
1050
1051 These ``optimal'' values of the damping coefficient are slightly
1052 larger than what were observed for DSF electrostatics for purely
1053 point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
1054 $r_c = 12$\AA appears to be an excellent compromise for mixed charge
1055 multipole systems.
1056
1057 To test the fidelity of the electrostatic methods at reproducing
1058 dynamics in a multipolar liquid, it is also useful to look at
1059 transport properties, particularly the diffusion constant,
1060 \begin{equation}
1061 D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
1062 \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
1063 \label{eq:diff}
1064 \end{equation}
1065 which measures long-time behavior and is sensitive to the forces on
1066 the multipoles. For the soft dipolar fluid and the SSDQ liquid
1067 systems, the self-diffusion constants (D) were calculated from linear
1068 fits to the long-time portion of the mean square displacement,
1069 $\langle r^{2}(t) \rangle$.\cite{Allen87}
1070
1071 In addition to translational diffusion, orientational relaxation times
1072 were calculated for comparisons with the Ewald simulations and with
1073 experiments. These values were determined from the same 1~ns
1074 microcanonical trajectories used for translational diffusion by
1075 calculating the orientational time correlation function,
1076 \begin{equation}
1077 C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1078 \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1079 \label{eq:OrientCorr}
1080 \end{equation}
1081 where $P_l$ is the Legendre polynomial of order $l$ and
1082 $\hat{\mathbf{A}}_\gamma$ is the space-frame unit vector for body axis
1083 $\gamma$ on a molecule.. Th body-fixed reference frame used for our
1084 models has the $z$-axis running along the dipoles, and for the SSDQ
1085 water model, the $y$-axis connects the two implied hydrogen atom
1086 positions. From the orientation autocorrelation functions, we can
1087 obtain time constants for rotational relaxation either by fitting an
1088 exponential function or by integrating the entire correlation
1089 function. In a good water model, these decay times would be
1090 comparable to water orientational relaxation times from nuclear
1091 magnetic resonance (NMR). The relaxation constant obtained from
1092 $C_2^y(t)$ is normally of experimental interest because it describes
1093 the relaxation of the principle axis connecting the hydrogen
1094 atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1095 of the dipole-dipole relaxation from a proton NMR signal and should
1096 provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1097
1098 Results for the diffusion constants and orientational relaxation times
1099 are shown in figure \ref{tab:Props}. From this data, it is apparent
1100 that the values for both $D$ and $\tau_2$ using the Ewald sum are
1101 reproduced with reasonable fidelity by the GSF method.
1102
1103 The $\tau_2$ results in \ref{tab:Props} show a much greater difference
1104 between the real-space and the Ewald results.
1105
1106 \begin{table}
1107 \label{tab:Props}
1108 \caption{Comparison of the structural and dynamic properties for the
1109 soft dipolar liquid test for all of the real-space methods.}
1110 \begin{tabular}{l|c|cccc|cccc|cccc}
1111 & Ewald & \multicolumn{4}{c|}{SP} & \multicolumn{4}{c|}{GSF} & \multicolumn{4}{c|}{TSF} \\
1112 $\alpha$ (\AA$^{-1}$) & &
1113 0.0 & 0.1 & 0.2 & 0.3 &
1114 0.0 & 0.1 & 0.2 & 0.3 &
1115 0.0 & 0.1 & 0.2 & 0.3 \\ \cline{2-6}\cline{6-10}\cline{10-14}
1116
1117 $\langle U_\mathrm{elect} \rangle /N$ &&&&&&&&&&&&&\\
1118 D ($10^{-4}~\mathrm{cm}^2/\mathrm{s}$)&
1119 470.2(6) &
1120 416.6(5) &
1121 379.6(5) &
1122 438.6(5) &
1123 476.0(6) &
1124 412.8(5) &
1125 421.1(5) &
1126 400.5(5) &
1127 437.5(6) &
1128 434.6(5) &
1129 411.4(5) &
1130 545.3(7) &
1131 459.6(6) \\
1132 $\tau_2$ (fs) &
1133 1.136 &
1134 1.041 &
1135 1.064 &
1136 1.109 &
1137 1.211 &
1138 1.119 &
1139 1.039 &
1140 1.058 &
1141 1.21 &
1142 1.15 &
1143 1.172 &
1144 1.153 &
1145 1.125 \\
1146 \end{tabular}
1147 \end{table}
1148
1149
1150 \section{CONCLUSION}
1151 In the first paper in this series, we generalized the
1152 charge-neutralized electrostatic energy originally developed by Wolf
1153 \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1154 up to quadrupolar order. The SP method is essentially a
1155 multipole-capable version of the Wolf model. The SP method for
1156 multipoles provides excellent agreement with Ewald-derived energies,
1157 forces and torques, and is suitable for Monte Carlo simulations,
1158 although the forces and torques retain discontinuities at the cutoff
1159 distance that prevents its use in molecular dynamics.
1160
1161 We also developed two natural extensions of the damped shifted-force
1162 (DSF) model originally proposed by Fennel and
1163 Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1164 smooth truncation of energies, forces, and torques at the real-space
1165 cutoff, and both converge to DSF electrostatics for point-charge
1166 interactions. The TSF model is based on a high-order truncated Taylor
1167 expansion which can be relatively perturbative inside the cutoff
1168 sphere. The GSF model takes the gradient from an images of the
1169 interacting multipole that has been projected onto the cutoff sphere
1170 to derive shifted force and torque expressions, and is a significantly
1171 more gentle approach.
1172
1173 Of the two newly-developed shifted force models, the GSF method
1174 produced quantitative agreement with Ewald energy, force, and torques.
1175 It also performs well in conserving energy in MD simulations. The
1176 Taylor-shifted (TSF) model provides smooth dynamics, but these take
1177 place on a potential energy surface that is significantly perturbed
1178 from Ewald-based electrostatics.
1179
1180 % The direct truncation of any electrostatic potential energy without
1181 % multipole neutralization creates large fluctuations in molecular
1182 % simulations. This fluctuation in the energy is very large for the case
1183 % of crystal because of long range of multipole ordering (Refer paper
1184 % I).\cite{PaperI} This is also significant in the case of the liquid
1185 % because of the local multipole ordering in the molecules. If the net
1186 % multipole within cutoff radius neutralized within cutoff sphere by
1187 % placing image multiples on the surface of the sphere, this fluctuation
1188 % in the energy reduced significantly. Also, the multipole
1189 % neutralization in the generalized SP method showed very good agreement
1190 % with the Ewald as compared to direct truncation for the evaluation of
1191 % the $\triangle E$ between the configurations. In MD simulations, the
1192 % energy conservation is very important. The conservation of the total
1193 % energy can be ensured by i) enforcing the smooth truncation of the
1194 % energy, force and torque in the cutoff radius and ii) making the
1195 % energy, force and torque consistent with each other. The GSF and TSF
1196 % methods ensure the consistency and smooth truncation of the energy,
1197 % force and torque at the cutoff radius, as a result show very good
1198 % total energy conservation. But the TSF method does not show good
1199 % agreement in the absolute value of the electrostatic energy, force and
1200 % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1201 % energy and torque accurately and also conserved energy.
1202
1203 The only cases we have found where the new GSF and SP real-space
1204 methods can be problematic are those which retain a bulk dipole moment
1205 at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1206 materials, uniform weighting of the orientational contributions can be
1207 important for converging the total energy. In these cases, the
1208 damping function which causes the non-uniform weighting can be
1209 replaced by the bare electrostatic kernel, and the energies return to
1210 the expected converged values.
1211
1212 Based on the results of this work, the GSF method is a suitable and
1213 efficient replacement for the Ewald sum for evaluating electrostatic
1214 interactions in MD simulations. Both methods retain excellent
1215 fidelity to the Ewald energies, forces and torques. Additionally, the
1216 energy drift and fluctuations from the GSF electrostatics are better
1217 than a multipolar Ewald sum for finite-sized reciprocal spaces.
1218 Because they use real-space cutoffs with moderate cutoff radii, the
1219 GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1220 increases. Additionally, they can be made extremely efficient using
1221 spline interpolations of the radial functions. They require no
1222 Fourier transforms or $k$-space sums, and guarantee the smooth
1223 handling of energies, forces, and torques as multipoles cross the
1224 real-space cutoff boundary.
1225
1226 \begin{acknowledgments}
1227 JDG acknowledges helpful discussions with Christopher
1228 Fennell. Support for this project was provided by the National
1229 Science Foundation under grant CHE-1362211. Computational time was
1230 provided by the Center for Research Computing (CRC) at the
1231 University of Notre Dame.
1232 \end{acknowledgments}
1233
1234 %\bibliographystyle{aip}
1235 \newpage
1236 \bibliography{references}
1237 \end{document}
1238
1239 %
1240 % ****** End of file aipsamp.tex ******