ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4175
Committed: Fri Jun 6 16:01:36 2014 UTC (10 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 59864 byte(s)
Log Message:
Latest edits

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{mathptm}
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 alternatives to the Ewald
51 Sum. II. Comparison of Methods} % Force line breaks with \\
52
53 \author{Madan Lamichhane}
54 \affiliation{Department of Physics, University
55 of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
56
57 \author{Kathie E. Newman}
58 \affiliation{Department of Physics, University
59 of Notre Dame, Notre Dame, IN 46556}
60
61 \author{J. Daniel Gezelter}%
62 \email{gezelter@nd.edu.}
63 \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
64 }%
65
66 \date{\today}% It is always \today, today,
67 % but any date may be explicitly specified
68
69 \begin{abstract}
70 We have tested the real-space shifted potential (SP),
71 gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
72 for multipoles that were developed in the first paper in this series
73 against a reference method. The tests were carried out in a variety
74 of condensed-phase environments which were designed to test all
75 levels of the multipole-multipole interactions. Comparisons of the
76 energy differences between configurations, molecular forces, and
77 torques were used to analyze how well the real-space models perform
78 relative to the more computationally expensive Ewald sum. We have
79 also investigated the energy conservation properties of the new
80 methods in molecular dynamics simulations using all of these
81 methods. The SP method shows excellent agreement with
82 configurational energy differences, forces, and torques, and would
83 be suitable for use in Monte Carlo calculations. Of the two new
84 shifted-force methods, the GSF approach shows the best agreement
85 with Ewald-derived energies, forces, and torques and exhibits energy
86 conservation properties that make it an excellent choice for
87 efficiently computing electrostatic interactions in molecular
88 dynamics simulations.
89 \end{abstract}
90
91 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92 % Classification Scheme.
93 \keywords{Electrostatics, Multipoles, Real-space}
94
95 \maketitle
96
97
98 \section{\label{sec:intro}Introduction}
99 Computing the interactions between electrostatic sites is one of the
100 most expensive aspects of molecular simulations, which is why there
101 have been significant efforts to develop practical, efficient and
102 convergent methods for handling these interactions. Ewald's method is
103 perhaps the best known and most accurate method for evaluating
104 energies, forces, and torques in explicitly-periodic simulation
105 cells. In this approach, the conditionally convergent electrostatic
106 energy is converted into two absolutely convergent contributions, one
107 which is carried out in real space with a cutoff radius, and one in
108 reciprocal space.\cite{Clarke:1986eu,Woodcock75}
109
110 When carried out as originally formulated, the reciprocal-space
111 portion of the Ewald sum exhibits relatively poor computational
112 scaling, making it prohibitive for large systems. By utilizing
113 particle meshes and three dimensional fast Fourier transforms (FFT),
114 the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
115 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
116 the computational cost from $O(N^2)$ down to $O(N \log
117 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
118
119 Because of the artificial periodicity required for the Ewald sum, the
120 method may require modification to compute interactions for
121 interfacial molecular systems such as membranes and liquid-vapor
122 interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
123 To simulate interfacial systems, Parry’s extension of the 3D Ewald sum
124 is appropriate for slab geometries.\cite{Parry:1975if} The inherent
125 periodicity in the Ewald’s method can also be problematic for
126 interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
127 methods that were developed to handle two-dimensional (2D)
128 electrostatic interactions in interfacial systems have not had similar
129 particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
131
132 \subsection{Real-space methods}
133 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
134 method for calculating electrostatic interactions between point
135 charges. They argued that the effective Coulomb interaction in
136 condensed systems is actually short ranged.\cite{Wolf92,Wolf95}. For
137 an ordered lattice (e.g. when computing the Madelung constant of an
138 ionic solid), the material can be considered as a set of ions
139 interacting with neutral dipolar or quadrupolar ``molecules'' giving
140 an effective distance dependence for the electrostatic interactions of
141 $r^{-5}$ (see figure \ref{fig:NaCl}. For this reason, careful
142 applications of Wolf's method are able to obtain accurate estimates of
143 Madelung constants using relatively short cutoff radii. Recently,
144 Fukuda used neutralization of the higher order moments for the
145 calculation of the electrostatic interaction of the point charges
146 system.\cite{Fukuda:2013sf}
147
148 \begin{figure}[h!]
149 \centering
150 \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
151 \caption{Top: NaCl crystal showing how spherical truncation can
152 breaking effective charge ordering, and how complete \ce{(NaCl)4}
153 molecules interact with the central ion. Bottom: A dipolar
154 crystal exhibiting similar behavior and illustrating how the
155 effective dipole-octupole interactions can be disrupted by
156 spherical truncation.}
157 \label{fig:NaCl}
158 \end{figure}
159
160 The direct truncation of interactions at a cutoff radius creates
161 truncation defects. Wolf \textit{et al.} further argued that
162 truncation errors are due to net charge remaining inside the cutoff
163 sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
164 placing an image charge on the surface of the cutoff sphere for every
165 real charge inside the cutoff. These charges are present for the
166 evaluation of both the pair interaction energy and the force, although
167 the force expression maintained a discontinuity at the cutoff sphere.
168 In the original Wolf formulation, the total energy for the charge and
169 image were not equal to the integral of their force expression, and as
170 a result, the total energy would not be conserved in molecular
171 dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
172 Fennel and Gezelter later proposed shifted force variants of the Wolf
173 method with commensurate force and energy expressions that do not
174 exhibit this problem.\cite{Fennell:2006lq} Related real-space
175 methods were also proposed by Chen \textit{et
176 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
177 and by Wu and Brooks.\cite{Wu:044107}
178
179 Considering the interaction of one central ion in an ionic crystal
180 with a portion of the crystal at some distance, the effective Columbic
181 potential is found to be decreasing as $r^{-5}$. If one views the
182 \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
183 \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
184 rapidly to the Madelung energy than the dipolar
185 approximation.\cite{Wolf92} To find the correct Madelung constant,
186 Lacman suggested that the NaCl structure could be constructed in a way
187 that the finite crystal terminates with complete \ce{(NaCl)4}
188 molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
189 by opposite charges. Similarly for each pair of charges, there is an
190 opposite pair of charge adjacent to it. The central ion sees what is
191 effectively a set of octupoles at large distances. These facts suggest
192 that the Madelung constants are relatively short ranged for perfect
193 ionic crystals.\cite{Wolf:1999dn}
194
195 One can make a similar argument for crystals of point multipoles. The
196 Luttinger and Tisza treatment of energy constants for dipolar lattices
197 utilizes 24 basis vectors that contain dipoles at the eight corners of
198 a unit cube. Only three of these basis vectors, $X_1, Y_1,
199 \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
200 zero net dipole and retain contributions only from higher order
201 multipoles. 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:NaCl}), causing significant
212 swings in the electrostatic energy as the cutoff radius is increased
213 (or as individual ions move back and forth across the boundary). This
214 is why the image charges were necessary for the Wolf sum to exhibit
215 rapid convergence. Similarly, the real-space truncation of point
216 multipole interactions breaks higher order multipole arrangements, and
217 image multipoles are required for real-space treatments of
218 electrostatic energies.
219
220 % Because of this reason, although the nature of electrostatic
221 % interaction short ranged, the hard cutoff sphere creates very large
222 % fluctuation in the electrostatic energy for the perfect crystal. In
223 % addition, the charge neutralized potential proposed by Wolf et
224 % al. converged to correct Madelung constant but still holds oscillation
225 % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
226 % oscillation in the energy around its fully converged value can be due
227 % to the non-neutralized value of the higher order moments within the
228 % cutoff sphere.
229
230 The forces and torques acting on atomic sites are the fundamental
231 factors driving dynamics in molecular simulations. Fennell and
232 Gezelter proposed the damped shifted force (DSF) energy kernel to
233 obtain consistent energies and forces on the atoms within the cutoff
234 sphere. Both the energy and the force go smoothly to zero as an atom
235 aproaches the cutoff radius. The comparisons of the accuracy these
236 quantities between the DSF kernel and SPME was surprisingly
237 good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
238 calculating electrostatic interactions in molecular systems with
239 relatively uniform charge
240 densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
241
242 \subsection{The damping function}
243 The damping function used in our research has been discussed in detail
244 in the first paper of this series.\cite{PaperI} The radial kernel
245 $1/r$ for the interactions between point charges can be replaced by
246 the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
247 accelerate the rate of convergence, where $\alpha$ is a damping
248 parameter with units of inverse distance. Altering the value of
249 $\alpha$ is equivalent to changing the width of Gaussian charge
250 distributions that replace each point charge -- Gaussian overlap
251 integrals yield complementary error functions when truncated at a
252 finite distance.
253
254 By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
255 cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
256 agreement with SPME for the interaction energies, forces and torques
257 for charge-charge interactions.\cite{Fennell:2006lq}
258
259 \subsection{Point multipoles in molecular modeling}
260 Coarse-graining approaches which treat entire molecular subsystems as
261 a single rigid body are now widely used. A common feature of many
262 coarse-graining approaches is simplification of the electrostatic
263 interactions between bodies so that fewer site-site interactions are
264 required to compute configurational energies. Many coarse-grained
265 molecular structures would normally consist of equal positive and
266 negative charges, and rather than use multiple site-site interactions,
267 the interaction between higher order multipoles can also be used to
268 evaluate a single molecule-molecule
269 interaction.\cite{Ren06,Essex10,Essex11}
270
271 Because electrons in a molecule are not localized at specific points,
272 the assignment of partial charges to atomic centers is a relatively
273 rough approximation. Atomic sites can also be assigned point
274 multipoles and polarizabilities to increase the accuracy of the
275 molecular model. Recently, water has been modeled with point
276 multipoles up to octupolar
277 order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
278 multipoles up to quadrupolar order have also been coupled with point
279 polarizabilities in the high-quality AMOEBA and iAMOEBA water
280 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}. But
281 using point multipole with the real space truncation without
282 accounting for multipolar neutrality will create energy conservation
283 issues in molecular dynamics (MD) simulations.
284
285 In this paper we test a set of real-space methods that were developed
286 for point multipolar interactions. These methods extend the damped
287 shifted force (DSF) and Wolf methods originally developed for
288 charge-charge interactions and generalize them for higher order
289 multipoles. The detailed mathematical development of these methods has
290 been presented in the first paper in this series, while this work
291 covers the testing the energies, forces, torques, and energy
292 conservation properties of the methods in realistic simulation
293 environments. In all cases, the methods are compared with the
294 reference method, a full multipolar Ewald treatment.
295
296
297 %\subsection{Conservation of total energy }
298 %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.
299
300 \section{\label{sec:method}Review of Methods}
301 Any real-space electrostatic method that is suitable for MD
302 simulations should have the electrostatic energy, forces and torques
303 between two sites go smoothly to zero as the distance between the
304 sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
305 this continuity at the cutoff is essential for energy conservation in
306 MD simulations. The mathematical details of the shifted potential
307 (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
308 methods have been discussed in detail in the previous paper in this
309 series.\cite{PaperI} Here we briefly review the new methods and
310 describe their essential features.
311
312 \subsection{Taylor-shifted force (TSF)}
313
314 The electrostatic potential energy between point multipoles can be
315 expressed as the product of two multipole operators and a Coulombic
316 kernel,
317 \begin{equation}
318 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
319 \end{equation}
320 where the multipole operator for site $\bf a$,
321 \begin{equation}
322 \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
323 + Q_{{\bf a}\alpha\beta}
324 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
325 \end{equation}
326 is expressed in terms of the point charge, $C_{\bf a}$, dipole,
327 $D_{{\bf a}\alpha}$, and quadrupole, $Q_{{\bf a}\alpha\beta}$, for
328 object $\bf a$. Note that in this work, we use the primitive
329 quadrupole tensor, $Q_{a\alpha,\beta}=\frac{1}{2}\sum_{k\;in\;a}q_k
330 r_{k\alpha}r_{k\beta}$ to represent point quadrupoles on a site.
331
332 Interactions between multipoles can be expressed as higher derivatives
333 of the bare Coulomb potential, so one way of ensuring that the forces
334 and torques vanish at the cutoff distance is to include a larger
335 number of terms in the truncated Taylor expansion, e.g.,
336 %
337 \begin{equation}
338 f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
339 \end{equation}
340 %
341 The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
342 Thus, for $f(r)=1/r$, we find
343 %
344 \begin{equation}
345 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} .
346 \end{equation}
347 This function is an approximate electrostatic potential that has
348 vanishing second derivatives at the cutoff radius, making it suitable
349 for shifting the forces and torques of charge-dipole interactions.
350
351 In general, the TSF potential for any multipole-multipole interaction
352 can be written
353 \begin{equation}
354 U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
355 \label{generic}
356 \end{equation}
357 with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
358 charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
359 $n=4$ for quadrupole-quadrupole. To ensure smooth convergence of the
360 energy, force, and torques, the required number of terms from Taylor
361 series expansion in $f_n(r)$ must be performed for different
362 multipole-multipole interactions.
363
364 To carry out the same procedure for a damped electrostatic kernel, we
365 replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
366 Many of the derivatives of the damped kernel are well known from
367 Smith's early work on multipoles for the Ewald
368 summation.\cite{Smith82,Smith98}
369
370 Note that increasing the value of $n$ will add additional terms to the
371 electrostatic potential, e.g., $f_2(r)$ includes orders up to
372 $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
373 functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
374 f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
375 for computing multipole energies, forces, and torques, and smooth
376 cutoffs of these quantities can be guaranteed as long as the number of
377 terms in the Taylor series exceeds the derivative order required.
378
379 For multipole-multipole interactions, following this procedure results
380 in separate radial functions for each distinct orientational
381 contribution to the potential, and ensures that the forces and torques
382 from {\it each} of these contributions will vanish at the cutoff
383 radius. For example, the direct dipole dot product ($\mathbf{D}_{i}
384 \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
385 dot products:
386 \begin{equation}
387 U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
388 \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
389 -\frac{1}{4\pi \epsilon_0}
390 \left( \mathbf{D}_{i} \cdot \hat{r} \right)
391 \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
392 \frac{g_2(r)}{r} \right)
393 \end{equation}
394
395 The electrostatic forces and torques acting on the central multipole
396 site due to another site within cutoff sphere are derived from
397 Eq.~\ref{generic}, accounting for the appropriate number of
398 derivatives. Complete energy, force, and torque expressions are
399 presented in the first paper in this series (Reference
400 \onlinecite{PaperI}).
401
402 \subsection{Gradient-shifted force (GSF)}
403
404 A second (and significantly simpler) method involves shifting the
405 gradient of the raw coulomb potential for each particular multipole
406 order. For example, the raw dipole-dipole potential energy may be
407 shifted smoothly by finding the gradient for two interacting dipoles
408 which have been projected onto the surface of the cutoff sphere
409 without changing their relative orientation,
410 \begin{displaymath}
411 U_{D_{i}D_{j}}(r_{ij}) = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(r_c)
412 - (r_{ij}-r_c) \hat{r}_{ij} \cdot
413 \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{r_c}
414 \end{displaymath}
415 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
416 and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
417 the signs are reversed for the dipole that has been projected onto the
418 cutoff sphere). In many ways, this simpler approach is closer in
419 spirit to the original shifted force method, in that it projects a
420 neutralizing multipole (and the resulting forces from this multipole)
421 onto a cutoff sphere. The resulting functional forms for the
422 potentials, forces, and torques turn out to be quite similar in form
423 to the Taylor-shifted approach, although the radial contributions are
424 significantly less perturbed by the Gradient-shifted approach than
425 they are in the Taylor-shifted method.
426
427 In general, the gradient shifted potential between a central multipole
428 and any multipolar site inside the cutoff radius is given by,
429 \begin{equation}
430 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
431 U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
432 \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert _{r_c} \right]
433 \label{generic2}
434 \end{equation}
435 where the sum describes a separate force-shifting that is applied to
436 each orientational contribution to the energy.
437
438 The third term converges more rapidly than the first two terms as a
439 function of radius, hence the contribution of the third term is very
440 small for large cutoff radii. The force and torque derived from
441 equation \ref{generic2} are consistent with the energy expression and
442 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
443 can be considered generalizations of the original DSF method for
444 higher order multipole interactions. GSF and TSF are also identical up
445 to the charge-dipole interaction but generate different expressions in
446 the energy, force and torque for higher order multipole-multipole
447 interactions. Complete energy, force, and torque expressions for the
448 GSF potential are presented in the first paper in this series
449 (Reference~\onlinecite{PaperI})
450
451
452 \subsection{Shifted potential (SP) }
453 A discontinuous truncation of the electrostatic potential at the
454 cutoff sphere introduces a severe artifact (oscillation in the
455 electrostatic energy) even for molecules with the higher-order
456 multipoles.\cite{PaperI} We have also formulated an extension of the
457 Wolf approach for point multipoles by simply projecting the image
458 multipole onto the surface of the cutoff sphere, and including the
459 interactions with the central multipole and the image. This
460 effectively shifts the total potential to zero at the cutoff radius,
461 \begin{equation}
462 U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
463 \label{eq:SP}
464 \end{equation}
465 where the sum describes separate potential shifting that is done for
466 each orientational contribution to the energy (e.g. the direct dipole
467 product contribution is shifted {\it separately} from the
468 dipole-distance terms in dipole-dipole interactions). Note that this
469 is not a simple shifting of the total potential at $r_c$. Each radial
470 contribution is shifted separately. One consequence of this is that
471 multipoles that reorient after leaving the cutoff sphere can re-enter
472 the cutoff sphere without perturbing the total energy.
473
474 The potential energy between a central multipole and other multipolar
475 sites then goes smoothly to zero as $r \rightarrow r_c$. However, the
476 force and torque obtained from the shifted potential (SP) are
477 discontinuous at $r_c$. Therefore, MD simulations will still
478 experience energy drift while operating under the SP potential, but it
479 may be suitable for Monte Carlo approaches where the configurational
480 energy differences are the primary quantity of interest.
481
482 \subsection{The Self term}
483 In the TSF, GSF, and SP methods, a self-interaction is retained for
484 the central multipole interacting with its own image on the surface of
485 the cutoff sphere. This self interaction is nearly identical with the
486 self-terms that arise in the Ewald sum for multipoles. Complete
487 expressions for the self terms are presented in the first paper in
488 this series (Reference \onlinecite{PaperI}).
489
490
491 \section{\label{sec:methodology}Methodology}
492
493 To understand how the real-space multipole methods behave in computer
494 simulations, it is vital to test against established methods for
495 computing electrostatic interactions in periodic systems, and to
496 evaluate the size and sources of any errors that arise from the
497 real-space cutoffs. In the first paper of this series, we compared
498 the dipolar and quadrupolar energy expressions against analytic
499 expressions for ordered dipolar and quadrupolar
500 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
501 used the multipolar Ewald sum as a reference method for comparing
502 energies, forces, and torques for molecular models that mimic
503 disordered and ordered condensed-phase systems. The parameters used
504 in the test-cases are given in table~\ref{tab:pars}.
505
506 \begin{table}
507 \label{tab:pars}
508 \caption{The parameters used in the systems used to evaluate the new
509 real-space methods. The most comprehensive test was a liquid
510 composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
511 ions). This test excercises all orders of the multipolar
512 interactions developed in the first paper.}
513 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
514 & \multicolumn{2}{c|}{LJ parameters} &
515 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
516 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
517 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
518 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
519 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
520 \AA\textsuperscript{2})} \\ \hline
521 Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
522 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & 10,000 & 17.6 &17.6 & 0 \\
523 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
524 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & 10,000 & 17.6&17.6&0 \\
525 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
526 \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
527 \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
528 \end{tabularx}
529 \end{table}
530 The systems consist of pure multipolar solids (both dipole and
531 quadrupole), pure multipolar liquids (both dipole and quadrupole), a
532 fluid composed of sites containing both dipoles and quadrupoles
533 simultaneously, and a final test case that includes ions with point
534 charges in addition to the multipolar fluid. The solid-phase
535 parameters were chosen so that the systems can explore some
536 orientational freedom for the multipolar sites, while maintaining
537 relatively strict translational order. The SSDQ model used here is
538 not a particularly accurate water model, but it does test
539 dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
540 interactions at roughly the same magnitudes. The last test case, SSDQ
541 water with dissolved ions, exercises \textit{all} levels of the
542 multipole-multipole interactions we have derived so far and represents
543 the most complete test of the new methods.
544
545 In the following section, we present results for the total
546 electrostatic energy, as well as the electrostatic contributions to
547 the force and torque on each molecule. These quantities have been
548 computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
549 and have been compared with the values obtaine from the multipolar
550 Ewald sum. In Mote Carlo (MC) simulations, the energy differences
551 between two configurations is the primary quantity that governs how
552 the simulation proceeds. These differences are the most imporant
553 indicators of the reliability of a method even if the absolute
554 energies are not exact. For each of the multipolar systems listed
555 above, we have compared the change in electrostatic potential energy
556 ($\Delta E$) between 250 statistically-independent configurations. In
557 molecular dynamics (MD) simulations, the forces and torques govern the
558 behavior of the simulation, so we also compute the electrostatic
559 contributions to the forces and torques.
560
561 \subsection{Implementation}
562 The real-space methods developed in the first paper in this series
563 have been implemented in our group's open source molecular simulation
564 program, OpenMD,\cite{openmd} which was used for all calculations in
565 this work. The complementary error function can be a relatively slow
566 function on some processors, so all of the radial functions are
567 precomputed on a fine grid and are spline-interpolated to provide
568 values when required.
569
570 Using the same simulation code, we compare to a multipolar Ewald sum
571 with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
572 the Ewald sum is a re-implementation of the algorithm originally
573 proposed by Smith that does not use the particle mesh or smoothing
574 approximations.\cite{Smith82,Smith98} In all cases, the quantities
575 being compared are the electrostatic contributions to energies, force,
576 and torques. All other contributions to these quantities (i.e. from
577 Lennard-Jones interactions) are removed prior to the comparisons.
578
579 The convergence parameter ($\alpha$) also plays a role in the balance
580 of the real-space and reciprocal-space portions of the Ewald
581 calculation. Typical molecular mechanics packages set this to a value
582 that depends on the cutoff radius and a tolerance (typically less than
583 $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
584 associated with increasing accuracy at the expense of computational
585 time spent on the reciprocal-space portion of the
586 summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
587 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
588 Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
589
590 The real-space models have self-interactions that provide
591 contributions to the energies only. Although the self interaction is
592 a rapid calculation, we note that in systems with fluctuating charges
593 or point polarizabilities, the self-term is not static and must be
594 recomputed at each time step.
595
596 \subsection{Model systems}
597 To sample independent configurations of multipolar crystals, a body
598 centered cubic (bcc) crystal which is a minimum energy structure for
599 point dipoles was generated using 3,456 molecules. The multipoles
600 were translationally locked in their respective crystal sites for
601 equilibration at a relatively low temperature (50K), so that dipoles
602 or quadrupoles could freely explore all accessible orientations. The
603 translational constraints were removed, and the crystals were
604 simulated for 10 ps in the microcanonical (NVE) ensemble with an
605 average temperature of 50 K. Configurations were sampled at equal
606 time intervals for the comparison of the configurational energy
607 differences. The crystals were not simulated close to the melting
608 points in order to avoid translational deformation away of the ideal
609 lattice geometry.
610
611 For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
612 system was created with 2048 molecules oriented randomly. These were
613
614 system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
615 temperature after equilibration. We collected 250 different
616 configurations in equal interval of time. For the ions mixed liquid
617 system, we converted 48 different molecules into 24 \ce{Na+} and 24
618 \ce{Cl-} ions and equilibrated. After equilibration, the system was run
619 at the same environment for 1ns and 250 configurations were
620 collected. While comparing energies, forces, and torques with Ewald
621 method, Lennard-Jones potentials were turned off and purely
622 electrostatic interaction had been compared.
623
624 \subsection{Accuracy of Energy Differences, Forces and Torques}
625 The pairwise summation techniques (outlined above) were evaluated for
626 use in MC simulations by studying the energy differences between
627 different configurations. We took the Ewald-computed energy
628 difference between two conformations to be the correct behavior. An
629 ideal performance by one of the new methods would reproduce these
630 energy differences exactly. The configurational energies being used
631 here contain only contributions from electrostatic interactions.
632 Lennard-Jones interactions were omitted from the comparison as they
633 should be identical for all methods.
634
635 Since none of the real-space methods provide exact energy differences,
636 we used least square regressions analysiss for the six different
637 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
638 with the multipolar Ewald reference method. Unitary results for both
639 the correlation (slope) and correlation coefficient for these
640 regressions indicate perfect agreement between the real-space method
641 and the multipolar Ewald sum.
642
643 Molecular systems were run long enough to explore independent
644 configurations and 250 configurations were recorded for comparison.
645 Each system provided 31,125 energy differences for a total of 186,750
646 data points. Similarly, the magnitudes of the forces and torques have
647 also been compared by using least squares regression analyses. In the
648 forces and torques comparison, the magnitudes of the forces acting in
649 each molecule for each configuration were evaluated. For example, our
650 dipolar liquid simulation contains 2048 molecules and there are 250
651 different configurations for each system resulting in 3,072,000 data
652 points for comparison of forces and torques.
653
654 \subsection{Analysis of vector quantities}
655 Getting the magnitudes of the force and torque vectors correct is only
656 part of the issue for carrying out accurate molecular dynamics
657 simulations. Because the real space methods reweight the different
658 orientational contributions to the energies, it is also important to
659 understand how the methods impact the \textit{directionality} of the
660 force and torque vectors. Fisher developed a probablity density
661 function to analyse directional data sets,
662 \begin{equation}
663 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
664 \label{eq:pdf}
665 \end{equation}
666 where $\kappa$ measures directional dispersion of the data around the
667 mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
668 estimated as a reciprocal of the circular variance.\cite{Allen91} To
669 quantify the directional error, forces obtained from the Ewald sum
670 were taken as the mean (or correct) direction and the angle between
671 the forces obtained via the Ewald sum and the real-space methods were
672 evaluated,
673 \begin{equation}
674 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
675 \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
676 \end{equation}
677 The total angular displacement of the vectors was calculated as,
678 \begin{equation}
679 R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
680 \label{eq:displacement}
681 \end{equation}
682 where $N$ is number of force vectors. The circular variance is
683 defined as
684 \begin{equation}
685 \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
686 \end{equation}
687 The circular variance takes on values between from 0 to 1, with 0
688 indicating a perfect directional match between the Ewald force vectors
689 and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
690 correspond to higher values of $\kappa$, which indicates tighter
691 clustering of the real-space force vectors around the Ewald forces.
692
693 A similar analysis was carried out for the electrostatic contribution
694 to the molecular torques as well as forces.
695
696 \subsection{Energy conservation}
697 To test conservation the energy for the methods, the mixed molecular
698 system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
699 ions was run for 1 ns in the microcanonical ensemble at an average
700 temperature of 300K. Each of the different electrostatic methods
701 (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
702 damping values. The molecular system was started with same initial
703 positions and velocities for all cutoff methods. The energy drift
704 ($\delta E_1$) and standard deviation of the energy about the slope
705 ($\delta E_0$) were evaluated from the total energy of the system as a
706 function of time. Although both measures are valuable at
707 investigating new methods for molecular dynamics, a useful interaction
708 model must allow for long simulation times with minimal energy drift.
709
710 \section{\label{sec:result}RESULTS}
711 \subsection{Configurational energy differences}
712 %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.
713 %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.
714 %\begin{figure}[h!]
715 % \centering
716 % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
717 % \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. }
718 % \label{fig:rcutConvergence_dipolarLiquid}
719 % \end{figure}
720 %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.
721 %\begin{figure}
722 % \centering
723 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
724 % \label{fig:barGraph1}
725 % \end{figure}
726 % \begin{figure}
727 % \centering
728 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
729 % \caption{}
730
731 % \label{fig:barGraph2}
732 % \end{figure}
733 %The correlation coefficient ($R^2$) and slope of the linear
734 %regression plots for the energy differences for all six different
735 %molecular systems is shown in figure 4a and 4b.The plot shows that
736 %the correlation coefficient improves for the SP cutoff method as
737 %compared to the undamped hard cutoff method in the case of SSDQC,
738 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
739 %crystal and liquid, the correlation coefficient is almost unchanged
740 %and close to 1. The correlation coefficient is smallest (0.696276
741 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
742 %charge-charge and charge-multipole interactions. Since the
743 %charge-charge and charge-multipole interaction is long ranged, there
744 %is huge deviation of correlation coefficient from 1. Similarly, the
745 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
746 %compared to interactions in the other multipolar systems, thus the
747 %correlation coefficient very close to 1 even for hard cutoff
748 %method. The idea of placing image multipole on the surface of the
749 %cutoff sphere improves the correlation coefficient and makes it close
750 %to 1 for all types of multipolar systems. Similarly the slope is
751 %hugely deviated from the correct value for the lower order
752 %multipole-multipole interaction and slightly deviated for higher
753 %order multipole – multipole interaction. The SP method improves both
754 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
755 %dipolar systems. The Slope is found to be deviated more in dipolar
756 %crystal as compared to liquid which is associated with the large
757 %fluctuation in the electrostatic energy in crystal. The GSF also
758 %produced better values of correlation coefficient and slope with the
759 %proper selection of the damping alpha (Interested reader can consult
760 %accompanying supporting material). The TSF method gives good value of
761 %correlation coefficient for the dipolar crystal, dipolar liquid,
762 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
763 %regression slopes are significantly deviated.
764
765 \begin{figure}
766 \centering
767 \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
768 \caption{Statistical analysis of the quality of configurational
769 energy differences for the real-space electrostatic methods
770 compared with the reference Ewald sum. Results with a value equal
771 to 1 (dashed line) indicate $\Delta E$ values indistinguishable
772 from those obtained using the multipolar Ewald sum. Different
773 values of the cutoff radius are indicated with different symbols
774 (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
775 triangles).}
776 \label{fig:slopeCorr_energy}
777 \end{figure}
778
779 The combined correlation coefficient and slope for all six systems is
780 shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
781 reproduce the Ewald configurational energy differences with remarkable
782 fidelity. Undamped hard cutoffs introduce a significant amount of
783 random scatter in the energy differences which is apparent in the
784 reduced value of the correlation coefficient for this method. This
785 can be easily understood as configurations which exhibit small
786 traversals of a few dipoles or quadrupoles out of the cutoff sphere
787 will see large energy jumps when hard cutoffs are used. The
788 orientations of the multipoles (particularly in the ordered crystals)
789 mean that these energy jumps can go in either direction, producing a
790 significant amount of random scatter, but no systematic error.
791
792 The TSF method produces energy differences that are highly correlated
793 with the Ewald results, but it also introduces a significant
794 systematic bias in the values of the energies, particularly for
795 smaller cutoff values. The TSF method alters the distance dependence
796 of different orientational contributions to the energy in a
797 non-uniform way, so the size of the cutoff sphere can have a large
798 effect, particularly for the crystalline systems.
799
800 Both the SP and GSF methods appear to reproduce the Ewald results with
801 excellent fidelity, particularly for moderate damping ($\alpha =
802 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
803 12$\AA). With the exception of the undamped hard cutoff, and the TSF
804 method with short cutoffs, all of the methods would be appropriate for
805 use in Monte Carlo simulations.
806
807 \subsection{Magnitude of the force and torque vectors}
808
809 The comparisons of the magnitudes of the forces and torques for the
810 data accumulated from all six systems are shown in Figures
811 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
812 correlation and slope for the forces agree well with the Ewald sum
813 even for the hard cutoffs.
814
815 For systems of molecules with only multipolar interactions, the pair
816 energy contributions are quite short ranged. Moreover, the force
817 decays more rapidly than the electrostatic energy, hence the hard
818 cutoff method can also produce reasonable agreement for this quantity.
819 Although the pure cutoff gives reasonably good electrostatic forces
820 for pairs of molecules included within each other's cutoff spheres,
821 the discontinuity in the force at the cutoff radius can potentially
822 cause energy conservation problems as molecules enter and leave the
823 cutoff spheres. This is discussed in detail in section
824 \ref{sec:conservation}.
825
826 The two shifted-force methods (GSF and TSF) exhibit a small amount of
827 systematic variation and scatter compared with the Ewald forces. The
828 shifted-force models intentionally perturb the forces between pairs of
829 molecules inside each other's cutoff spheres in order to correct the
830 energy conservation issues, and this perturbation is evident in the
831 statistics accumulated for the molecular forces. The GSF
832 perturbations are minimal, particularly for moderate damping and and
833 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
834 reasonable agreement in the correlation coefficient but again the
835 systematic error in the forces is concerning if replication of Ewald
836 forces is desired.
837
838 \begin{figure}
839 \centering
840 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
841 \caption{Statistical analysis of the quality of the force vector
842 magnitudes for the real-space electrostatic methods compared with
843 the reference Ewald sum. Results with a value equal to 1 (dashed
844 line) indicate force magnitude values indistinguishable from those
845 obtained using the multipolar Ewald sum. Different values of the
846 cutoff radius are indicated with different symbols (9\AA\ =
847 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
848 \label{fig:slopeCorr_force}
849 \end{figure}
850
851
852 \begin{figure}
853 \centering
854 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
855 \caption{Statistical analysis of the quality of the torque vector
856 magnitudes for the real-space electrostatic methods compared with
857 the reference Ewald sum. Results with a value equal to 1 (dashed
858 line) indicate force magnitude values indistinguishable from those
859 obtained using the multipolar Ewald sum. Different values of the
860 cutoff radius are indicated with different symbols (9\AA\ =
861 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
862 \label{fig:slopeCorr_torque}
863 \end{figure}
864
865 The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
866 significantly influenced by the choice of real-space method. The
867 torque expressions have the same distance dependence as the energies,
868 which are naturally longer-ranged expressions than the inter-site
869 forces. Torques are also quite sensitive to orientations of
870 neighboring molecules, even those that are near the cutoff distance.
871
872 The results shows that the torque from the hard cutoff method
873 reproduces the torques in quite good agreement with the Ewald sum.
874 The other real-space methods can cause some deviations, but excellent
875 agreement with the Ewald sum torques is recovered at moderate values
876 of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
877 radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
878 in the slope when compared with the Ewald torques even for larger
879 cutoff radii. It appears that the severity of the perturbations in
880 the TSF method are most in evidence for the torques.
881
882 \subsection{Directionality of the force and torque vectors}
883
884 The accurate evaluation of force and torque directions is just as
885 important for molecular dynamics simulations as the magnitudes of
886 these quantities. Force and torque vectors for all six systems were
887 analyzed using Fisher statistics, and the quality of the vector
888 directionality is shown in terms of circular variance
889 ($\mathrm{Var}(\theta$) in figure
890 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
891 from the new real-space methods exhibit nearly-ideal Fisher probability
892 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
893 exhibit the best vectorial agreement with the Ewald sum. The force and
894 torque vectors from GSF method also show good agreement with the Ewald
895 method, which can also be systematically improved by using moderate
896 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
897 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
898 to a distribution with 95\% of force vectors within $6.37^\circ$ of
899 the corresponding Ewald forces. The TSF method produces the poorest
900 agreement with the Ewald force directions.
901
902 Torques are again more perturbed than the forces by the new real-space
903 methods, but even here the variance is reasonably small. For the same
904 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
905 the circular variance was 0.01415, corresponds to a distribution which
906 has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
907 results. Again, the direction of the force and torque vectors can be
908 systematically improved by varying $\alpha$ and $r_c$.
909
910 \begin{figure}
911 \centering
912 \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
913 \caption{The circular variance of the direction of the force and
914 torque vectors obtained from the real-space methods around the
915 reference Ewald vectors. A variance equal to 0 (dashed line)
916 indicates direction of the force or torque vectors are
917 indistinguishable from those obtained from the Ewald sum. Here
918 different symbols represent different values of the cutoff radius
919 (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
920 \label{fig:slopeCorr_circularVariance}
921 \end{figure}
922
923 \subsection{Energy conservation\label{sec:conservation}}
924
925 We have tested the conservation of energy one can expect to see with
926 the new real-space methods using the SSDQ water model with a small
927 fraction of solvated ions. This is a test system which exercises all
928 orders of multipole-multipole interactions derived in the first paper
929 in this series and provides the most comprehensive test of the new
930 methods. A liquid-phase system was created with 2000 water molecules
931 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
932 temperature of 300K. After equilibration, this liquid-phase system
933 was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
934 a cutoff radius of 12\AA. The value of the damping coefficient was
935 also varied from the undamped case ($\alpha = 0$) to a heavily damped
936 case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods. A
937 sample was also run using the multipolar Ewald sum with the same
938 real-space cutoff.
939
940 In figure~\ref{fig:energyDrift} we show the both the linear drift in
941 energy over time, $\delta E_1$, and the standard deviation of energy
942 fluctuations around this drift $\delta E_0$. Both of the
943 shifted-force methods (GSF and TSF) provide excellent energy
944 conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
945 while the hard cutoff is essentially unusable for molecular dynamics.
946 SP provides some benefit over the hard cutoff because the energetic
947 jumps that happen as particles leave and enter the cutoff sphere are
948 somewhat reduced, but like the Wolf method for charges, the SP method
949 would not be as useful for molecular dynamics as either of the
950 shifted-force methods.
951
952 We note that for all tested values of the cutoff radius, the new
953 real-space methods can provide better energy conservation behavior
954 than the multipolar Ewald sum, even when utilizing a relatively large
955 $k$-space cutoff values.
956
957 \begin{figure}
958 \centering
959 \includegraphics[width=\textwidth]{newDrift.pdf}
960 \label{fig:energyDrift}
961 \caption{Analysis of the energy conservation of the real-space
962 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
963 energy over time and $\delta \mathrm{E}_0$ is the standard deviation
964 of energy fluctuations around this drift. All simulations were of a
965 2000-molecule simulation of SSDQ water with 48 ionic charges at 300
966 K starting from the same initial configuration. All runs utilized
967 the same real-space cutoff, $r_c = 12$\AA.}
968 \end{figure}
969
970
971 \section{CONCLUSION}
972 In the first paper in this series, we generalized the
973 charge-neutralized electrostatic energy originally developed by Wolf
974 \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
975 up to quadrupolar order. The SP method is essentially a
976 multipole-capable version of the Wolf model. The SP method for
977 multipoles provides excellent agreement with Ewald-derived energies,
978 forces and torques, and is suitable for Monte Carlo simulations,
979 although the forces and torques retain discontinuities at the cutoff
980 distance that prevents its use in molecular dynamics.
981
982 We also developed two natural extensions of the damped shifted-force
983 (DSF) model originally proposed by Fennel and
984 Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
985 smooth truncation of energies, forces, and torques at the real-space
986 cutoff, and both converge to DSF electrostatics for point-charge
987 interactions. The TSF model is based on a high-order truncated Taylor
988 expansion which can be relatively perturbative inside the cutoff
989 sphere. The GSF model takes the gradient from an images of the
990 interacting multipole that has been projected onto the cutoff sphere
991 to derive shifted force and torque expressions, and is a significantly
992 more gentle approach.
993
994 Of the two newly-developed shifted force models, the GSF method
995 produced quantitative agreement with Ewald energy, force, and torques.
996 It also performs well in conserving energy in MD simulations. The
997 Taylor-shifted (TSF) model provides smooth dynamics, but these take
998 place on a potential energy surface that is significantly perturbed
999 from Ewald-based electrostatics.
1000
1001 % The direct truncation of any electrostatic potential energy without
1002 % multipole neutralization creates large fluctuations in molecular
1003 % simulations. This fluctuation in the energy is very large for the case
1004 % of crystal because of long range of multipole ordering (Refer paper
1005 % I).\cite{PaperI} This is also significant in the case of the liquid
1006 % because of the local multipole ordering in the molecules. If the net
1007 % multipole within cutoff radius neutralized within cutoff sphere by
1008 % placing image multiples on the surface of the sphere, this fluctuation
1009 % in the energy reduced significantly. Also, the multipole
1010 % neutralization in the generalized SP method showed very good agreement
1011 % with the Ewald as compared to direct truncation for the evaluation of
1012 % the $\triangle E$ between the configurations. In MD simulations, the
1013 % energy conservation is very important. The conservation of the total
1014 % energy can be ensured by i) enforcing the smooth truncation of the
1015 % energy, force and torque in the cutoff radius and ii) making the
1016 % energy, force and torque consistent with each other. The GSF and TSF
1017 % methods ensure the consistency and smooth truncation of the energy,
1018 % force and torque at the cutoff radius, as a result show very good
1019 % total energy conservation. But the TSF method does not show good
1020 % agreement in the absolute value of the electrostatic energy, force and
1021 % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1022 % energy and torque accurately and also conserved energy.
1023
1024 The only cases we have found where the new GSF and SP real-space
1025 methods can be problematic are those which retain a bulk dipole moment
1026 at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1027 materials, uniform weighting of the orientational contributions can be
1028 important for converging the total energy. In these cases, the
1029 damping function which causes the non-uniform weighting can be
1030 replaced by the bare electrostatic kernel, and the energies return to
1031 the expected converged values.
1032
1033 Based on the results of this work, the GSF method is a suitable and
1034 efficient replacement for the Ewald sum for evaluating electrostatic
1035 interactions in MD simulations. Both methods retain excellent
1036 fidelity to the Ewald energies, forces and torques. Additionally, the
1037 energy drift and fluctuations from the GSF electrostatics are better
1038 than a multipolar Ewald sum for finite-sized reciprocal spaces.
1039 Because they use real-space cutoffs with moderate cutoff radii, the
1040 GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1041 increases. Additionally, they can be made extremely efficient using
1042 spline interpolations of the radial functions. They require no
1043 Fourier transforms or $k$-space sums, and guarantee the smooth
1044 handling of energies, forces, and torques as multipoles cross the
1045 real-space cutoff boundary.
1046
1047 %\bibliographystyle{aip}
1048 \newpage
1049 \bibliography{references}
1050 \end{document}
1051
1052 %
1053 % ****** End of file aipsamp.tex ******