ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4171
Committed: Wed Jun 4 19:31:06 2014 UTC (10 years ago) by gezelter
Content type: application/x-tex
File size: 53105 byte(s)
Log Message:
New figure

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[version=3]{mhchem} % this is a great package for formatting chemical reactions
40 \usepackage{url}
41 \usepackage[english]{babel}
42
43
44 \begin{document}
45
46 \preprint{AIP/123-QED}
47
48 \title[Efficient electrostatics for condensed-phase multipoles]{Real space alternatives to the Ewald
49 Sum. II. Comparison of Simulation Methodologies} % Force line breaks with \\
50
51 \author{Madan Lamichhane}
52 \affiliation{Department of Physics, University
53 of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
54
55 \author{Kathie E. Newman}
56 \affiliation{Department of Physics, University
57 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%\\This line break forced with \textbackslash\textbackslash
62 }%
63
64 \date{\today}% It is always \today, today,
65 % but any date may be explicitly specified
66
67 \begin{abstract}
68 We have tested our recently developed shifted potential, gradient-shifted force, and Taylor-shifted force methods for the higher-order multipoles against Ewald’s method in different types of liquid and crystalline system. In this paper, we have also investigated the conservation of total energy in the molecular dynamic simulation using all of these methods. The shifted potential method shows better agreement with the Ewald in the energy differences between different configurations as compared to the direct truncation. Both the gradient shifted force and Taylor-shifted force methods reproduce very good energy conservation. But the absolute energy, force and torque evaluated from the gradient shifted force method shows better result as compared to taylor-shifted force method. Hence the gradient-shifted force method suitably mimics the electrostatic interaction in the molecular dynamic simulation.
69 \end{abstract}
70
71 \pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
72 % Classification Scheme.
73 \keywords{Electrostatics, Multipoles, Real-space}
74
75 \maketitle
76
77
78 \section{\label{sec:intro}Introduction}
79 Computing the interactions between electrostatic sites is one of the
80 most expensive aspects of molecular simulations, which is why there
81 have been significant efforts to develop practical, efficient and
82 convergent methods for handling these interactions. Ewald's method is
83 perhaps the best known and most accurate method for evaluating
84 energies, forces, and torques in explicitly-periodic simulation
85 cells. In this approach, the conditionally convergent electrostatic
86 energy is converted into two absolutely convergent contributions, one
87 which is carried out in real space with a cutoff radius, and one in
88 reciprocal space.\cite{Clarke:1986eu,Woodcock75}
89
90 When carried out as originally formulated, the reciprocal-space
91 portion of the Ewald sum exhibits relatively poor computational
92 scaling, making it prohibitive for large systems. By utilizing
93 particle meshes and three dimensional fast Fourier transforms (FFT),
94 the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
95 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
96 the computational cost from $O(N^2)$ down to $O(N \log
97 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
98
99 Because of the artificial periodicity required for the Ewald sum, the
100 method may require modification to compute interactions for
101 interfacial molecular systems such as membranes and liquid-vapor
102 interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
103 To simulate interfacial systems, Parry’s extension of the 3D Ewald sum
104 is appropriate for slab geometries.\cite{Parry:1975if} The inherent
105 periodicity in the Ewald’s method can also be problematic for
106 interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
107 methods that were developed to handle two-dimensional (2D)
108 electrostatic interactions in interfacial systems have not had similar
109 particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
110 Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
111
112 \subsection{Real-space methods}
113 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
114 method for calculating electrostatic interactions between point
115 charges. They argued that the effective Coulomb interaction in
116 condensed systems is actually short ranged.\cite{Wolf92,Wolf95}. For
117 an ordered lattice (e.g. when computing the Madelung constant of an
118 ionic solid), the material can be considered as a set of ions
119 interacting with neutral dipolar or quadrupolar ``molecules'' giving
120 an effective distance dependence for the electrostatic interactions of
121 $r^{-5}$ (see figure \ref{fig:NaCl}. For this reason, careful
122 applications of Wolf's method are able to obtain accurate estimates of
123 Madelung constants using relatively short cutoff radii. Recently,
124 Fukuda used neutralization of the higher order moments for the
125 calculation of the electrostatic interaction of the point charges
126 system.\cite{Fukuda:2013sf}
127
128 \begin{figure}[h!]
129 \centering
130 \includegraphics[width=0.50 \textwidth]{chargesystem.pdf}
131 \caption{Top: NaCl crystal showing how spherical truncation can
132 breaking effective charge ordering, and how complete \ce{(NaCl)4}
133 molecules interact with the central ion. Bottom: A dipolar
134 crystal exhibiting similar behavior and illustrating how the
135 effective dipole-octupole interactions can be disrupted by
136 spherical truncation.}
137 \label{fig:NaCl}
138 \end{figure}
139
140 The direct truncation of interactions at a cutoff radius creates
141 truncation defects. Wolf \textit{et al.} further argued that
142 truncation errors are due to net charge remaining inside the cutoff
143 sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
144 placing an image charge on the surface of the cutoff sphere for every
145 real charge inside the cutoff. These charges are present for the
146 evaluation of both the pair interaction energy and the force, although
147 the force expression maintained a discontinuity at the cutoff sphere.
148 In the original Wolf formulation, the total energy for the charge and
149 image were not equal to the integral of their force expression, and as
150 a result, the total energy would not be conserved in molecular
151 dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
152 Fennel and Gezelter later proposed shifted force variants of the Wolf
153 method with commensurate force and energy expressions that do not
154 exhibit this problem.\cite{Fennell:2006lq} Related real-space
155 methods were also proposed by Chen \textit{et
156 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
157 and by Wu and Brooks.\cite{Wu:044107}
158
159 Considering the interaction of one central ion in an ionic crystal
160 with a portion of the crystal at some distance, the effective Columbic
161 potential is found to be decreasing as $r^{-5}$. If one views the
162 \ce{NaCl} crystal as simple cubic (SC) structure with an octupolar
163 \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
164 rapidly to the Madelung energy than the dipolar
165 approximation.\cite{Wolf92} To find the correct Madelung constant,
166 Lacman suggested that the NaCl structure could be constructed in a way
167 that the finite crystal terminates with complete \ce{(NaCl)4}
168 molecules.\cite{Lacman65} Any charge in a NaCl crystal is surrounded
169 by opposite charges. Similarly for each pair of charges, there is an
170 opposite pair of charge adjacent to it. The central ion sees what is
171 effectively a set of octupoles at large distances. These facts suggest
172 that the Madelung constants are relatively short ranged for perfect
173 ionic crystals.\cite{Wolf:1999dn}
174
175 One can make a similar argument for crystals of point multipoles. The
176 Luttinger and Tisza treatment of energy constants for dipolar lattices
177 utilizes 24 basis vectors that contain dipoles at the eight corners of
178 a unit cube. Only three of these basis vectors, $X_1, Y_1,
179 \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
180 zero net dipole and retain contributions only from higher order
181 multipoles. The effective interaction between a dipole at the center
182 of a crystal and a group of eight dipoles farther away is
183 significantly shorter ranged than the $r^{-3}$ that one would expect
184 for raw dipole-dipole interactions. Only in crystals which retain a
185 bulk dipole moment (e.g. ferroelectrics) does the analogy with the
186 ionic crystal break down -- ferroelectric dipolar crystals can exist,
187 while ionic crystals with net charge in each unit cell would be
188 unstable.
189
190 In ionic crystals, real-space truncation can break the effective
191 multipolar arrangements (see Fig. \ref{fig:NaCl}), causing significant
192 swings in the electrostatic energy as the cutoff radius is increased
193 (or as individual ions move back and forth across the boundary). This
194 is why the image charges were necessary for the Wolf sum to exhibit
195 rapid convergence. Similarly, the real-space truncation of point
196 multipole interactions breaks higher order multipole arrangements, and
197 image multipoles are required for real-space treatments of
198 electrostatic energies.
199
200 % Because of this reason, although the nature of electrostatic
201 % interaction short ranged, the hard cutoff sphere creates very large
202 % fluctuation in the electrostatic energy for the perfect crystal. In
203 % addition, the charge neutralized potential proposed by Wolf et
204 % al. converged to correct Madelung constant but still holds oscillation
205 % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
206 % oscillation in the energy around its fully converged value can be due
207 % to the non-neutralized value of the higher order moments within the
208 % cutoff sphere.
209
210 The forces and torques acting on atomic sites are the fundamental
211 factors driving dynamics in molecular simulations. Fennell and
212 Gezelter proposed the damped shifted force (DSF) energy kernel to
213 obtain consistent energies and forces on the atoms within the cutoff
214 sphere. Both the energy and the force go smoothly to zero as an atom
215 aproaches the cutoff radius. The comparisons of the accuracy these
216 quantities between the DSF kernel and SPME was surprisingly
217 good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
218 calculating electrostatic interactions in molecular systems with
219 relatively uniform charge
220 densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
221
222 \subsection{The damping function}
223 The damping function used in our research has been discussed in detail
224 in the first paper of this series.\cite{PaperI} The radial kernel
225 $1/r$ for the interactions between point charges can be replaced by
226 the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
227 accelerate the rate of convergence, where $\alpha$ is a damping
228 parameter with units of inverse distance. Altering the value of
229 $\alpha$ is equivalent to changing the width of Gaussian charge
230 distributions that replace each point charge -- Gaussian overlap
231 integrals yield complementary error functions when truncated at a
232 finite distance.
233
234 By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
235 cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
236 agreement with SPME for the interaction energies, forces and torques
237 for charge-charge interactions.\cite{Fennell:2006lq}
238
239 \subsection{Point multipoles in molecular modeling}
240 Coarse-graining approaches which treat entire molecular subsystems as
241 a single rigid body are now widely used. A common feature of many
242 coarse-graining approaches is simplification of the electrostatic
243 interactions between bodies so that fewer site-site interactions are
244 required to compute configurational energies. Many coarse-grained
245 molecular structures would normally consist of equal positive and
246 negative charges, and rather than use multiple site-site interactions,
247 the interaction between higher order multipoles can also be used to
248 evaluate a single molecule-molecule
249 interaction.\cite{Ren06,Essex10,Essex11}
250
251 Because electrons in a molecule are not localized at specific points,
252 the assignment of partial charges to atomic centers is a relatively
253 rough approximation. Atomic sites can also be assigned point
254 multipoles and polarizabilities to increase the accuracy of the
255 molecular model. Recently, water has been modeled with point
256 multipoles up to octupolar
257 order.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
258 multipoles up to quadrupolar order have also been coupled with point
259 polarizabilities in the high-quality AMOEBA and iAMOEBA water
260 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk}. But
261 using point multipole with the real space truncation without
262 accounting for multipolar neutrality will create energy conservation
263 issues in molecular dynamics (MD) simulations.
264
265 In this paper we test a set of real-space methods that were developed
266 for point multipolar interactions. These methods extend the damped
267 shifted force (DSF) and Wolf methods originally developed for
268 charge-charge interactions and generalize them for higher order
269 multipoles. The detailed mathematical development of these methods has
270 been presented in the first paper in this series, while this work
271 covers the testing the energies, forces, torques, and energy
272 conservation properties of the methods in realistic simulation
273 environments. In all cases, the methods are compared with the
274 reference method, a full multipolar Ewald treatment.
275
276
277 %\subsection{Conservation of total energy }
278 %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.
279
280 \section{\label{sec:method}Review of Methods}
281 Any real-space electrostatic method that is suitable for MD
282 simulations should have the electrostatic energy, forces and torques
283 between two sites go smoothly to zero as the distance between the
284 sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
285 this continuity at the cutoff is essential for energy conservation in
286 MD simulations. The mathematical details of the shifted potential
287 (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
288 methods have been discussed in detail in the previous paper in this
289 series.\cite{PaperI} Here we briefly review the new methods and
290 describe their essential features.
291
292 \subsection{Taylor-shifted force (TSF)}
293
294 The electrostatic potential energy between point multipoles can be
295 expressed as the product of two multipole operators and a Coulombic
296 kernel,
297 \begin{equation}
298 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
299 \end{equation}
300 where the multipole operator for site $\bf a$,
301 \begin{equation}
302 \hat{M}_{\bf a} = C_{\bf a} - D_{{\bf a}\alpha} \frac{\partial}{\partial r_{\alpha}}
303 + Q_{{\bf a}\alpha\beta}
304 \frac{\partial^2}{\partial r_{\alpha} \partial r_{\beta}} + \dots
305 \end{equation}
306 is expressed in terms of the point charge, $C_{\bf a}$, dipole,
307 $D_{{\bf a}\alpha}$, and quadrupole, $Q_{{\bf a}\alpha\beta}$, for
308 object $\bf a$. Note that in this work, we use the primitive
309 quadrupole tensor, $Q_{a\alpha,\beta}=\frac{1}{2}\sum_{k\;in\;a}q_k
310 r_{k\alpha}r_{k\beta}$ to represent point quadrupoles on a site.
311
312 Interactions between multipoles can be expressed as higher derivatives
313 of the bare Coulomb potential, so one way of ensuring that the forces
314 and torques vanish at the cutoff distance is to include a larger
315 number of terms in the truncated Taylor expansion, e.g.,
316 %
317 \begin{equation}
318 f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-R_c)^m}{m!} f^{(m)} \Big \lvert _{R_c} .
319 \end{equation}
320 %
321 The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
322 Thus, for $f(r)=1/r$, we find
323 %
324 \begin{equation}
325 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} .
326 \end{equation}
327 This function is an approximate electrostatic potential that has
328 vanishing second derivatives at the cutoff radius, making it suitable
329 for shifting the forces and torques of charge-dipole interactions.
330
331 In general, the TSF potential for any multipole-multipole interaction
332 can be written
333 \begin{equation}
334 U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
335 \label{generic}
336 \end{equation}
337 with $n=0$ for charge-charge, $n=1$ for charge-dipole, $n=2$ for
338 charge-quadrupole and dipole-dipole, $n=3$ for dipole-quadrupole, and
339 $n=4$ for quadrupole-quadrupole. To ensure smooth convergence of the
340 energy, force, and torques, the required number of terms from Taylor
341 series expansion in $f_n(r)$ must be performed for different
342 multipole-multipole interactions.
343
344 To carry out the same procedure for a damped electrostatic kernel, we
345 replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
346 Many of the derivatives of the damped kernel are well known from
347 Smith's early work on multipoles for the Ewald
348 summation.\cite{Smith82,Smith98}
349
350 Note that increasing the value of $n$ will add additional terms to the
351 electrostatic potential, e.g., $f_2(r)$ includes orders up to
352 $(r-R_c)^3/R_c^4$, and so on. Successive derivatives of the $f_n(r)$
353 functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
354 f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
355 for computing multipole energies, forces, and torques, and smooth
356 cutoffs of these quantities can be guaranteed as long as the number of
357 terms in the Taylor series exceeds the derivative order required.
358
359 For multipole-multipole interactions, following this procedure results
360 in separate radial functions for each distinct orientational
361 contribution to the potential, and ensures that the forces and torques
362 from {\it each} of these contributions will vanish at the cutoff
363 radius. For example, the direct dipole dot product ($\mathbf{D}_{i}
364 \cdot \mathbf{D}_{j}$) is treated differently than the dipole-distance
365 dot products:
366 \begin{equation}
367 U_{D_{i}D_{j}}(r)= -\frac{1}{4\pi \epsilon_0} \left( \mathbf{D}_{i} \cdot
368 \mathbf{D}_{j} \right) \frac{g_2(r)}{r}
369 -\frac{1}{4\pi \epsilon_0}
370 \left( \mathbf{D}_{i} \cdot \hat{r} \right)
371 \left( \mathbf{D}_{j} \cdot \hat{r} \right) \left(h_2(r) -
372 \frac{g_2(r)}{r} \right)
373 \end{equation}
374
375 The electrostatic forces and torques acting on the central multipole
376 site due to another site within cutoff sphere are derived from
377 Eq.~\ref{generic}, accounting for the appropriate number of
378 derivatives. Complete energy, force, and torque expressions are
379 presented in the first paper in this series (Reference
380 \citep{PaperI}).
381
382 \subsection{Gradient-shifted force (GSF)}
383
384 A second (and significantly simpler) method involves shifting the
385 gradient of the raw coulomb potential for each particular multipole
386 order. For example, the raw dipole-dipole potential energy may be
387 shifted smoothly by finding the gradient for two interacting dipoles
388 which have been projected onto the surface of the cutoff sphere
389 without changing their relative orientation,
390 \begin{displaymath}
391 U_{D_{i}D_{j}}(r_{ij}) = U_{D_{i}D_{j}}(r_{ij}) - U_{D_{i}D_{j}}(R_c)
392 - (r_{ij}-R_c) \hat{r}_{ij} \cdot
393 \vec{\nabla} V_{D_{i}D_{j}}(r) \Big \lvert _{R_c}
394 \end{displaymath}
395 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{i}$
396 and $\mathbf{D}_{j}$, are retained at the cutoff distance (although
397 the signs are reversed for the dipole that has been projected onto the
398 cutoff sphere). In many ways, this simpler approach is closer in
399 spirit to the original shifted force method, in that it projects a
400 neutralizing multipole (and the resulting forces from this multipole)
401 onto a cutoff sphere. The resulting functional forms for the
402 potentials, forces, and torques turn out to be quite similar in form
403 to the Taylor-shifted approach, although the radial contributions are
404 significantly less perturbed by the Gradient-shifted approach than
405 they are in the Taylor-shifted method.
406
407 In general, the gradient shifted potential between a central multipole
408 and any multipolar site inside the cutoff radius is given by,
409 \begin{equation}
410 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
411 U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
412 \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert _{r_c} \right]
413 \label{generic2}
414 \end{equation}
415 where the sum describes a separate force-shifting that is applied to
416 each orientational contribution to the energy.
417
418 The third term converges more rapidly than the first two terms as a
419 function of radius, hence the contribution of the third term is very
420 small for large cutoff radii. The force and torque derived from
421 equation \ref{generic2} are consistent with the energy expression and
422 approach zero as $r \rightarrow R_c$. Both the GSF and TSF methods
423 can be considered generalizations of the original DSF method for
424 higher order multipole interactions. GSF and TSF are also identical up
425 to the charge-dipole interaction but generate different expressions in
426 the energy, force and torque for higher order multipole-multipole
427 interactions. Complete energy, force, and torque expressions for the
428 GSF potential are presented in the first paper in this series
429 (Reference \citep{PaperI})
430
431
432 \subsection{Shifted potential (SP) }
433 A discontinuous truncation of the electrostatic potential at the
434 cutoff sphere introduces a severe artifact (oscillation in the
435 electrostatic energy) even for molecules with the higher-order
436 multipoles.\cite{PaperI} We have also formulated an extension of the
437 Wolf approach for point multipoles by simply projecting the image
438 multipole onto the surface of the cutoff sphere, and including the
439 interactions with the central multipole and the image. This
440 effectively shifts the total potential to zero at the cutoff radius,
441 \begin{equation}
442 U_{SP}(\vec r)=\sum U(\vec r) - U(\vec r_c)
443 \label{eq:SP}
444 \end{equation}
445 where the sum describes separate potential shifting that is done for
446 each orientational contribution to the energy (e.g. the direct dipole
447 product contribution is shifted {\it separately} from the
448 dipole-distance terms in dipole-dipole interactions). Note that this
449 is not a simple shifting of the total potential at $R_c$. Each radial
450 contribution is shifted separately. One consequence of this is that
451 multipoles that reorient after leaving the cutoff sphere can re-enter
452 the cutoff sphere without perturbing the total energy.
453
454 The potential energy between a central multipole and other multipolar
455 sites then goes smoothly to zero as $r \rightarrow R_c$. However, the
456 force and torque obtained from the shifted potential (SP) are
457 discontinuous at $R_c$. Therefore, MD simulations will still
458 experience energy drift while operating under the SP potential, but it
459 may be suitable for Monte Carlo approaches where the configurational
460 energy differences are the primary quantity of interest.
461
462 \subsection{The Self term}
463 In the TSF, GSF, and SP methods, a self-interaction is retained for
464 the central multipole interacting with its own image on the surface of
465 the cutoff sphere. This self interaction is nearly identical with the
466 self-terms that arise in the Ewald sum for multipoles. Complete
467 expressions for the self terms are presented in the first paper in
468 this series (Reference \citep{PaperI})
469
470
471 \section{\label{sec:methodology}Methodology}
472
473 To understand how the real-space multipole methods behave in computer
474 simulations, it is vital to test against established methods for
475 computing electrostatic interactions in periodic systems, and to
476 evaluate the size and sources of any errors that arise from the
477 real-space cutoffs. In the first paper of this series, we compared
478 the dipolar and quadrupolar energy expressions against analytic
479 expressions for ordered dipolar and quadrupolar
480 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} This work uses the
481 multipolar Ewald sum as a reference method for comparing energies,
482 forces, and torques for molecular models that mimic disordered and
483 ordered condensed-phase systems. These test-cases include:
484
485 \begin{itemize}
486 \item Soft Dipolar fluids ($\sigma = , \epsilon = , |D| = $)
487 \item Soft Dipolar solids ($\sigma = , \epsilon = , |D| = $)
488 \item Soft Quadrupolar fluids ($\sigma = , \epsilon = , Q_{xx} = ...$)
489 \item Soft Quadrupolar solids ($\sigma = , \epsilon = , Q_{xx} = ...$)
490 \item A mixed multipole model for water
491 \item A mixed multipole models for water with dissolved ions
492 \end{itemize}
493 This last test case exercises all levels of the multipole-multipole
494 interactions we have derived so far and represents the most complete
495 test of the new methods.
496
497 In the following section, we present results for the total
498 electrostatic energy, as well as the electrostatic contributions to
499 the force and torque on each molecule. These quantities have been
500 computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
501 and have been compared with the values obtaine from the multipolar
502 Ewald sum. In Mote Carlo (MC) simulations, the energy differences
503 between two configurations is the primary quantity that governs how
504 the simulation proceeds. These differences are the most imporant
505 indicators of the reliability of a method even if the absolute
506 energies are not exact. For each of the multipolar systems listed
507 above, we have compared the change in electrostatic potential energy
508 ($\Delta E$) between 250 statistically-independent configurations. In
509 molecular dynamics (MD) simulations, the forces and torques govern the
510 behavior of the simulation, so we also compute the electrostatic
511 contributions to the forces and torques.
512
513 \subsection{Model systems}
514 To sample independent configurations of multipolar crystals, a body
515 centered cubic (BCC) crystal which is a minimum energy structure for
516 point dipoles was generated using 3,456 molecules. The multipoles
517 were translationally locked in their respective crystal sites for
518 equilibration at a relatively low temperature (50K), so that dipoles
519 or quadrupoles could freely explore all accessible orientations. The
520 translational constraints were removed, and the crystals were
521 simulated for 10 ps in the microcanonical (NVE) ensemble with an
522 average temperature of 50 K. Configurations were sampled at equal
523 time intervals for the comparison of the configurational energy
524 differences. The crystals were not simulated close to the melting
525 points in order to avoid translational deformation away of the ideal
526 lattice geometry.
527
528 For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
529 system was created with 2048 molecules oriented randomly. These were
530
531 system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
532 temperature after equilibration. We collected 250 different
533 configurations in equal interval of time. For the ions mixed liquid
534 system, we converted 48 different molecules into 24 \ce{Na+} and 24
535 \ce{Cl-} ions and equilibrated. After equilibration, the system was run
536 at the same environment for 1ns and 250 configurations were
537 collected. While comparing energies, forces, and torques with Ewald
538 method, Lennard-Jones potentials were turned off and purely
539 electrostatic interaction had been compared.
540
541 \subsection{Accuracy of Energy Differences, Forces and Torques}
542 The pairwise summation techniques (outlined above) were evaluated for
543 use in MC simulations by studying the energy differences between
544 different configurations. We took the Ewald-computed energy
545 difference between two conformations to be the correct behavior. An
546 ideal performance by one of the new methods would reproduce these
547 energy differences exactly. The configurational energies being used
548 here contain only contributions from electrostatic interactions.
549 Lennard-Jones interactions were omitted from the comparison as they
550 should be identical for all methods.
551
552 Since none of the real-space methods provide exact energy differences,
553 we used least square regressions analysiss for the six different
554 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
555 with the multipolar Ewald reference method. Unitary results for both
556 the correlation (slope) and correlation coefficient for these
557 regressions indicate perfect agreement between the real-space method
558 and the multipolar Ewald sum.
559
560 Molecular systems were run long enough to explore independent
561 configurations and 250 configurations were recorded for comparison.
562 Each system provided 31,125 energy differences for a total of 186,750
563 data points. Similarly, the magnitudes of the forces and torques have
564 also been compared by using least squares regression analyses. In the
565 forces and torques comparison, the magnitudes of the forces acting in
566 each molecule for each configuration were evaluated. For example, our
567 dipolar liquid simulation contains 2048 molecules and there are 250
568 different configurations for each system resulting in 3,072,000 data
569 points for comparison of forces and torques.
570
571 \subsection{Analysis of vector quantities}
572 Getting the magnitudes of the force and torque vectors correct is only
573 part of the issue for carrying out accurate molecular dynamics
574 simulations. Because the real space methods reweight the different
575 orientational contributions to the energies, it is also important to
576 understand how the methods impact the \textit{directionality} of the
577 force and torque vectors. Fisher developed a probablity density
578 function to analyse directional data sets,
579 \begin{equation}
580 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
581 \label{eq:pdf}
582 \end{equation}
583 where $\kappa$ measures directional dispersion of the data around the
584 mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
585 estimated as a reciprocal of the circular variance.\cite{Allen91} To
586 quantify the directional error, forces obtained from the Ewald sum
587 were taken as the mean (or correct) direction and the angle between
588 the forces obtained via the Ewald sum and the real-space methods were
589 evaluated,
590 \begin{equation}
591 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
592 \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
593 \end{equation}
594 The total angular displacement of the vectors was calculated as,
595 \begin{equation}
596 R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
597 \label{eq:displacement}
598 \end{equation}
599 where $N$ is number of force vectors. The circular variance is
600 defined as
601 \begin{equation}
602 \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
603 \end{equation}
604 The circular variance takes on values between from 0 to 1, with 0
605 indicating a perfect directional match between the Ewald force vectors
606 and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
607 correspond to higher values of $\kappa$, which indicates tighter
608 clustering of the real-space force vectors around the Ewald forces.
609
610 A similar analysis was carried out for the electrostatic contribution
611 to the molecular torques as well as forces.
612
613 \subsection{Energy conservation}
614 To test conservation the energy for the methods, the mixed molecular
615 system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
616 ions was run for 1 ns in the microcanonical ensemble at an average
617 temperature of 300K. Each of the different electrostatic methods
618 (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
619 damping values. The molecular system was started with same initial
620 positions and velocities for all cutoff methods. The energy drift
621 ($\delta E_1$) and standard deviation of the energy about the slope
622 ($\delta E_0$) were evaluated from the total energy of the system as a
623 function of time. Although both measures are valuable at
624 investigating new methods for molecular dynamics, a useful interaction
625 model must allow for long simulation times with minimal energy drift.
626
627 \section{\label{sec:result}RESULTS}
628 \subsection{Configurational energy differences}
629 %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.
630 %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.
631 %\begin{figure}[h!]
632 % \centering
633 % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
634 % \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. }
635 % \label{fig:rcutConvergence_dipolarLiquid}
636 % \end{figure}
637 %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.
638 %\begin{figure}
639 % \centering
640 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
641 % \label{fig:barGraph1}
642 % \end{figure}
643 % \begin{figure}
644 % \centering
645 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
646 % \caption{}
647
648 % \label{fig:barGraph2}
649 % \end{figure}
650 %The correlation coefficient ($R^2$) and slope of the linear regression plots for the energy differences for all six different molecular systems is shown in figure 4a and 4b.The plot shows that the correlation coefficient improves for the SP cutoff method as compared to the undamped hard cutoff method in the case of SSDQC, SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar crystal and liquid, the correlation coefficient is almost unchanged and close to 1. The correlation coefficient is smallest (0.696276 for $r_c$ = 9 $A^o$) for the SSDQC liquid because of the presence of charge-charge and charge-multipole interactions. Since the charge-charge and charge-multipole interaction is long ranged, there is huge deviation of correlation coefficient from 1. Similarly, the quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with compared to interactions in the other multipolar systems, thus the correlation coefficient very close to 1 even for hard cutoff method. The idea of placing image multipole on the surface of the cutoff sphere improves the correlation coefficient and makes it close to 1 for all types of multipolar systems. Similarly the slope is hugely deviated from the correct value for the lower order multipole-multipole interaction and slightly deviated for higher order multipole – multipole interaction. The SP method improves both correlation coefficient ($R^2$) and slope significantly in SSDQC and dipolar systems. The Slope is found to be deviated more in dipolar crystal as compared to liquid which is associated with the large fluctuation in the electrostatic energy in crystal. The GSF also produced better values of correlation coefficient and slope with the proper selection of the damping alpha (Interested reader can consult accompanying supporting material). The TSF method gives good value of correlation coefficient for the dipolar crystal, dipolar liquid, SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the regression slopes are significantly deviated.
651 \begin{figure}
652 \centering
653 \includegraphics[width=0.50 \textwidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
654 \caption{The correlation coefficient and regression slope of configurational energy differences for a given method with compared with the reference Ewald method. The value of result equal to 1(dashed line) indicates energy difference is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\ = circle, 12 \AA\ = square 15 \AA\ = inverted triangle)}
655 \label{fig:slopeCorr_energy}
656 \end{figure}
657 The combined correlation coefficient and slope for all six systems is shown in Figure ~\ref{fig:slopeCorr_energy}. The correlation coefficient for the undamped hard cutoff method is does not have good agreement with the Ewald because of the fluctuation of the electrostatic energy in the direct truncation method. This deviation in correlation coefficient is improved by using SP, GSF, and TSF method. But the TSF method worsens the regression slope stating that this method produces statistically more biased result as compared to Ewald. Also the GSF method slightly deviate slope but it can be alleviated by using proper value of damping alpha and cutoff radius. The SP method shows good agreement with Ewald method for all values of damping alpha and radii.
658 \subsection{Magnitude of the force and torque vectors}
659 The comparison of the magnitude of the combined forces and torques for the data accumulated from all system types are shown in Figure ~\ref{fig:slopeCorr_force}. The correlation and slope for the forces agree with the Ewald even for the hard cutoff method. For the system of molecules with higher order multipoles, the interaction is short ranged. Moreover, the force decays more rapidly than the electrostatic energy hence the hard cutoff method also produces good results. Although the pure cutoff gives the good match of the electrostatic force, the discontinuity in the force at the cutoff radius causes problem in the total energy conservation in MD simulations, which will be discussed in detail in subsection D. The correlation coefficient for GSF method also perfectly matches with Ewald but the slope is slightly deviated (due to extra term obtained from the angular differentiation). This deviation in the slope can be alleviated with proper selection of the damping alpha and radii ($\alpha = 0.2$ and $r_c = 12 A^o$ are good choice). The TSF method shows good agreement in the correlation coefficient but the slope is not good as compared to the Ewald.
660 \begin{figure}
661 \centering
662 \includegraphics[width=0.50 \textwidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
663 \caption{The correlation coefficient and regression slope of the magnitude of the force for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 \AA\ = circle, 12 \AA\ = square 15 \AA\ = inverted triangle). }
664 \label{fig:slopeCorr_force}
665 \end{figure}
666 The torques appears to be very influenced because of extra term generated when the potential energy is modified to get consistent force and torque. The result shows that the torque from the hard cutoff method has good agreement with Ewald. As the potential is modified to make it consistent with the force and torque, the correlation and slope is deviated as shown in Figure~\ref{fig:slopeCorr_torque} for SP, GSF and TSF cutoff methods. But the proper value of the damping alpha and radius can improve the agreement of the GSF with the Ewald method. The TSF method shows worst agreement in the slope as compared to Ewald even for larger cutoff radii.
667 \begin{figure}
668 \centering
669 \includegraphics[width=0.5 \textwidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
670 \caption{The correlation coefficient and regression slope of the magnitude of the torque for a given method with compared to the reference Ewald method. The value of result equal to 1(dashed line) indicates, the magnitude of the force from a method is indistinguishable from the Ewald method. Here different symbols represent different value of the cutoff radius (9 $A^o$ = circle, 12 $A^o$ = square 15 $A^o$ = inverted triangle).}
671 \label{fig:slopeCorr_torque}
672 \end{figure}
673 \subsection{Directionality of the force and torque vectors}
674 The accurate evaluation of the direction of the force and torques are also important for the dynamic simulation.In our research, the direction data sets were computed from the purposed method and compared with Ewald using Fisher statistics and results are expressed in terms of circular variance ($Var(\theta$).The force and torque vectors from the purposed method followed Fisher probability distribution function expressed in equation~\ref{eq:pdf}. The circular variance for the force and torque vectors of each molecule in the 250 configurations for all system types is shown in Figure~\ref{fig:slopeCorr_circularVariance}. The direction of the force and torque vectors from hard and SP cutoff methods showed best directional agreement with the Ewald. The force and torque vectors from GSF method also showed good agreement with the Ewald method, which can also be improved by varying damping alpha and cutoff radius.For $\alpha = 0.2$ and $r_c = 12 A^o$, $ Var(\theta) $ for direction of the force was found to be 0.002061 and corresponding value of $\kappa $ was 485.20. Integration of equation ~\ref{eq:pdf} for that corresponding value of $\kappa$ showed that 95\% of force vectors are with in $6.37^o$. The TSF method is the poorest in evaluating accurate direction with compared to Hard, SP, and GSF methods. The circular variance for the direction of the torques is larger as compared to force. For same $\alpha = 0.2, r_c = 12 A^o$ and GSF method, the circular variance was 0.01415, which showed 95\% of torque vectors are within $16.75^o$.The direction of the force and torque vectors can be improved by varying $\alpha$ and $r_c$.
675
676 \begin{figure}
677 \centering
678 \includegraphics[width=0.5 \textwidth]{Variance_forceNtorque_modified-crop.pdf}
679 \caption{The circular variance of the data sets of the
680 direction of the force and torque vectors obtained from a
681 given method about reference Ewald method. The result equal
682 to 0 (dashed line) indicates direction of the vectors are
683 indistinguishable from the Ewald method. Here different
684 symbols represent different value of the cutoff radius (9
685 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
686 \label{fig:slopeCorr_circularVariance}
687 \end{figure}
688 \subsection{Total energy conservation}
689 We have tested the conservation of energy in the SSDQC liquid system
690 by running system for 1ns in the Hard, SP, GSF and TSF method. The
691 Hard cutoff method shows very high energy drifts 433.53
692 KCal/Mol/ns/particle and energy fluctuation 32574.53 Kcal/Mol
693 (measured by the SD from the slope) for the undamped case, which makes
694 it completely unusable in MD simulations. The SP method also shows
695 large value of energy drift 1.289 Kcal/Mol/ns/particle and energy
696 fluctuation 0.01953 Kcal/Mol. The energy fluctuation in the SP method
697 is due to the non-vanishing nature of the torque and force at the
698 cutoff radius. We can improve the energy conservation in some extent
699 by the proper selection of the damping alpha but the improvement is
700 not good enough, which can be observed in Figure 9a and 9b .The GSF
701 and TSF shows very low value of energy drift 0.09016, 0.07371
702 KCal/Mol/ns/particle and fluctuation 3.016E-6, 1.217E-6 Kcal/Mol
703 respectively for the undamped case. Since the absolute value of the
704 evaluated electrostatic energy, force and torque from TSF method are
705 deviated from the Ewald, it does not mimic MD simulations
706 appropriately. The electrostatic energy, force and torque from the GSF
707 method have very good agreement with the Ewald. In addition, the
708 energy drift and energy fluctuation from the GSF method is much better
709 than Ewald’s method for reciprocal space vector value ($k_f$) equal to
710 7 as shown in Figure~\ref{fig:energyDrift} and
711 ~\ref{fig:fluctuation}. We can improve the total energy fluctuation
712 and drift for the Ewald’s method by increasing size of the reciprocal
713 space, which extremely increseses the simulation time. In our current
714 simulation, the simulation time for the Hard, SP, and GSF methods are
715 about 5.5 times faster than the Ewald method.
716
717 In Fig.~\ref{fig:energyDrift}, $\delta \mbox{E}_1$ is a measure of the
718 linear energy drift in units of $\mbox{kcal mol}^{-1}$ per particle
719 over a nanosecond of simulation time, and $\delta \mbox{E}_0$ is the
720 standard deviation of the energy fluctuations in units of $\mbox{kcal
721 mol}^{-1}$ per particle. In the bottom plot, it is apparent that the
722 energy drift is reduced by a significant amount (2 to 3 orders of
723 magnitude improvement at all values of the damping coefficient) by
724 chosing either of the shifted-force methods over the hard or SP
725 methods. We note that the two shifted-force method can give
726 significantly better energy conservation than the multipolar Ewald sum
727 with the same choice of real-space cutoffs.
728
729 \begin{figure}
730 \centering
731 \includegraphics[width=\textwidth]{newDrift.pdf}
732 \label{fig:energyDrift}
733 \caption{Analysis of the energy conservation of the real space
734 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
735 energy over time and $\delta \mathrm{E}_0$ is the standard deviation
736 of energy fluctuations around this drift. All simulations were of a
737 2000-molecule simulation of SSDQ water with 48 ionic charges at 300
738 K starting from the same initial configuration.}
739 \end{figure}
740
741 \section{CONCLUSION}
742 We have generalized the charged neutralized potential energy originally developed by the Wolf et al.\cite{Wolf:1999dn} for the charge-charge interaction to the charge-multipole and multipole-multipole interaction in the SP method for higher order multipoles. Also, we have developed GSF and TSF methods by implementing the modification purposed by Fennel and Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the higher order multipoles to ensure consistency and smooth truncation of the electrostatic energy, force, and torque for the spherical truncation. The SP methods for multipoles proved its suitability in MC simulations. On the other hand, the results from the GSF method produced good agreement with the Ewald's energy, force, and torque. Also, it shows very good energy conservation in MD simulations.
743 The direct truncation of any molecular system without multipole neutralization creates the fluctuation in the electrostatic energy. This fluctuation in the energy is very large for the case of crystal because of long range of multipole ordering (Refer paper I).\cite{PaperI} This is also significant in the case of the liquid because of the local multipole ordering in the molecules. If the net multipole within cutoff radius neutralized within cutoff sphere by placing image multiples on the surface of the sphere, this fluctuation in the energy reduced significantly. Also, the multipole neutralization in the generalized SP method showed very good agreement with the Ewald as compared to direct truncation for the evaluation of the $\triangle E$ between the configurations.
744 In MD simulations, the energy conservation is very important. The
745 conservation of the total energy can be ensured by i) enforcing the
746 smooth truncation of the energy, force and torque in the cutoff radius
747 and ii) making the energy, force and torque consistent with each
748 other. The GSF and TSF methods ensure the consistency and smooth
749 truncation of the energy, force and torque at the cutoff radius, as a
750 result show very good total energy conservation. But the TSF method
751 does not show good agreement in the absolute value of the
752 electrostatic energy, force and torque with the Ewald. The GSF method
753 has mimicked Ewald’s force, energy and torque accurately and also
754 conserved energy. Therefore, the GSF method is the suitable method for
755 evaluating required force field in MD simulations. In addition, the
756 energy drift and fluctuation from the GSF method is much better than
757 Ewald’s method for finite-sized reciprocal space.
758
759 Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
760 that can be made extremely efficient using spline interpolations of
761 the radial functions. They require no Fourier transforms or $k$-space
762 sums, and guarantee the smooth handling of energies, forces, and
763 torques as multipoles cross the real-space cutoff boundary.
764
765 %\bibliographystyle{aip}
766 \newpage
767 \bibliography{references}
768 \end{document}
769
770 %
771 % ****** End of file aipsamp.tex ******