ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4174
Committed: Thu Jun 5 19:55:14 2014 UTC (10 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 54384 byte(s)
Log Message:
Latest changes

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} In this work, we
481 used the multipolar Ewald sum as a reference method for comparing
482 energies, forces, and torques for molecular models that mimic
483 disordered and ordered condensed-phase systems. These test-cases
484 include:
485 \begin{itemize}
486 \item Soft Dipolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $|D| = 2.35$)
487 \item Soft Dipolar solids ($\sigma = 2.837$, $\epsilon =1.0$, $|D| = 2.35$)
488 \item Soft Quadrupolar fluids ($\sigma = 3.051$, $\epsilon =0.152$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
489 \item Soft Quadrupolar solids ($\sigma = 2.837$, $\epsilon = 1.0$, $Q_{\alpha\alpha} =\left\{-1,-1,-2.5\right\}$)
490 \item A mixed multipole model (SSDQ) for water ($\sigma = 3.051$, $\epsilon = 0.152$, $D_z = 2.35$, $Q_{\alpha\alpha} =\left\{-1.35,0,-0.68\right\}$)
491 \item A mixed multipole models for water with 48 dissolved ions, 24
492 \ce{Na+}: ($\sigma = 2.579$, $\epsilon =0.118$, $q = 1e$) and 24
493 \ce{Cl-}: ($\sigma = 4.445$, $\epsilon =0.1$l, $q = -1e$)
494 \end{itemize}
495 All Lennard-Jones parameters are in units of \AA\ $(\sigma)$ and kcal
496 / mole $(\epsilon)$. Partial charges are reported in electrons, while
497 dipoles are in Debye units, and quadrupoles are in units of Debye-\AA.
498
499 The last test case exercises all levels of the multipole-multipole
500 interactions we have derived so far and represents the most complete
501 test of the new methods. In the following section, we present results
502 for the total electrostatic energy, as well as the electrostatic
503 contributions to the force and torque on each molecule. These
504 quantities have been computed using the SP, TSF, and GSF methods, as
505 well as a hard cutoff, and have been compared with the values obtaine
506 from the multipolar Ewald sum. In Mote Carlo (MC) simulations, the
507 energy differences between two configurations is the primary quantity
508 that governs how the simulation proceeds. These differences are the
509 most imporant indicators of the reliability of a method even if the
510 absolute energies are not exact. For each of the multipolar systems
511 listed above, we have compared the change in electrostatic potential
512 energy ($\Delta E$) between 250 statistically-independent
513 configurations. In molecular dynamics (MD) simulations, the forces
514 and torques govern the behavior of the simulation, so we also compute
515 the electrostatic contributions to the forces and torques.
516
517 \subsection{Model systems}
518 To sample independent configurations of multipolar crystals, a body
519 centered cubic (bcc) crystal which is a minimum energy structure for
520 point dipoles was generated using 3,456 molecules. The multipoles
521 were translationally locked in their respective crystal sites for
522 equilibration at a relatively low temperature (50K), so that dipoles
523 or quadrupoles could freely explore all accessible orientations. The
524 translational constraints were removed, and the crystals were
525 simulated for 10 ps in the microcanonical (NVE) ensemble with an
526 average temperature of 50 K. Configurations were sampled at equal
527 time intervals for the comparison of the configurational energy
528 differences. The crystals were not simulated close to the melting
529 points in order to avoid translational deformation away of the ideal
530 lattice geometry.
531
532 For dipolar, quadrupolar, and mixed-multipole liquid simulations, each
533 system was created with 2048 molecules oriented randomly. These were
534
535 system with 2,048 molecules simulated for 1ns in NVE ensemble at 300 K
536 temperature after equilibration. We collected 250 different
537 configurations in equal interval of time. For the ions mixed liquid
538 system, we converted 48 different molecules into 24 \ce{Na+} and 24
539 \ce{Cl-} ions and equilibrated. After equilibration, the system was run
540 at the same environment for 1ns and 250 configurations were
541 collected. While comparing energies, forces, and torques with Ewald
542 method, Lennard-Jones potentials were turned off and purely
543 electrostatic interaction had been compared.
544
545 \subsection{Accuracy of Energy Differences, Forces and Torques}
546 The pairwise summation techniques (outlined above) were evaluated for
547 use in MC simulations by studying the energy differences between
548 different configurations. We took the Ewald-computed energy
549 difference between two conformations to be the correct behavior. An
550 ideal performance by one of the new methods would reproduce these
551 energy differences exactly. The configurational energies being used
552 here contain only contributions from electrostatic interactions.
553 Lennard-Jones interactions were omitted from the comparison as they
554 should be identical for all methods.
555
556 Since none of the real-space methods provide exact energy differences,
557 we used least square regressions analysiss for the six different
558 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
559 with the multipolar Ewald reference method. Unitary results for both
560 the correlation (slope) and correlation coefficient for these
561 regressions indicate perfect agreement between the real-space method
562 and the multipolar Ewald sum.
563
564 Molecular systems were run long enough to explore independent
565 configurations and 250 configurations were recorded for comparison.
566 Each system provided 31,125 energy differences for a total of 186,750
567 data points. Similarly, the magnitudes of the forces and torques have
568 also been compared by using least squares regression analyses. In the
569 forces and torques comparison, the magnitudes of the forces acting in
570 each molecule for each configuration were evaluated. For example, our
571 dipolar liquid simulation contains 2048 molecules and there are 250
572 different configurations for each system resulting in 3,072,000 data
573 points for comparison of forces and torques.
574
575 \subsection{Analysis of vector quantities}
576 Getting the magnitudes of the force and torque vectors correct is only
577 part of the issue for carrying out accurate molecular dynamics
578 simulations. Because the real space methods reweight the different
579 orientational contributions to the energies, it is also important to
580 understand how the methods impact the \textit{directionality} of the
581 force and torque vectors. Fisher developed a probablity density
582 function to analyse directional data sets,
583 \begin{equation}
584 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
585 \label{eq:pdf}
586 \end{equation}
587 where $\kappa$ measures directional dispersion of the data around the
588 mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
589 estimated as a reciprocal of the circular variance.\cite{Allen91} To
590 quantify the directional error, forces obtained from the Ewald sum
591 were taken as the mean (or correct) direction and the angle between
592 the forces obtained via the Ewald sum and the real-space methods were
593 evaluated,
594 \begin{equation}
595 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
596 \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
597 \end{equation}
598 The total angular displacement of the vectors was calculated as,
599 \begin{equation}
600 R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
601 \label{eq:displacement}
602 \end{equation}
603 where $N$ is number of force vectors. The circular variance is
604 defined as
605 \begin{equation}
606 \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
607 \end{equation}
608 The circular variance takes on values between from 0 to 1, with 0
609 indicating a perfect directional match between the Ewald force vectors
610 and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
611 correspond to higher values of $\kappa$, which indicates tighter
612 clustering of the real-space force vectors around the Ewald forces.
613
614 A similar analysis was carried out for the electrostatic contribution
615 to the molecular torques as well as forces.
616
617 \subsection{Energy conservation}
618 To test conservation the energy for the methods, the mixed molecular
619 system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
620 ions was run for 1 ns in the microcanonical ensemble at an average
621 temperature of 300K. Each of the different electrostatic methods
622 (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
623 damping values. The molecular system was started with same initial
624 positions and velocities for all cutoff methods. The energy drift
625 ($\delta E_1$) and standard deviation of the energy about the slope
626 ($\delta E_0$) were evaluated from the total energy of the system as a
627 function of time. Although both measures are valuable at
628 investigating new methods for molecular dynamics, a useful interaction
629 model must allow for long simulation times with minimal energy drift.
630
631 \section{\label{sec:result}RESULTS}
632 \subsection{Configurational energy differences}
633 %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.
634 %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.
635 %\begin{figure}[h!]
636 % \centering
637 % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
638 % \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. }
639 % \label{fig:rcutConvergence_dipolarLiquid}
640 % \end{figure}
641 %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.
642 %\begin{figure}
643 % \centering
644 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
645 % \label{fig:barGraph1}
646 % \end{figure}
647 % \begin{figure}
648 % \centering
649 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
650 % \caption{}
651
652 % \label{fig:barGraph2}
653 % \end{figure}
654 %The correlation coefficient ($R^2$) and slope of the linear
655 %regression plots for the energy differences for all six different
656 %molecular systems is shown in figure 4a and 4b.The plot shows that
657 %the correlation coefficient improves for the SP cutoff method as
658 %compared to the undamped hard cutoff method in the case of SSDQC,
659 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
660 %crystal and liquid, the correlation coefficient is almost unchanged
661 %and close to 1. The correlation coefficient is smallest (0.696276
662 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
663 %charge-charge and charge-multipole interactions. Since the
664 %charge-charge and charge-multipole interaction is long ranged, there
665 %is huge deviation of correlation coefficient from 1. Similarly, the
666 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
667 %compared to interactions in the other multipolar systems, thus the
668 %correlation coefficient very close to 1 even for hard cutoff
669 %method. The idea of placing image multipole on the surface of the
670 %cutoff sphere improves the correlation coefficient and makes it close
671 %to 1 for all types of multipolar systems. Similarly the slope is
672 %hugely deviated from the correct value for the lower order
673 %multipole-multipole interaction and slightly deviated for higher
674 %order multipole – multipole interaction. The SP method improves both
675 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
676 %dipolar systems. The Slope is found to be deviated more in dipolar
677 %crystal as compared to liquid which is associated with the large
678 %fluctuation in the electrostatic energy in crystal. The GSF also
679 %produced better values of correlation coefficient and slope with the
680 %proper selection of the damping alpha (Interested reader can consult
681 %accompanying supporting material). The TSF method gives good value of
682 %correlation coefficient for the dipolar crystal, dipolar liquid,
683 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
684 %regression slopes are significantly deviated.
685
686 \begin{figure}
687 \centering
688 \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
689 \caption{Statistical analysis of the quality of configurational
690 energy differences for the real-space electrostatic methods
691 compared with the reference Ewald sum. Results with a value equal
692 to 1 (dashed line) indicate $\Delta E$ values indistinguishable
693 from those obtained using the multipolar Ewald sum. Different
694 values of the cutoff radius are indicated with different symbols
695 (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
696 triangles).}
697 \label{fig:slopeCorr_energy}
698 \end{figure}
699
700 The combined correlation coefficient and slope for all six systems is
701 shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
702 reproduce the Ewald-derived configurational energy differences with
703 remarkable fidelity. Undamped hard cutoffs introduce a significant
704 amount of random scatter in the energy differences which is apparent
705 in the reduced value of the correlation coefficient for this method.
706 This can be understood easily as configurations which exhibit only
707 small traversals of a few dipoles or quadrupoles out of the cutoff
708 sphere will see large energy jumps when hard cutoffs are used. The
709 orientations of the multipoles (particularly in the ordered crystals)
710 mean that these jumps can go either up or down in energy, producing a
711 significant amount of random scatter.
712
713 The TSF method produces energy differences that are highly correlated
714 with the Ewald results, but it also introduces a significant
715 systematic bias in the values of the energies, particularly for
716 smaller cutoff values. The TSF method alters the distance dependence
717 of different orientational contributions to the energy in a
718 non-uniform way, so the size of the cutoff sphere can have a large
719 effect on crystalline systems.
720
721 Both the SP and GSF methods appear to reproduce the Ewald results with
722 excellent fidelity, particularly for moderate damping ($\alpha =
723 0.1-0.2$\AA$^{-1}$) and commonly-used cutoff values ($r_c = 12$\AA).
724 With the exception of the undamped hard cutoff, and the TSF method
725 with short cutoffs, all of the methods would be appropriate for use in
726 Monte Carlo simulations.
727
728 \subsection{Magnitude of the force and torque vectors}
729
730 The comparison of the magnitude of the combined forces and torques for
731 the data accumulated from all system types are shown in Figures
732 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
733 correlation and slope for the forces agree well with the Ewald sum
734 even for the hard cutoff method.
735
736 For the system of molecules with higher order multipoles, the
737 interaction is quite short ranged. Moreover, the force decays more
738 rapidly than the electrostatic energy hence the hard cutoff method can
739 also produces reasonable agreement. Although the pure cutoff gives
740 the good match of the electrostatic force for pairs of molecules
741 included within the cutoff sphere, the discontinuity in the force at
742 the cutoff radius can potentially cause problems the total energy
743 conservation as molecules enter and leave the cutoff sphere. This is
744 discussed in detail in section \ref{sec:}.
745
746 The two shifted-force methods (GSF and TSF) exhibit a small amount of
747 systematic variation and scatter compared with the Ewald forces. The
748 shifted-force models intentionally perturb the forces between pairs of
749 molecules inside the cutoff sphere in order to correct the energy
750 conservation issues, so it is not particularly surprising that this
751 perturbation is evident in these same molecular forces. The GSF
752 perturbations are minimal, particularly for moderate damping and and
753 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
754 reasonable agreement in the correlation coefficient but again the
755 systematic error in the forces is concerning if replication of Ewald
756 forces is desired.
757
758 \begin{figure}
759 \centering
760 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
761 \caption{Statistical analysis of the quality of the force vector
762 magnitudes for the real-space electrostatic methods compared with
763 the reference Ewald sum. Results with a value equal to 1 (dashed
764 line) indicate force magnitude values indistinguishable from those
765 obtained using the multipolar Ewald sum. Different values of the
766 cutoff radius are indicated with different symbols (9\AA\ =
767 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
768 \label{fig:slopeCorr_force}
769 \end{figure}
770
771
772 \begin{figure}
773 \centering
774 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
775 \caption{Statistical analysis of the quality of the torque vector
776 magnitudes for the real-space electrostatic methods compared with
777 the reference Ewald sum. Results with a value equal to 1 (dashed
778 line) indicate force magnitude values indistinguishable from those
779 obtained using the multipolar Ewald sum. Different values of the
780 cutoff radius are indicated with different symbols (9\AA\ =
781 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
782 \label{fig:slopeCorr_torque}
783 \end{figure}
784
785 The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
786 significantly influenced by the choice of real-space method. The
787 torque expressions have the same distance dependence as the energies,
788 which are naturally longer-ranged expressions than the inter-site
789 forces. Torques are also quite sensitive to orientations of
790 neighboring molecules, even those that are near the cutoff distance.
791
792 The results shows that the torque from the hard cutoff method
793 reproduces the torques in quite good agreement with the Ewald sum.
794 The other real-space methods can cause some significant deviations,
795 but excellent agreement with the Ewald sum torques is recovered at
796 moderate values of the damping coefficient ($\alpha =
797 0.1-0.2$\AA$^{-1}$) and cutoff radius ($r_c \ge 12$\AA). The TSF
798 method exhibits the only fair agreement in the slope as compared to
799 Ewald even for larger cutoff radii. It appears that the severity of
800 the perturbations in the TSF method are most apparent in the torques.
801
802 \subsection{Directionality of the force and torque vectors}
803
804 The accurate evaluation of force and torque directions is just as
805 important for molecular dynamics simulations as the magnitudes of
806 these quantities. Force and torque vectors for all six systems were
807 analyzed using Fisher statistics, and the quality of the vector
808 directionality is shown in terms of circular variance
809 ($\mathrm{Var}(\theta$) in figure
810 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
811 from the new real-space method exhibit nearly-ideal Fisher probability
812 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
813 exhibit the best vectorial agreement with the Ewald sum. The force and
814 torque vectors from GSF method also show good agreement with the Ewald
815 method, which can also be systematically improved by using moderate
816 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
817 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
818 to a distribution with 95\% of force vectors within $6.37^\circ$ of the
819 corresponding Ewald forces. The TSF method produces the poorest
820 agreement with the Ewald force directions.
821
822 Torques are again more perturbed by the new real-space methods, than
823 forces, but even here the variance is reasonably small. For the same
824 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
825 the circular variance was 0.01415, corresponds to a distribution which
826 has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
827 results. Again, the direction of the force and torque vectors can be
828 systematically improved by varying $\alpha$ and $r_c$.
829
830 \begin{figure}
831 \centering
832 \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
833 \caption{The circular variance of the direction of the force and
834 torque vectors obtained from the real-space methods around the
835 reference Ewald vectors. A variance equal to 0 (dashed line)
836 indicates direction of the force or torque vectors are
837 indistinguishable from those obtained from the Ewald sum. Here
838 different symbols represent different values of the cutoff radius
839 (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
840 \label{fig:slopeCorr_circularVariance}
841 \end{figure}
842
843 \subsection{Energy conservation}
844
845 We have tested the conservation of energy one can expect to see with
846 the new real-space methods using the SSDQ water model with a small
847 fraction of solvated ions. This is a test system which exercises all
848 orders of multipole-multipole interactions derived in the first paper
849 in this series and provides the most comprehensive test of the new
850 methods. A liquid-phase system was created with 2000 water molecules
851 and 48 dissolved ions at a density of 0.98 g cm${-3}$ and a
852 temperature of 300K. After equilibration, this liquid-phase system
853 was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
854 a cutoff radius of 9\AA. The value of the damping coefficient was
855 also varied from the undamped case ($\alpha = 0$) to a heavily damped
856 case ($\alpha = 0.3$ \AA$^{-1}$) for the real space methods. A sample
857 was also run using the multipolar Ewald sum.
858
859 In figure~\ref{fig:energyDrift} we show the both the linear drift in
860 energy over time, $\delta E_1$, and the standard deviation of energy
861 fluctuations around this drift $\delta E_0$. Both of the
862 shifted-force methods (GSF and TSF) provide excellent energy
863 conservation (drift less than $10^{-6}$ kcal / mol / ns / particle),
864 while the hard cutoff is essentially unusable for molecular dynamics.
865 SP provides some benefit over the hard cutoff because the energetic
866 jumps that happen as particles leave and enter the cutoff sphere are
867 somewhat reduced.
868
869 We note that for all tested values of the cutoff radius, the new
870 real-space methods can provide better energy conservation behavior
871 than the multipolar Ewald sum, even when utilizing a relatively large
872 $k$-space cutoff values.
873
874 \begin{figure}
875 \centering
876 \includegraphics[width=\textwidth]{newDrift.pdf}
877 \label{fig:energyDrift}
878 \caption{Analysis of the energy conservation of the real-space
879 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
880 energy over time and $\delta \mathrm{E}_0$ is the standard deviation
881 of energy fluctuations around this drift. All simulations were of a
882 2000-molecule simulation of SSDQ water with 48 ionic charges at 300
883 K starting from the same initial configuration.}
884 \end{figure}
885
886
887 \section{CONCLUSION}
888 We have generalized the charged neutralized potential energy
889 originally developed by the Wolf et al.\cite{Wolf:1999dn} for the
890 charge-charge interaction to the charge-multipole and
891 multipole-multipole interaction in the SP method for higher order
892 multipoles. Also, we have developed GSF and TSF methods by
893 implementing the modification purposed by Fennel and
894 Gezelter\cite{Fennell:2006lq} for the charge-charge interaction to the
895 higher order multipoles to ensure consistency and smooth truncation of
896 the electrostatic energy, force, and torque for the spherical
897 truncation. The SP methods for multipoles proved its suitability in MC
898 simulations. On the other hand, the results from the GSF method
899 produced good agreement with the Ewald's energy, force, and
900 torque. Also, it shows very good energy conservation in MD
901 simulations. The direct truncation of any molecular system without
902 multipole neutralization creates the fluctuation in the electrostatic
903 energy. This fluctuation in the energy is very large for the case of
904 crystal because of long range of multipole ordering (Refer paper
905 I).\cite{PaperI} This is also significant in the case of the liquid
906 because of the local multipole ordering in the molecules. If the net
907 multipole within cutoff radius neutralized within cutoff sphere by
908 placing image multiples on the surface of the sphere, this fluctuation
909 in the energy reduced significantly. Also, the multipole
910 neutralization in the generalized SP method showed very good agreement
911 with the Ewald as compared to direct truncation for the evaluation of
912 the $\triangle E$ between the configurations. In MD simulations, the
913 energy conservation is very important. The conservation of the total
914 energy can be ensured by i) enforcing the smooth truncation of the
915 energy, force and torque in the cutoff radius and ii) making the
916 energy, force and torque consistent with each other. The GSF and TSF
917 methods ensure the consistency and smooth truncation of the energy,
918 force and torque at the cutoff radius, as a result show very good
919 total energy conservation. But the TSF method does not show good
920 agreement in the absolute value of the electrostatic energy, force and
921 torque with the Ewald. The GSF method has mimicked Ewald’s force,
922 energy and torque accurately and also conserved energy. Therefore, the
923 GSF method is the suitable method for evaluating required force field
924 in MD simulations. In addition, the energy drift and fluctuation from
925 the GSF method is much better than Ewald’s method for finite-sized
926 reciprocal space.
927
928 Note that the TSF, GSF, and SP models are $\mathcal{O}(N)$ methods
929 that can be made extremely efficient using spline interpolations of
930 the radial functions. They require no Fourier transforms or $k$-space
931 sums, and guarantee the smooth handling of energies, forces, and
932 torques as multipoles cross the real-space cutoff boundary.
933
934 %\bibliographystyle{aip}
935 \newpage
936 \bibliography{references}
937 \end{document}
938
939 %
940 % ****** End of file aipsamp.tex ******