ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
Revision: 4181
Committed: Thu Jun 12 14:58:06 2014 UTC (10 years, 3 months ago) by gezelter
Content type: application/x-tex
File size: 62477 byte(s)
Log Message:
Edits

File Contents

# Content
1 % ****** Start of file aipsamp.tex ******
2 %
3 % This file is part of the AIP files in the AIP distribution for REVTeX 4.
4 % Version 4.1 of REVTeX, October 2009
5 %
6 % Copyright (c) 2009 American Institute of Physics.
7 %
8 % See the AIP README file for restrictions and more information.
9 %
10 % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
11 % as well as the rest of the prerequisites for REVTeX 4.1
12 %
13 % It also requires running BibTeX. The commands are as follows:
14 %
15 % 1) latex aipsamp
16 % 2) bibtex aipsamp
17 % 3) latex aipsamp
18 % 4) latex aipsamp
19 %
20 % Use this file as a source of example code for your aip document.
21 % Use the file aiptemplate.tex as a template for your document.
22 \documentclass[%
23 aip,jcp,
24 amsmath,amssymb,
25 preprint,
26 %reprint,%
27 %author-year,%
28 %author-numerical,%
29 ]{revtex4-1}
30
31 \usepackage{graphicx}% Include figure files
32 \usepackage{dcolumn}% Align table columns on decimal point
33 %\usepackage{bm}% bold math
34 %\usepackage[mathlines]{lineno}% Enable numbering of text and display math
35 %\linenumbers\relax % Commence numbering lines
36 \usepackage{amsmath}
37 \usepackage{times}
38 \usepackage{mathptmx}
39 \usepackage{tabularx}
40 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
41 \usepackage{url}
42 \usepackage[english]{babel}
43
44 \newcolumntype{Y}{>{\centering\arraybackslash}X}
45
46 \begin{document}
47
48 %\preprint{AIP/123-QED}
49
50 \title{Real space alternatives to the Ewald
51 Sum. II. Comparison of Methods} % Force line breaks with \\
52
53 \author{Madan Lamichhane}
54 \affiliation{Department of Physics, University
55 of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
56
57 \author{Kathie E. Newman}
58 \affiliation{Department of Physics, University
59 of Notre Dame, Notre Dame, IN 46556}
60
61 \author{J. Daniel Gezelter}%
62 \email{gezelter@nd.edu.}
63 \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
64 }%
65
66 \date{\today}% It is always \today, today,
67 % but any date may be explicitly specified
68
69 \begin{abstract}
70 We have tested the real-space shifted potential (SP),
71 gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
72 for multipoles that were developed in the first paper in this series
73 against a reference method. The tests were carried out in a variety
74 of condensed-phase environments which were designed to test all
75 levels of the multipole-multipole interactions. Comparisons of the
76 energy differences between configurations, molecular forces, and
77 torques were used to analyze how well the real-space models perform
78 relative to the more computationally expensive Ewald sum. We have
79 also investigated the energy conservation properties of the new
80 methods in molecular dynamics simulations using all of these
81 methods. The SP method shows excellent agreement with
82 configurational energy differences, forces, and torques, and would
83 be suitable for use in Monte Carlo calculations. Of the two new
84 shifted-force methods, the GSF approach shows the best agreement
85 with Ewald-derived energies, forces, and torques and exhibits energy
86 conservation properties that make it an excellent choice for
87 efficiently computing electrostatic interactions in molecular
88 dynamics simulations.
89 \end{abstract}
90
91 %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
92 % Classification Scheme.
93 \keywords{Electrostatics, Multipoles, Real-space}
94
95 \maketitle
96
97
98 \section{\label{sec:intro}Introduction}
99 Computing the interactions between electrostatic sites is one of the
100 most expensive aspects of molecular simulations, which is why there
101 have been significant efforts to develop practical, efficient and
102 convergent methods for handling these interactions. Ewald's method is
103 perhaps the best known and most accurate method for evaluating
104 energies, forces, and torques in explicitly-periodic simulation
105 cells. In this approach, the conditionally convergent electrostatic
106 energy is converted into two absolutely convergent contributions, one
107 which is carried out in real space with a cutoff radius, and one in
108 reciprocal space.\cite{Clarke:1986eu,Woodcock75}
109
110 When carried out as originally formulated, the reciprocal-space
111 portion of the Ewald sum exhibits relatively poor computational
112 scaling, making it prohibitive for large systems. By utilizing
113 particle meshes and three dimensional fast Fourier transforms (FFT),
114 the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
115 (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
116 the computational cost from $O(N^2)$ down to $O(N \log
117 N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
118
119 Because of the artificial periodicity required for the Ewald sum, the
120 method may require modification to compute interactions for
121 interfacial molecular systems such as membranes and liquid-vapor
122 interfaces.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
123 To simulate interfacial systems, Parry's extension of the 3D Ewald sum
124 is appropriate for slab geometries.\cite{Parry:1975if} The inherent
125 periodicity in the Ewald’s method can also be problematic for
126 interfacial molecular systems.\cite{Fennell:2006lq} Modified Ewald
127 methods that were developed to handle two-dimensional (2D)
128 electrostatic interactions in interfacial systems have not had similar
129 particle-mesh treatments.\cite{Parry:1975if, Parry:1976fq, Clarke77,
130 Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq}
131
132 \subsection{Real-space methods}
133 Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
134 method for calculating electrostatic interactions between point
135 charges. They argued that the effective Coulomb interaction in
136 condensed systems is actually short ranged.\cite{Wolf92,Wolf95} For an
137 ordered lattice (e.g., when computing the Madelung constant of an
138 ionic solid), the material can be considered as a set of ions
139 interacting with neutral dipolar or quadrupolar ``molecules'' giving
140 an effective distance dependence for the electrostatic interactions of
141 $r^{-5}$ (see figure \ref{fig:schematic}). For this reason, careful
142 applications of Wolf's method are able to obtain accurate estimates of
143 Madelung constants using relatively short cutoff radii. Recently,
144 Fukuda used neutralization of the higher order moments for the
145 calculation of the electrostatic interaction of the point charges
146 system.\cite{Fukuda:2013sf}
147
148 \begin{figure}
149 \centering
150 \includegraphics[width=\linewidth]{schematic.pdf}
151 \caption{Top: Ionic systems exhibit local clustering of dissimilar
152 charges (in the smaller grey circle), so interactions are
153 effectively charge-multipole in order at longer distances. With
154 hard cutoffs, motion of individual charges in and out of the
155 cutoff sphere can break the effective multipolar ordering.
156 Bottom: dipolar crystals and fluids have a similar effective
157 \textit{quadrupolar} ordering (in the smaller grey circles), and
158 orientational averaging helps to reduce the effective range of the
159 interactions in the fluid. Placement of reversed image multipoles
160 on the surface of the cutoff sphere recovers the effective
161 higher-order multipole behavior.}
162 \label{fig:schematic}
163 \end{figure}
164
165 The direct truncation of interactions at a cutoff radius creates
166 truncation defects. Wolf \textit{et al.} further argued that
167 truncation errors are due to net charge remaining inside the cutoff
168 sphere.\cite{Wolf:1999dn} To neutralize this charge they proposed
169 placing an image charge on the surface of the cutoff sphere for every
170 real charge inside the cutoff. These charges are present for the
171 evaluation of both the pair interaction energy and the force, although
172 the force expression maintained a discontinuity at the cutoff sphere.
173 In the original Wolf formulation, the total energy for the charge and
174 image were not equal to the integral of their force expression, and as
175 a result, the total energy would not be conserved in molecular
176 dynamics (MD) simulations.\cite{Zahn:2002hc} Zahn \textit{et al.} and
177 Fennel and Gezelter later proposed shifted force variants of the Wolf
178 method with commensurate force and energy expressions that do not
179 exhibit this problem.\cite{Fennell:2006lq} Related real-space
180 methods were also proposed by Chen \textit{et
181 al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
182 and by Wu and Brooks.\cite{Wu:044107}
183
184 Considering the interaction of one central ion in an ionic crystal
185 with a portion of the crystal at some distance, the effective Columbic
186 potential is found to decrease as $r^{-5}$. If one views the \ce{NaCl}
187 crystal as a simple cubic (SC) structure with an octupolar
188 \ce{(NaCl)4} basis, the electrostatic energy per ion converges more
189 rapidly to the Madelung energy than the dipolar
190 approximation.\cite{Wolf92} To find the correct Madelung constant,
191 Lacman suggested that the NaCl structure could be constructed in a way
192 that the finite crystal terminates with complete \ce{(NaCl)4}
193 molecules.\cite{Lacman65} The central ion sees what is effectively a
194 set of octupoles at large distances. These facts suggest that the
195 Madelung constants are relatively short ranged for perfect ionic
196 crystals.\cite{Wolf:1999dn}
197
198 One can make a similar argument for crystals of point multipoles. The
199 Luttinger and Tisza treatment of energy constants for dipolar lattices
200 utilizes 24 basis vectors that contain dipoles at the eight corners of
201 a unit cube. Only three of these basis vectors, $X_1, Y_1,
202 \mathrm{~and~} Z_1,$ retain net dipole moments, while the rest have
203 zero net dipole and retain contributions only from higher order
204 multipoles. The effective interaction between a dipole at the center
205 of a crystal and a group of eight dipoles farther away is
206 significantly shorter ranged than the $r^{-3}$ that one would expect
207 for raw dipole-dipole interactions. Only in crystals which retain a
208 bulk dipole moment (e.g. ferroelectrics) does the analogy with the
209 ionic crystal break down -- ferroelectric dipolar crystals can exist,
210 while ionic crystals with net charge in each unit cell would be
211 unstable.
212
213 In ionic crystals, real-space truncation can break the effective
214 multipolar arrangements (see Fig. \ref{fig:schematic}), causing
215 significant swings in the electrostatic energy as individual ions move
216 back and forth across the boundary. This is why the image charges are
217 necessary for the Wolf sum to exhibit rapid convergence. Similarly,
218 the real-space truncation of point multipole interactions breaks
219 higher order multipole arrangements, and image multipoles are required
220 for real-space treatments of electrostatic energies.
221
222 The shorter effective range of electrostatic interactions is not
223 limited to perfect crystals, but can also apply in disordered fluids.
224 Even at elevated temperatures, there is, on average, local charge
225 balance in an ionic liquid, where each positive ion has surroundings
226 dominated by negaitve ions and vice versa. The reversed-charge images
227 on the cutoff sphere that are integral to the Wolf and DSF approaches
228 retain the effective multipolar interactions as the charges traverse
229 the cutoff boundary.
230
231 In multipolar fluids (see Fig. \ref{fig:schematic}) there is
232 significant orientational averaging that additionally reduces the
233 effect of long-range multipolar interactions. The image multipoles
234 that are introduced in the TSF, GSF, and SP methods mimic this effect
235 and reduce the effective range of the multipolar interactions as
236 interacting molecules traverse each other's cutoff boundaries.
237
238 % Because of this reason, although the nature of electrostatic
239 % interaction short ranged, the hard cutoff sphere creates very large
240 % fluctuation in the electrostatic energy for the perfect crystal. In
241 % addition, the charge neutralized potential proposed by Wolf et
242 % al. converged to correct Madelung constant but still holds oscillation
243 % in the energy about correct Madelung energy.\cite{Wolf:1999dn}. This
244 % oscillation in the energy around its fully converged value can be due
245 % to the non-neutralized value of the higher order moments within the
246 % cutoff sphere.
247
248 The forces and torques acting on atomic sites are the fundamental
249 factors driving dynamics in molecular simulations. Fennell and
250 Gezelter proposed the damped shifted force (DSF) energy kernel to
251 obtain consistent energies and forces on the atoms within the cutoff
252 sphere. Both the energy and the force go smoothly to zero as an atom
253 aproaches the cutoff radius. The comparisons of the accuracy these
254 quantities between the DSF kernel and SPME was surprisingly
255 good.\cite{Fennell:2006lq} The DSF method has seen increasing use for
256 calculating electrostatic interactions in molecular systems with
257 relatively uniform charge
258 densities.\cite{Shi:2013ij,Kannam:2012rr,Acevedo13,Space12,English08,Lawrence13,Vergne13}
259
260 \subsection{The damping function}
261 The damping function used in our research has been discussed in detail
262 in the first paper of this series.\cite{PaperI} The radial kernel
263 $1/r$ for the interactions between point charges can be replaced by
264 the complementary error function $\mathrm{erfc}(\alpha r)/r$ to
265 accelerate the rate of convergence, where $\alpha$ is a damping
266 parameter with units of inverse distance. Altering the value of
267 $\alpha$ is equivalent to changing the width of Gaussian charge
268 distributions that replace each point charge -- Gaussian overlap
269 integrals yield complementary error functions when truncated at a
270 finite distance.
271
272 By using suitable value of damping alpha ($\alpha \sim 0.2$) for a
273 cutoff radius ($r_{­c}=9 A$), Fennel and Gezelter produced very good
274 agreement with SPME for the interaction energies, forces and torques
275 for charge-charge interactions.\cite{Fennell:2006lq}
276
277 \subsection{Point multipoles in molecular modeling}
278 Coarse-graining approaches which treat entire molecular subsystems as
279 a single rigid body are now widely used. A common feature of many
280 coarse-graining approaches is simplification of the electrostatic
281 interactions between bodies so that fewer site-site interactions are
282 required to compute configurational energies. Many coarse-grained
283 molecular structures would normally consist of equal positive and
284 negative charges, and rather than use multiple site-site interactions,
285 the interaction between higher order multipoles can also be used to
286 evaluate a single molecule-molecule
287 interaction.\cite{Ren06,Essex10,Essex11}
288
289 Because electrons in a molecule are not localized at specific points,
290 the assignment of partial charges to atomic centers is a relatively
291 rough approximation. Atomic sites can also be assigned point
292 multipoles and polarizabilities to increase the accuracy of the
293 molecular model. Recently, water has been modeled with point
294 multipoles up to octupolar order using the soft sticky
295 dipole-quadrupole-octupole (SSDQO)
296 model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
297 multipoles up to quadrupolar order have also been coupled with point
298 polarizabilities in the high-quality AMOEBA and iAMOEBA water
299 models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} But
300 using point multipole with the real space truncation without
301 accounting for multipolar neutrality will create energy conservation
302 issues in molecular dynamics (MD) simulations.
303
304 In this paper we test a set of real-space methods that were developed
305 for point multipolar interactions. These methods extend the damped
306 shifted force (DSF) and Wolf methods originally developed for
307 charge-charge interactions and generalize them for higher order
308 multipoles. The detailed mathematical development of these methods has
309 been presented in the first paper in this series, while this work
310 covers the testing the energies, forces, torques, and energy
311 conservation properties of the methods in realistic simulation
312 environments. In all cases, the methods are compared with the
313 reference method, a full multipolar Ewald treatment.
314
315
316 %\subsection{Conservation of total energy }
317 %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.
318
319 \section{\label{sec:method}Review of Methods}
320 Any real-space electrostatic method that is suitable for MD
321 simulations should have the electrostatic energy, forces and torques
322 between two sites go smoothly to zero as the distance between the
323 sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$. Requiring
324 this continuity at the cutoff is essential for energy conservation in
325 MD simulations. The mathematical details of the shifted potential
326 (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
327 methods have been discussed in detail in the previous paper in this
328 series.\cite{PaperI} Here we briefly review the new methods and
329 describe their essential features.
330
331 \subsection{Taylor-shifted force (TSF)}
332
333 The electrostatic potential energy between point multipoles can be
334 expressed as the product of two multipole operators and a Coulombic
335 kernel,
336 \begin{equation}
337 U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r} \label{kernel}.
338 \end{equation}
339 where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
340 expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
341 a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
342 $\bf a$.
343
344 % Interactions between multipoles can be expressed as higher derivatives
345 % of the bare Coulomb potential, so one way of ensuring that the forces
346 % and torques vanish at the cutoff distance is to include a larger
347 % number of terms in the truncated Taylor expansion, e.g.,
348 % %
349 % \begin{equation}
350 % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert _{r_c} .
351 % \end{equation}
352 % %
353 % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
354 % Thus, for $f(r)=1/r$, we find
355 % %
356 % \begin{equation}
357 % 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} .
358 % \end{equation}
359 % This function is an approximate electrostatic potential that has
360 % vanishing second derivatives at the cutoff radius, making it suitable
361 % for shifting the forces and torques of charge-dipole interactions.
362
363 The TSF potential for any multipole-multipole interaction can be
364 written
365 \begin{equation}
366 U^{\text{TSF}}= (\text{prefactor}) (\text{derivatives}) f_n(r)
367 \label{generic}
368 \end{equation}
369 where $f_n(r)$ is a shifted kernel that is appropriate for the order
370 of the interaction (see Ref. \onlinecite{PaperI}), with $n=0$ for
371 charge-charge, $n=1$ for charge-dipole, $n=2$ for charge-quadrupole
372 and dipole-dipole, $n=3$ for dipole-quadrupole, and $n=4$ for
373 quadrupole-quadrupole. To ensure smooth convergence of the energy,
374 force, and torques, a Taylor expansion with $n$ terms must be
375 performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
376
377 % To carry out the same procedure for a damped electrostatic kernel, we
378 % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
379 % Many of the derivatives of the damped kernel are well known from
380 % Smith's early work on multipoles for the Ewald
381 % summation.\cite{Smith82,Smith98}
382
383 % Note that increasing the value of $n$ will add additional terms to the
384 % electrostatic potential, e.g., $f_2(r)$ includes orders up to
385 % $(r-r_c)^3/r_c^4$, and so on. Successive derivatives of the $f_n(r)$
386 % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
387 % f^{\prime\prime}_2(r)$, etc. These higher derivatives are required
388 % for computing multipole energies, forces, and torques, and smooth
389 % cutoffs of these quantities can be guaranteed as long as the number of
390 % terms in the Taylor series exceeds the derivative order required.
391
392 For multipole-multipole interactions, following this procedure results
393 in separate radial functions for each of the distinct orientational
394 contributions to the potential, and ensures that the forces and
395 torques from each of these contributions will vanish at the cutoff
396 radius. For example, the direct dipole dot product
397 ($\mathbf{D}_{\bf a}
398 \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
399 dot products:
400 \begin{equation}
401 U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
402 \mathbf{D}_{\bf a} \cdot
403 \mathbf{D}_{\bf b} \right) v_{21}(r) +
404 \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
405 \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
406 \end{equation}
407
408 For the Taylor shifted (TSF) method with the undamped kernel,
409 $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} - \frac{8}{r_c^3} +
410 \frac{6}{r r_c^2}$ and $v_{22}(r) = \frac{3}{r^3} + \frac{3 r}{r_c^4}
411 - \frac{6}{r r_c^2}$. In these functions, one can easily see the
412 connection to unmodified electrostatics as well as the smooth
413 transition to zero in both these functions as $r\rightarrow r_c$. The
414 electrostatic forces and torques acting on the central multipole due
415 to another site within cutoff sphere are derived from
416 Eq.~\ref{generic}, accounting for the appropriate number of
417 derivatives. Complete energy, force, and torque expressions are
418 presented in the first paper in this series (Reference
419 \onlinecite{PaperI}).
420
421 \subsection{Gradient-shifted force (GSF)}
422
423 A second (and conceptually simpler) method involves shifting the
424 gradient of the raw Coulomb potential for each particular multipole
425 order. For example, the raw dipole-dipole potential energy may be
426 shifted smoothly by finding the gradient for two interacting dipoles
427 which have been projected onto the surface of the cutoff sphere
428 without changing their relative orientation,
429 \begin{equation}
430 U_{D_{\bf a}D_{\bf b}}(r) = U_{D_{\bf a}D_{\bf b}}(r) -
431 U_{D_{\bf a} D_{\bf b}}(r_c)
432 - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
433 \vec{\nabla} U_{D_{\bf a}D_{\bf b}}(r) \Big \lvert _{r_c}
434 \end{equation}
435 Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
436 a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
437 (although the signs are reversed for the dipole that has been
438 projected onto the cutoff sphere). In many ways, this simpler
439 approach is closer in spirit to the original shifted force method, in
440 that it projects a neutralizing multipole (and the resulting forces
441 from this multipole) onto a cutoff sphere. The resulting functional
442 forms for the potentials, forces, and torques turn out to be quite
443 similar in form to the Taylor-shifted approach, although the radial
444 contributions are significantly less perturbed by the gradient-shifted
445 approach than they are in the Taylor-shifted method.
446
447 For the gradient shifted (GSF) method with the undamped kernel,
448 $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
449 $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
450 Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
451 because the Taylor expansion retains only one term, they are
452 significantly less perturbed than the TSF functions.
453
454 In general, the gradient shifted potential between a central multipole
455 and any multipolar site inside the cutoff radius is given by,
456 \begin{equation}
457 U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
458 U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{r}
459 \cdot \nabla U(\mathbf{r},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \Big \lvert _{r_c} \right]
460 \label{generic2}
461 \end{equation}
462 where the sum describes a separate force-shifting that is applied to
463 each orientational contribution to the energy.
464
465 The third term converges more rapidly than the first two terms as a
466 function of radius, hence the contribution of the third term is very
467 small for large cutoff radii. The force and torque derived from
468 equation \ref{generic2} are consistent with the energy expression and
469 approach zero as $r \rightarrow r_c$. Both the GSF and TSF methods
470 can be considered generalizations of the original DSF method for
471 higher order multipole interactions. GSF and TSF are also identical up
472 to the charge-dipole interaction but generate different expressions in
473 the energy, force and torque for higher order multipole-multipole
474 interactions. Complete energy, force, and torque expressions for the
475 GSF potential are presented in the first paper in this series
476 (Reference~\onlinecite{PaperI})
477
478
479 \subsection{Shifted potential (SP) }
480 A discontinuous truncation of the electrostatic potential at the
481 cutoff sphere introduces a severe artifact (oscillation in the
482 electrostatic energy) even for molecules with the higher-order
483 multipoles.\cite{PaperI} We have also formulated an extension of the
484 Wolf approach for point multipoles by simply projecting the image
485 multipole onto the surface of the cutoff sphere, and including the
486 interactions with the central multipole and the image. This
487 effectively shifts the total potential to zero at the cutoff radius,
488 \begin{equation}
489 U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
490 U(\mathbf{r}_c,\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
491 \label{eq:SP}
492 \end{equation}
493 where the sum describes separate potential shifting that is done for
494 each orientational contribution to the energy (e.g. the direct dipole
495 product contribution is shifted {\it separately} from the
496 dipole-distance terms in dipole-dipole interactions). Note that this
497 is not a simple shifting of the total potential at $r_c$. Each radial
498 contribution is shifted separately. One consequence of this is that
499 multipoles that reorient after leaving the cutoff sphere can re-enter
500 the cutoff sphere without perturbing the total energy.
501
502 For the shifted potential (SP) method with the undamped kernel,
503 $v_{21}(r) = -\frac{1}{r^3} + \frac{1}{r_c^3}$ and $v_{22}(r) =
504 \frac{3}{r^3} - \frac{3}{r_c^3}$. The potential energy between a
505 central multipole and other multipolar sites goes smoothly to zero as
506 $r \rightarrow r_c$. However, the force and torque obtained from the
507 shifted potential (SP) are discontinuous at $r_c$. MD simulations
508 will still experience energy drift while operating under the SP
509 potential, but it may be suitable for Monte Carlo approaches where the
510 configurational energy differences are the primary quantity of
511 interest.
512
513 \subsection{The Self Term}
514 In the TSF, GSF, and SP methods, a self-interaction is retained for
515 the central multipole interacting with its own image on the surface of
516 the cutoff sphere. This self interaction is nearly identical with the
517 self-terms that arise in the Ewald sum for multipoles. Complete
518 expressions for the self terms are presented in the first paper in
519 this series (Reference \onlinecite{PaperI}).
520
521
522 \section{\label{sec:methodology}Methodology}
523
524 To understand how the real-space multipole methods behave in computer
525 simulations, it is vital to test against established methods for
526 computing electrostatic interactions in periodic systems, and to
527 evaluate the size and sources of any errors that arise from the
528 real-space cutoffs. In the first paper of this series, we compared
529 the dipolar and quadrupolar energy expressions against analytic
530 expressions for ordered dipolar and quadrupolar
531 arrays.\cite{Sauer,LT,Nagai01081960,Nagai01091963} In this work, we
532 used the multipolar Ewald sum as a reference method for comparing
533 energies, forces, and torques for molecular models that mimic
534 disordered and ordered condensed-phase systems. The parameters used
535 in the test cases are given in table~\ref{tab:pars}.
536
537 \begin{table}
538 \label{tab:pars}
539 \caption{The parameters used in the systems used to evaluate the new
540 real-space methods. The most comprehensive test was a liquid
541 composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
542 ions). This test excercises all orders of the multipolar
543 interactions developed in the first paper.}
544 \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
545 & \multicolumn{2}{c|}{LJ parameters} &
546 \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
547 Test system & $\sigma$& $\epsilon$ & $C$ & $D$ &
548 $Q_{xx}$ & $Q_{yy}$ & $Q_{zz}$ & mass & $I_{xx}$ & $I_{yy}$ &
549 $I_{zz}$ \\ \cline{6-8}\cline{10-12}
550 & (\AA) & (kcal/mol) & (e) & (debye) & \multicolumn{3}{c|}{(debye \AA)} & (amu) & \multicolumn{3}{c}{(amu
551 \AA\textsuperscript{2})} \\ \hline
552 Soft Dipolar fluid & 3.051 & 0.152 & & 2.35 & & & & 18.0153 & 1.77&0.6145& 1.155 \\
553 Soft Dipolar solid & 2.837 & 1.0 & & 2.35 & & & & $10^4$ & 17.6 &17.6 & 0 \\
554 Soft Quadrupolar fluid & 3.051 & 0.152 & & & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155 \\
555 Soft Quadrupolar solid & 2.837 & 1.0 & & & -1&-1&-2.5 & $10^4$ & 17.6&17.6&0 \\
556 SSDQ water & 3.051 & 0.152 & & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
557 \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
558 \ce{Cl-} & 4.445 & 0.1 & -1& & & & & 35.4527& & & \\ \hline
559 \end{tabularx}
560 \end{table}
561 The systems consist of pure multipolar solids (both dipole and
562 quadrupole), pure multipolar liquids (both dipole and quadrupole), a
563 fluid composed of sites containing both dipoles and quadrupoles
564 simultaneously, and a final test case that includes ions with point
565 charges in addition to the multipolar fluid. The solid-phase
566 parameters were chosen so that the systems can explore some
567 orientational freedom for the multipolar sites, while maintaining
568 relatively strict translational order. The SSDQ model used here is
569 not a particularly accurate water model, but it does test
570 dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
571 interactions at roughly the same magnitudes. The last test case, SSDQ
572 water with dissolved ions, exercises \textit{all} levels of the
573 multipole-multipole interactions we have derived so far and represents
574 the most complete test of the new methods.
575
576 In the following section, we present results for the total
577 electrostatic energy, as well as the electrostatic contributions to
578 the force and torque on each molecule. These quantities have been
579 computed using the SP, TSF, and GSF methods, as well as a hard cutoff,
580 and have been compared with the values obtained from the multipolar
581 Ewald sum. In Monte Carlo (MC) simulations, the energy differences
582 between two configurations is the primary quantity that governs how
583 the simulation proceeds. These differences are the most imporant
584 indicators of the reliability of a method even if the absolute
585 energies are not exact. For each of the multipolar systems listed
586 above, we have compared the change in electrostatic potential energy
587 ($\Delta E$) between 250 statistically-independent configurations. In
588 molecular dynamics (MD) simulations, the forces and torques govern the
589 behavior of the simulation, so we also compute the electrostatic
590 contributions to the forces and torques.
591
592 \subsection{Implementation}
593 The real-space methods developed in the first paper in this series
594 have been implemented in our group's open source molecular simulation
595 program, OpenMD,\cite{openmd} which was used for all calculations in
596 this work. The complementary error function can be a relatively slow
597 function on some processors, so all of the radial functions are
598 precomputed on a fine grid and are spline-interpolated to provide
599 values when required.
600
601 Using the same simulation code, we compare to a multipolar Ewald sum
602 with a reciprocal space cutoff, $k_\mathrm{max} = 7$. Our version of
603 the Ewald sum is a re-implementation of the algorithm originally
604 proposed by Smith that does not use the particle mesh or smoothing
605 approximations.\cite{Smith82,Smith98} In all cases, the quantities
606 being compared are the electrostatic contributions to energies, force,
607 and torques. All other contributions to these quantities (i.e. from
608 Lennard-Jones interactions) are removed prior to the comparisons.
609
610 The convergence parameter ($\alpha$) also plays a role in the balance
611 of the real-space and reciprocal-space portions of the Ewald
612 calculation. Typical molecular mechanics packages set this to a value
613 that depends on the cutoff radius and a tolerance (typically less than
614 $1 \times 10^{-4}$ kcal/mol). Smaller tolerances are typically
615 associated with increasing accuracy at the expense of computational
616 time spent on the reciprocal-space portion of the
617 summation.\cite{Perram88,Essmann:1995pb} A default tolerance of $1 \times
618 10^{-8}$ kcal/mol was used in all Ewald calculations, resulting in
619 Ewald coefficient 0.3119 \AA$^{-1}$ for a cutoff radius of 12 \AA.
620
621 The real-space models have self-interactions that provide
622 contributions to the energies only. Although the self interaction is
623 a rapid calculation, we note that in systems with fluctuating charges
624 or point polarizabilities, the self-term is not static and must be
625 recomputed at each time step.
626
627 \subsection{Model systems}
628 To sample independent configurations of the multipolar crystals, body
629 centered cubic (bcc) crystals, which exhibit the minimum energy
630 structures for point dipoles, were generated using 3,456 molecules.
631 The multipoles were translationally locked in their respective crystal
632 sites for equilibration at a relatively low temperature (50K) so that
633 dipoles or quadrupoles could freely explore all accessible
634 orientations. The translational constraints were then removed, the
635 systems were re-equilibrated, and the crystals were simulated for an
636 additional 10 ps in the microcanonical (NVE) ensemble with an average
637 temperature of 50 K. The balance between moments of inertia and
638 particle mass were chosen to allow orientational sampling without
639 significant translational motion. Configurations were sampled at
640 equal time intervals in order to compare configurational energy
641 differences. The crystals were simulated far from the melting point
642 in order to avoid translational deformation away of the ideal lattice
643 geometry.
644
645 For dipolar, quadrupolar, and mixed-multipole \textit{liquid}
646 simulations, each system was created with 2,048 randomly-oriented
647 molecules. These were equilibrated at a temperature of 300K for 1 ns.
648 Each system was then simulated for 1 ns in the microcanonical (NVE)
649 ensemble. We collected 250 different configurations at equal time
650 intervals. For the liquid system that included ionic species, we
651 converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
652 \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
653 run under the same conditions for 1 ns. A total of 250 configurations
654 were collected. In the following comparisons of energies, forces, and
655 torques, the Lennard-Jones potentials were turned off and only the
656 purely electrostatic quantities were compared with the same values
657 obtained via the Ewald sum.
658
659 \subsection{Accuracy of Energy Differences, Forces and Torques}
660 The pairwise summation techniques (outlined above) were evaluated for
661 use in MC simulations by studying the energy differences between
662 different configurations. We took the Ewald-computed energy
663 difference between two conformations to be the correct behavior. An
664 ideal performance by one of the new methods would reproduce these
665 energy differences exactly. The configurational energies being used
666 here contain only contributions from electrostatic interactions.
667 Lennard-Jones interactions were omitted from the comparison as they
668 should be identical for all methods.
669
670 Since none of the real-space methods provide exact energy differences,
671 we used least square regressions analysis for the six different
672 molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
673 with the multipolar Ewald reference method. Unitary results for both
674 the correlation (slope) and correlation coefficient for these
675 regressions indicate perfect agreement between the real-space method
676 and the multipolar Ewald sum.
677
678 Molecular systems were run long enough to explore independent
679 configurations and 250 configurations were recorded for comparison.
680 Each system provided 31,125 energy differences for a total of 186,750
681 data points. Similarly, the magnitudes of the forces and torques have
682 also been compared using least squares regression analysis. In the
683 forces and torques comparison, the magnitudes of the forces acting in
684 each molecule for each configuration were evaluated. For example, our
685 dipolar liquid simulation contains 2048 molecules and there are 250
686 different configurations for each system resulting in 3,072,000 data
687 points for comparison of forces and torques.
688
689 \subsection{Analysis of vector quantities}
690 Getting the magnitudes of the force and torque vectors correct is only
691 part of the issue for carrying out accurate molecular dynamics
692 simulations. Because the real space methods reweight the different
693 orientational contributions to the energies, it is also important to
694 understand how the methods impact the \textit{directionality} of the
695 force and torque vectors. Fisher developed a probablity density
696 function to analyse directional data sets,
697 \begin{equation}
698 p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
699 \label{eq:pdf}
700 \end{equation}
701 where $\kappa$ measures directional dispersion of the data around the
702 mean direction.\cite{fisher53} This quantity $(\kappa)$ can be
703 estimated as a reciprocal of the circular variance.\cite{Allen91} To
704 quantify the directional error, forces obtained from the Ewald sum
705 were taken as the mean (or correct) direction and the angle between
706 the forces obtained via the Ewald sum and the real-space methods were
707 evaluated,
708 \begin{equation}
709 \cos\theta_i = \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
710 \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
711 \end{equation}
712 The total angular displacement of the vectors was calculated as,
713 \begin{equation}
714 R = \sqrt{\left(\sum\limits_{i=1}^N \cos\theta_i\right)^2 + \left(\sum\limits_{i=1}^N \sin\theta_i\right)^2}
715 \label{eq:displacement}
716 \end{equation}
717 where $N$ is number of force vectors. The circular variance is
718 defined as
719 \begin{equation}
720 \mathrm{Var}(\theta) \approx 1/\kappa = 1 - R/N
721 \end{equation}
722 The circular variance takes on values between from 0 to 1, with 0
723 indicating a perfect directional match between the Ewald force vectors
724 and the real-space forces. Lower values of $\mathrm{Var}(\theta)$
725 correspond to higher values of $\kappa$, which indicates tighter
726 clustering of the real-space force vectors around the Ewald forces.
727
728 A similar analysis was carried out for the electrostatic contribution
729 to the molecular torques as well as forces.
730
731 \subsection{Energy conservation}
732 To test conservation the energy for the methods, the mixed molecular
733 system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
734 ions was run for 1 ns in the microcanonical ensemble at an average
735 temperature of 300K. Each of the different electrostatic methods
736 (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
737 damping values. The molecular system was started with same initial
738 positions and velocities for all cutoff methods. The energy drift
739 ($\delta E_1$) and standard deviation of the energy about the slope
740 ($\delta E_0$) were evaluated from the total energy of the system as a
741 function of time. Although both measures are valuable at
742 investigating new methods for molecular dynamics, a useful interaction
743 model must allow for long simulation times with minimal energy drift.
744
745 \section{\label{sec:result}RESULTS}
746 \subsection{Configurational energy differences}
747 %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.
748 %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.
749 %\begin{figure}[h!]
750 % \centering
751 % \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
752 % \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. }
753 % \label{fig:rcutConvergence_dipolarLiquid}
754 % \end{figure}
755 %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.
756 %\begin{figure}
757 % \centering
758 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
759 % \label{fig:barGraph1}
760 % \end{figure}
761 % \begin{figure}
762 % \centering
763 % \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
764 % \caption{}
765
766 % \label{fig:barGraph2}
767 % \end{figure}
768 %The correlation coefficient ($R^2$) and slope of the linear
769 %regression plots for the energy differences for all six different
770 %molecular systems is shown in figure 4a and 4b.The plot shows that
771 %the correlation coefficient improves for the SP cutoff method as
772 %compared to the undamped hard cutoff method in the case of SSDQC,
773 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
774 %crystal and liquid, the correlation coefficient is almost unchanged
775 %and close to 1. The correlation coefficient is smallest (0.696276
776 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
777 %charge-charge and charge-multipole interactions. Since the
778 %charge-charge and charge-multipole interaction is long ranged, there
779 %is huge deviation of correlation coefficient from 1. Similarly, the
780 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
781 %compared to interactions in the other multipolar systems, thus the
782 %correlation coefficient very close to 1 even for hard cutoff
783 %method. The idea of placing image multipole on the surface of the
784 %cutoff sphere improves the correlation coefficient and makes it close
785 %to 1 for all types of multipolar systems. Similarly the slope is
786 %hugely deviated from the correct value for the lower order
787 %multipole-multipole interaction and slightly deviated for higher
788 %order multipole – multipole interaction. The SP method improves both
789 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
790 %dipolar systems. The Slope is found to be deviated more in dipolar
791 %crystal as compared to liquid which is associated with the large
792 %fluctuation in the electrostatic energy in crystal. The GSF also
793 %produced better values of correlation coefficient and slope with the
794 %proper selection of the damping alpha (Interested reader can consult
795 %accompanying supporting material). The TSF method gives good value of
796 %correlation coefficient for the dipolar crystal, dipolar liquid,
797 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
798 %regression slopes are significantly deviated.
799
800 \begin{figure}
801 \centering
802 \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
803 \caption{Statistical analysis of the quality of configurational
804 energy differences for the real-space electrostatic methods
805 compared with the reference Ewald sum. Results with a value equal
806 to 1 (dashed line) indicate $\Delta E$ values indistinguishable
807 from those obtained using the multipolar Ewald sum. Different
808 values of the cutoff radius are indicated with different symbols
809 (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
810 triangles).}
811 \label{fig:slopeCorr_energy}
812 \end{figure}
813
814 The combined correlation coefficient and slope for all six systems is
815 shown in Figure ~\ref{fig:slopeCorr_energy}. Most of the methods
816 reproduce the Ewald configurational energy differences with remarkable
817 fidelity. Undamped hard cutoffs introduce a significant amount of
818 random scatter in the energy differences which is apparent in the
819 reduced value of the correlation coefficient for this method. This
820 can be easily understood as configurations which exhibit small
821 traversals of a few dipoles or quadrupoles out of the cutoff sphere
822 will see large energy jumps when hard cutoffs are used. The
823 orientations of the multipoles (particularly in the ordered crystals)
824 mean that these energy jumps can go in either direction, producing a
825 significant amount of random scatter, but no systematic error.
826
827 The TSF method produces energy differences that are highly correlated
828 with the Ewald results, but it also introduces a significant
829 systematic bias in the values of the energies, particularly for
830 smaller cutoff values. The TSF method alters the distance dependence
831 of different orientational contributions to the energy in a
832 non-uniform way, so the size of the cutoff sphere can have a large
833 effect, particularly for the crystalline systems.
834
835 Both the SP and GSF methods appear to reproduce the Ewald results with
836 excellent fidelity, particularly for moderate damping ($\alpha =
837 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
838 12$\AA). With the exception of the undamped hard cutoff, and the TSF
839 method with short cutoffs, all of the methods would be appropriate for
840 use in Monte Carlo simulations.
841
842 \subsection{Magnitude of the force and torque vectors}
843
844 The comparisons of the magnitudes of the forces and torques for the
845 data accumulated from all six systems are shown in Figures
846 ~\ref{fig:slopeCorr_force} and \ref{fig:slopeCorr_torque}. The
847 correlation and slope for the forces agree well with the Ewald sum
848 even for the hard cutoffs.
849
850 For systems of molecules with only multipolar interactions, the pair
851 energy contributions are quite short ranged. Moreover, the force
852 decays more rapidly than the electrostatic energy, hence the hard
853 cutoff method can also produce reasonable agreement for this quantity.
854 Although the pure cutoff gives reasonably good electrostatic forces
855 for pairs of molecules included within each other's cutoff spheres,
856 the discontinuity in the force at the cutoff radius can potentially
857 cause energy conservation problems as molecules enter and leave the
858 cutoff spheres. This is discussed in detail in section
859 \ref{sec:conservation}.
860
861 The two shifted-force methods (GSF and TSF) exhibit a small amount of
862 systematic variation and scatter compared with the Ewald forces. The
863 shifted-force models intentionally perturb the forces between pairs of
864 molecules inside each other's cutoff spheres in order to correct the
865 energy conservation issues, and this perturbation is evident in the
866 statistics accumulated for the molecular forces. The GSF
867 perturbations are minimal, particularly for moderate damping and
868 commonly-used cutoff values ($r_c = 12$\AA). The TSF method shows
869 reasonable agreement in the correlation coefficient but again the
870 systematic error in the forces is concerning if replication of Ewald
871 forces is desired.
872
873 \begin{figure}
874 \centering
875 \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
876 \caption{Statistical analysis of the quality of the force vector
877 magnitudes for the real-space electrostatic methods compared with
878 the reference Ewald sum. Results with a value equal to 1 (dashed
879 line) indicate force magnitude values indistinguishable from those
880 obtained using the multipolar Ewald sum. Different values of the
881 cutoff radius are indicated with different symbols (9\AA\ =
882 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
883 \label{fig:slopeCorr_force}
884 \end{figure}
885
886
887 \begin{figure}
888 \centering
889 \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
890 \caption{Statistical analysis of the quality of the torque vector
891 magnitudes for the real-space electrostatic methods compared with
892 the reference Ewald sum. Results with a value equal to 1 (dashed
893 line) indicate force magnitude values indistinguishable from those
894 obtained using the multipolar Ewald sum. Different values of the
895 cutoff radius are indicated with different symbols (9\AA\ =
896 circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
897 \label{fig:slopeCorr_torque}
898 \end{figure}
899
900 The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
901 significantly influenced by the choice of real-space method. The
902 torque expressions have the same distance dependence as the energies,
903 which are naturally longer-ranged expressions than the inter-site
904 forces. Torques are also quite sensitive to orientations of
905 neighboring molecules, even those that are near the cutoff distance.
906
907 The results shows that the torque from the hard cutoff method
908 reproduces the torques in quite good agreement with the Ewald sum.
909 The other real-space methods can cause some deviations, but excellent
910 agreement with the Ewald sum torques is recovered at moderate values
911 of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
912 radius ($r_c \ge 12$\AA). The TSF method exhibits only fair agreement
913 in the slope when compared with the Ewald torques even for larger
914 cutoff radii. It appears that the severity of the perturbations in
915 the TSF method are most in evidence for the torques.
916
917 \subsection{Directionality of the force and torque vectors}
918
919 The accurate evaluation of force and torque directions is just as
920 important for molecular dynamics simulations as the magnitudes of
921 these quantities. Force and torque vectors for all six systems were
922 analyzed using Fisher statistics, and the quality of the vector
923 directionality is shown in terms of circular variance
924 ($\mathrm{Var}(\theta)$) in figure
925 \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
926 from the new real-space methods exhibit nearly-ideal Fisher probability
927 distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
928 exhibit the best vectorial agreement with the Ewald sum. The force and
929 torque vectors from GSF method also show good agreement with the Ewald
930 method, which can also be systematically improved by using moderate
931 damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
932 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
933 to a distribution with 95\% of force vectors within $6.37^\circ$ of
934 the corresponding Ewald forces. The TSF method produces the poorest
935 agreement with the Ewald force directions.
936
937 Torques are again more perturbed than the forces by the new real-space
938 methods, but even here the variance is reasonably small. For the same
939 method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
940 the circular variance was 0.01415, corresponds to a distribution which
941 has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
942 results. Again, the direction of the force and torque vectors can be
943 systematically improved by varying $\alpha$ and $r_c$.
944
945 \begin{figure}
946 \centering
947 \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
948 \caption{The circular variance of the direction of the force and
949 torque vectors obtained from the real-space methods around the
950 reference Ewald vectors. A variance equal to 0 (dashed line)
951 indicates direction of the force or torque vectors are
952 indistinguishable from those obtained from the Ewald sum. Here
953 different symbols represent different values of the cutoff radius
954 (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
955 \label{fig:slopeCorr_circularVariance}
956 \end{figure}
957
958 \subsection{Energy conservation\label{sec:conservation}}
959
960 We have tested the conservation of energy one can expect to see with
961 the new real-space methods using the SSDQ water model with a small
962 fraction of solvated ions. This is a test system which exercises all
963 orders of multipole-multipole interactions derived in the first paper
964 in this series and provides the most comprehensive test of the new
965 methods. A liquid-phase system was created with 2000 water molecules
966 and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
967 temperature of 300K. After equilibration, this liquid-phase system
968 was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
969 a cutoff radius of 12\AA. The value of the damping coefficient was
970 also varied from the undamped case ($\alpha = 0$) to a heavily damped
971 case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods. A
972 sample was also run using the multipolar Ewald sum with the same
973 real-space cutoff.
974
975 In figure~\ref{fig:energyDrift} we show the both the linear drift in
976 energy over time, $\delta E_1$, and the standard deviation of energy
977 fluctuations around this drift $\delta E_0$. Both of the
978 shifted-force methods (GSF and TSF) provide excellent energy
979 conservation (drift less than $10^{-5}$ kcal / mol / ns / particle),
980 while the hard cutoff is essentially unusable for molecular dynamics.
981 SP provides some benefit over the hard cutoff because the energetic
982 jumps that happen as particles leave and enter the cutoff sphere are
983 somewhat reduced, but like the Wolf method for charges, the SP method
984 would not be as useful for molecular dynamics as either of the
985 shifted-force methods.
986
987 We note that for all tested values of the cutoff radius, the new
988 real-space methods can provide better energy conservation behavior
989 than the multipolar Ewald sum, even when utilizing a relatively large
990 $k$-space cutoff values.
991
992 \begin{figure}
993 \centering
994 \includegraphics[width=\textwidth]{newDrift_12.pdf}
995 \label{fig:energyDrift}
996 \caption{Analysis of the energy conservation of the real-space
997 electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
998 energy over time (in kcal / mol / particle / ns) and $\delta
999 \mathrm{E}_0$ is the standard deviation of energy fluctuations
1000 around this drift (in kcal / mol / particle). All simulations were
1001 of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
1002 300 K starting from the same initial configuration. All runs
1003 utilized the same real-space cutoff, $r_c = 12$\AA.}
1004 \end{figure}
1005
1006
1007 \section{CONCLUSION}
1008 In the first paper in this series, we generalized the
1009 charge-neutralized electrostatic energy originally developed by Wolf
1010 \textit{et al.}\cite{Wolf:1999dn} to multipole-multipole interactions
1011 up to quadrupolar order. The SP method is essentially a
1012 multipole-capable version of the Wolf model. The SP method for
1013 multipoles provides excellent agreement with Ewald-derived energies,
1014 forces and torques, and is suitable for Monte Carlo simulations,
1015 although the forces and torques retain discontinuities at the cutoff
1016 distance that prevents its use in molecular dynamics.
1017
1018 We also developed two natural extensions of the damped shifted-force
1019 (DSF) model originally proposed by Fennel and
1020 Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1021 smooth truncation of energies, forces, and torques at the real-space
1022 cutoff, and both converge to DSF electrostatics for point-charge
1023 interactions. The TSF model is based on a high-order truncated Taylor
1024 expansion which can be relatively perturbative inside the cutoff
1025 sphere. The GSF model takes the gradient from an images of the
1026 interacting multipole that has been projected onto the cutoff sphere
1027 to derive shifted force and torque expressions, and is a significantly
1028 more gentle approach.
1029
1030 Of the two newly-developed shifted force models, the GSF method
1031 produced quantitative agreement with Ewald energy, force, and torques.
1032 It also performs well in conserving energy in MD simulations. The
1033 Taylor-shifted (TSF) model provides smooth dynamics, but these take
1034 place on a potential energy surface that is significantly perturbed
1035 from Ewald-based electrostatics.
1036
1037 % The direct truncation of any electrostatic potential energy without
1038 % multipole neutralization creates large fluctuations in molecular
1039 % simulations. This fluctuation in the energy is very large for the case
1040 % of crystal because of long range of multipole ordering (Refer paper
1041 % I).\cite{PaperI} This is also significant in the case of the liquid
1042 % because of the local multipole ordering in the molecules. If the net
1043 % multipole within cutoff radius neutralized within cutoff sphere by
1044 % placing image multiples on the surface of the sphere, this fluctuation
1045 % in the energy reduced significantly. Also, the multipole
1046 % neutralization in the generalized SP method showed very good agreement
1047 % with the Ewald as compared to direct truncation for the evaluation of
1048 % the $\triangle E$ between the configurations. In MD simulations, the
1049 % energy conservation is very important. The conservation of the total
1050 % energy can be ensured by i) enforcing the smooth truncation of the
1051 % energy, force and torque in the cutoff radius and ii) making the
1052 % energy, force and torque consistent with each other. The GSF and TSF
1053 % methods ensure the consistency and smooth truncation of the energy,
1054 % force and torque at the cutoff radius, as a result show very good
1055 % total energy conservation. But the TSF method does not show good
1056 % agreement in the absolute value of the electrostatic energy, force and
1057 % torque with the Ewald. The GSF method has mimicked Ewald’s force,
1058 % energy and torque accurately and also conserved energy.
1059
1060 The only cases we have found where the new GSF and SP real-space
1061 methods can be problematic are those which retain a bulk dipole moment
1062 at large distances (e.g. the $Z_1$ dipolar lattice). In ferroelectric
1063 materials, uniform weighting of the orientational contributions can be
1064 important for converging the total energy. In these cases, the
1065 damping function which causes the non-uniform weighting can be
1066 replaced by the bare electrostatic kernel, and the energies return to
1067 the expected converged values.
1068
1069 Based on the results of this work, the GSF method is a suitable and
1070 efficient replacement for the Ewald sum for evaluating electrostatic
1071 interactions in MD simulations. Both methods retain excellent
1072 fidelity to the Ewald energies, forces and torques. Additionally, the
1073 energy drift and fluctuations from the GSF electrostatics are better
1074 than a multipolar Ewald sum for finite-sized reciprocal spaces.
1075 Because they use real-space cutoffs with moderate cutoff radii, the
1076 GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1077 increases. Additionally, they can be made extremely efficient using
1078 spline interpolations of the radial functions. They require no
1079 Fourier transforms or $k$-space sums, and guarantee the smooth
1080 handling of energies, forces, and torques as multipoles cross the
1081 real-space cutoff boundary.
1082
1083 \begin{acknowledgments}
1084 JDG acknowledges helpful discussions with Christopher
1085 Fennell. Support for this project was provided by the National
1086 Science Foundation under grant CHE-1362211. Computational time was
1087 provided by the Center for Research Computing (CRC) at the
1088 University of Notre Dame.
1089 \end{acknowledgments}
1090
1091 %\bibliographystyle{aip}
1092 \newpage
1093 \bibliography{references}
1094 \end{document}
1095
1096 %
1097 % ****** End of file aipsamp.tex ******