ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/multipole/multipole_2/multipole2.tex
(Generate patch)

Comparing trunk/multipole/multipole_2/multipole2.tex (file contents):
Revision 4185 by gezelter, Sun Jun 15 00:18:18 2014 UTC vs.
Revision 4208 by gezelter, Fri Aug 8 15:11:49 2014 UTC

# Line 47 | Line 47 | preprint,
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 \\
50 > \title{Real space electrostatics for multipoles. II. Comparisons with
51 >  the Ewald Sum}
52  
53   \author{Madan Lamichhane}
54 < \affiliation{Department of Physics, University
55 < of Notre Dame, Notre Dame, IN 46556}%Lines break automatically or can be forced with \\
54 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
55  
56   \author{Kathie E. Newman}
57 < \affiliation{Department of Physics, University
59 < of Notre Dame, Notre Dame, IN 46556}
57 > \affiliation{Department of Physics, University of Notre Dame, Notre Dame, IN 46556}
58  
59   \author{J. Daniel Gezelter}%
60   \email{gezelter@nd.edu.}
61 < \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556%\\This line break forced with \textbackslash\textbackslash
62 < }%
61 > \affiliation{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556
62 > }
63  
64 < \date{\today}% It is always \today, today,
67 <             %  but any date may be explicitly specified
64 > \date{\today}
65  
66   \begin{abstract}
67 <  We have tested the real-space shifted potential (SP),
68 <  gradient-shifted force (GSF), and Taylor-shifted force (TSF) methods
69 <  for multipole interactions that were developed in the first paper in
70 <  this series, using the multipolar Ewald sum as a reference
71 <  method. The tests were carried out in a variety of condensed-phase
72 <  environments which were designed to test all levels of the
73 <  multipole-multipole interactions.  Comparisons of the energy
74 <  differences between configurations, molecular forces, and torques
75 <  were used to analyze how well the real-space models perform relative
76 <  to the more computationally expensive Ewald treatment.  We have also
77 <  investigated the energy conservation properties of the new methods
78 <  in molecular dynamics simulations. The SP method shows excellent
79 <  agreement with configurational energy differences, forces, and
80 <  torques, and would be suitable for use in Monte Carlo calculations.
81 <  Of the two new shifted-force methods, the GSF approach shows the
82 <  best agreement with Ewald-derived energies, forces, and torques and
83 <  exhibits energy conservation properties that make it an excellent
84 <  choice for efficient computation of electrostatic interactions in
85 <  molecular dynamics simulations.
67 >  We report on tests of the shifted potential (SP), gradient shifted
68 >  force (GSF), and Taylor shifted force (TSF) real-space methods for
69 >  multipole interactions developed in the first paper in this series,
70 >  using the multipolar Ewald sum as a reference method. The tests were
71 >  carried out in a variety of condensed-phase environments designed to
72 >  test up to quadrupole-quadrupole interactions.  Comparisons of the
73 >  energy differences between configurations, molecular forces, and
74 >  torques were used to analyze how well the real-space models perform
75 >  relative to the more computationally expensive Ewald treatment.  We
76 >  have also investigated the energy conservation, structural, and
77 >  dynamical properties of the new methods in molecular dynamics
78 >  simulations. The SP method shows excellent agreement with
79 >  configurational energy differences, forces, and torques, and would
80 >  be suitable for use in Monte Carlo calculations.  Of the two new
81 >  shifted-force methods, the GSF approach shows the best agreement
82 >  with Ewald-derived energies, forces, and torques and also exhibits
83 >  energy conservation properties that make it an excellent choice for
84 >  efficient computation of electrostatic interactions in molecular
85 >  dynamics simulations.  Both SP and GSF are able to reproduce
86 >  structural and dyanamical properties in the liquid models with
87 >  excellent fidelity.
88   \end{abstract}
89  
90   %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
# Line 94 | Line 93 | of Notre Dame, Notre Dame, IN 46556}
93  
94   \maketitle
95  
97
96   \section{\label{sec:intro}Introduction}
97   Computing the interactions between electrostatic sites is one of the
98   most expensive aspects of molecular simulations. There have been
# Line 105 | Line 103 | space. BETTER CITATIONS\cite{Clarke:1986eu,Woodcock75}
103   the conditionally convergent electrostatic energy is converted into
104   two absolutely convergent contributions, one which is carried out in
105   real space with a cutoff radius, and one in reciprocal
106 < space. BETTER CITATIONS\cite{Clarke:1986eu,Woodcock75}
106 > space.\cite{Ewald21,deLeeuw80,Smith81,Allen87}
107  
108   When carried out as originally formulated, the reciprocal-space
109   portion of the Ewald sum exhibits relatively poor computational
110 < scaling, making it prohibitive for large systems. By utilizing
111 < particle meshes and three dimensional fast Fourier transforms (FFT),
112 < the particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
113 < (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME) methods can decrease
114 < the computational cost from $O(N^2)$ down to $O(N \log
115 < N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}.
110 > scaling, making it prohibitive for large systems. By utilizing a
111 > particle mesh and three dimensional fast Fourier transforms (FFT), the
112 > particle-mesh Ewald (PME), particle-particle particle-mesh Ewald
113 > (P\textsuperscript{3}ME), and smooth particle mesh Ewald (SPME)
114 > methods can decrease the computational cost from $O(N^2)$ down to $O(N
115 > \log
116 > N)$.\cite{Takada93,Gunsteren94,Gunsteren95,Darden:1993pd,Essmann:1995pb}
117  
118   Because of the artificial periodicity required for the Ewald sum,
119   interfacial molecular systems such as membranes and liquid-vapor
120 < interfaces require modifications to the
121 < method.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
122 < Parry's extension of the three dimensional Ewald sum is appropriate
123 < for slab geometries.\cite{Parry:1975if} Modified Ewald methods that
124 < were developed to handle two-dimensional (2D) electrostatic
125 < interactions in interfacial systems have not seen similar
126 < particle-mesh treatments,\cite{Parry:1975if, Parry:1976fq, Clarke77,
127 <  Perram79,Rhee:1989kl,Spohr:1997sf,Yeh:1999oq} and still scale poorly
128 < with system size. The inherent periodicity in the Ewald’s method can
129 < also be problematic for interfacial molecular
130 < systems.\cite{Fennell:2006lq}
120 > interfaces require modifications to the method.  Parry's extension of
121 > the three dimensional Ewald sum is appropriate for slab
122 > geometries.\cite{Parry:1975if} Modified Ewald methods that were
123 > developed to handle two-dimensional (2-D) electrostatic
124 > interactions.\cite{Parry:1975if,Parry:1976fq,Clarke77,Perram79,Rhee:1989kl}
125 > These methods were originally quite computationally
126 > expensive.\cite{Spohr97,Yeh99} There have been several successful
127 > efforts that reduced the computational cost of 2-D lattice summations,
128 > bringing them more in line with the scaling for the full 3-D
129 > treatments.\cite{Yeh99,Kawata01,Arnold02,deJoannis02,Brodka04} The
130 > inherent periodicity required by the Ewald method can also be
131 > problematic in a number of protein/solvent and ionic solution
132 > environments.\cite{Roberts94,Roberts95,Luty96,Hunenberger99a,Hunenberger99b,Weber00,Fennell:2006lq}
133  
134   \subsection{Real-space methods}
135   Wolf \textit{et al.}\cite{Wolf:1999dn} proposed a real space $O(N)$
# Line 150 | Line 151 | application of Wolf's method are able to obtain accura
151   what is effectively a set of octupoles at large distances. These facts
152   suggest that the Madelung constants are relatively short ranged for
153   perfect ionic crystals.\cite{Wolf:1999dn} For this reason, careful
154 < application of Wolf's method are able to obtain accurate estimates of
154 > application of Wolf's method can provide accurate estimates of
155   Madelung constants using relatively short cutoff radii.
156  
157   Direct truncation of interactions at a cutoff radius creates numerical
158 < errors.  Wolf \textit{et al.}  argued that truncation errors are due
158 > errors.  Wolf \textit{et al.} suggest that truncation errors are due
159   to net charge remaining inside the cutoff sphere.\cite{Wolf:1999dn} To
160   neutralize this charge they proposed placing an image charge on the
161   surface of the cutoff sphere for every real charge inside the cutoff.
162   These charges are present for the evaluation of both the pair
163   interaction energy and the force, although the force expression
164 < maintained a discontinuity at the cutoff sphere.  In the original Wolf
164 > maintains a discontinuity at the cutoff sphere.  In the original Wolf
165   formulation, the total energy for the charge and image were not equal
166 < to the integral of their force expression, and as a result, the total
166 > to the integral of the force expression, and as a result, the total
167   energy would not be conserved in molecular dynamics (MD)
168   simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennel and
169   Gezelter later proposed shifted force variants of the Wolf method with
170   commensurate force and energy expressions that do not exhibit this
171 < problem.\cite{Fennell:2006lq} Related real-space methods were also
172 < proposed by Chen \textit{et
171 > problem.\cite{Zahn:2002hc,Fennell:2006lq} Related real-space methods
172 > were also proposed by Chen \textit{et
173    al.}\cite{Chen:2004du,Chen:2006ii,Denesyuk:2008ez,Rodgers:2006nw}
174 < and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has used
175 < neutralization of the higher order moments for the calculation of the
176 < electrostatic interaction of the point charge
176 < systems.\cite{Fukuda:2013sf}
174 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfuly
175 > used additional neutralization of higher order moments for systems of
176 > point charges.\cite{Fukuda:2013sf}
177  
178   \begin{figure}
179    \centering
180 <  \includegraphics[width=\linewidth]{schematic.pdf}
180 >  \includegraphics[width=\linewidth]{schematic.eps}
181    \caption{Top: Ionic systems exhibit local clustering of dissimilar
182      charges (in the smaller grey circle), so interactions are
183      effectively charge-multipole at longer distances.  With hard
# Line 188 | Line 188 | systems.\cite{Fukuda:2013sf}
188      orientational averaging helps to reduce the effective range of the
189      interactions in the fluid.  Placement of reversed image multipoles
190      on the surface of the cutoff sphere recovers the effective
191 <    higher-order multipole behavior.}
192 <  \label{fig:schematic}
191 >    higher-order multipole behavior. \label{fig:schematic}}
192   \end{figure}
193  
194   One can make a similar effective range argument for crystals of point
195   \textit{multipoles}. The Luttinger and Tisza treatment of energy
196   constants for dipolar lattices utilizes 24 basis vectors that contain
197 < dipoles at the eight corners of a unit cube.  Only three of these
198 < basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
197 > dipoles at the eight corners of a unit cube.\cite{LT} Only three of
198 > these basis vectors, $X_1, Y_1, \mathrm{~and~} Z_1,$ retain net dipole
199   moments, while the rest have zero net dipole and retain contributions
200 < only from higher order multipoles.  The lowest energy crystalline
200 > only from higher order multipoles.  The lowest-energy crystalline
201   structures are built out of basis vectors that have only residual
202   quadrupolar moments (e.g. the $Z_5$ array). In these low energy
203   structures, the effective interaction between a dipole at the center
# Line 221 | Line 220 | Even at elevated temperatures, there is, on average, l
220  
221   The shorter effective range of electrostatic interactions is not
222   limited to perfect crystals, but can also apply in disordered fluids.
223 < Even at elevated temperatures, there is, on average, local charge
224 < balance in an ionic liquid, where each positive ion has surroundings
225 < dominated by negaitve ions and vice versa.  The reversed-charge images
226 < on the cutoff sphere that are integral to the Wolf and DSF approaches
227 < retain the effective multipolar interactions as the charges traverse
228 < the cutoff boundary.
223 > Even at elevated temperatures, there is local charge balance in an
224 > ionic liquid, where each positive ion has surroundings dominated by
225 > negative ions and vice versa.  The reversed-charge images on the
226 > cutoff sphere that are integral to the Wolf and DSF approaches retain
227 > the effective multipolar interactions as the charges traverse the
228 > cutoff boundary.
229  
230   In multipolar fluids (see Fig. \ref{fig:schematic}) there is
231   significant orientational averaging that additionally reduces the
# Line 235 | Line 234 | interacting molecules traverse each other's cutoff bou
234   and reduce the effective range of the multipolar interactions as
235   interacting molecules traverse each other's cutoff boundaries.
236  
237 < % Because of this reason, although the nature of electrostatic
238 < % interaction short ranged, the hard cutoff sphere creates very large
239 < % fluctuation in the electrostatic energy for the perfect crystal. In
240 < % addition, the charge neutralized potential proposed by Wolf et
241 < % al. converged to correct Madelung constant but still holds oscillation
242 < % in the energy about correct Madelung energy.\cite{Wolf:1999dn}.  This
243 < % oscillation in the energy around its fully converged value can be due
244 < % to the non-neutralized value of the higher order moments within the
245 < % cutoff sphere.
237 > Forces and torques acting on atomic sites are fundamental in driving
238 > dynamics in molecular simulations, and the damped shifted force (DSF)
239 > energy kernel provides consistent energies and forces on charged atoms
240 > within the cutoff sphere. Both the energy and the force go smoothly to
241 > zero as an atom aproaches the cutoff radius. The comparisons of the
242 > accuracy these quantities between the DSF kernel and SPME was
243 > surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
244 > has seen increasing use in molecular systems with relatively uniform
245 > charge
246 > densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
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
248   \subsection{The damping function}
249   The damping function has been discussed in detail in the first paper
250   of this series.\cite{PaperI} The $1/r$ Coulombic kernel for the
# Line 282 | Line 270 | Because electrons in a molecule are not localized at s
270   required to compute configurational
271   energies.\cite{Ren06,Essex10,Essex11}
272  
273 < Because electrons in a molecule are not localized at specific points,
274 < the assignment of partial charges to atomic centers is always an
275 < approximation.  Atomic sites can also be assigned point multipoles and
276 < polarizabilities to increase the accuracy of the molecular model.
277 < Recently, water has been modeled with point multipoles up to octupolar
278 < order using the soft sticky dipole-quadrupole-octupole (SSDQO)
273 > Additionally, because electrons in a molecule are not localized at
274 > specific points, the assignment of partial charges to atomic centers
275 > is always an approximation.  For increased accuracy, atomic sites can
276 > also be assigned point multipoles and polarizabilities.  Recently,
277 > water has been modeled with point multipoles up to octupolar order
278 > using the soft sticky dipole-quadrupole-octupole (SSDQO)
279   model.\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} Atom-centered point
280   multipoles up to quadrupolar order have also been coupled with point
281   polarizabilities in the high-quality AMOEBA and iAMOEBA water
282   models.\cite{Ren:2003uq,Ren:2004kx,Ponder:2010vl,Wang:2013fk} However,
283   truncating point multipoles without smoothing the forces and torques
284 < will create energy conservation issues in molecular dynamics simulations.
284 > can create energy conservation issues in molecular dynamics
285 > simulations.
286  
287   In this paper we test a set of real-space methods that were developed
288   for point multipolar interactions.  These methods extend the damped
289   shifted force (DSF) and Wolf methods originally developed for
290   charge-charge interactions and generalize them for higher order
291 < multipoles. The detailed mathematical development of these methods has
292 < been presented in the first paper in this series, while this work
293 < covers the testing the energies, forces, torques, and energy
291 > multipoles.  The detailed mathematical development of these methods
292 > has been presented in the first paper in this series, while this work
293 > covers the testing of energies, forces, torques, and energy
294   conservation properties of the methods in realistic simulation
295   environments.  In all cases, the methods are compared with the
296 < reference method, a full multipolar Ewald treatment.
296 > reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
297  
298  
310 %\subsection{Conservation of total energy }
311 %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.
312
299   \section{\label{sec:method}Review of Methods}
300   Any real-space electrostatic method that is suitable for MD
301   simulations should have the electrostatic energy, forces and torques
302   between two sites go smoothly to zero as the distance between the
303 < sites, $r_{\bf ab}$ approaches the cutoff radius, $r_c$.  Requiring
303 > sites, $r_{ab}$ approaches the cutoff radius, $r_c$.  Requiring
304   this continuity at the cutoff is essential for energy conservation in
305   MD simulations.  The mathematical details of the shifted potential
306   (SP), gradient-shifted-force (GSF) and Taylor shifted-force (TSF)
# Line 328 | Line 314 | U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1
314   expressed as the product of two multipole operators and a Coulombic
315   kernel,
316   \begin{equation}
317 < U_{\bf{ab}}(r)=\hat{M}_{\bf a} \hat{M}_{\bf b} \frac{1}{r}  \label{kernel}.
317 > U_{ab}(r)= M_{a} M_{b} \frac{1}{r}  \label{kernel}.
318   \end{equation}
319 < where the multipole operator for site $\bf a$, $\hat{M}_{\bf a}$, is
320 < expressed in terms of the point charge, $C_{\bf a}$, dipole, ${\bf D}_{\bf
321 <    a}$, and quadrupole, $\mathbf{Q}_{\bf a}$, for object
336 < $\bf a$, etc.
319 > where the multipole operator for site $a$, $M_{a}$, is
320 > expressed in terms of the point charge, $C_{a}$, dipole, ${\bf D}_{a}$, and quadrupole, $\mathsf{Q}_{a}$, for object
321 > $a$, etc.
322  
338 % Interactions between multipoles can be expressed as higher derivatives
339 % of the bare Coulomb potential, so one way of ensuring that the forces
340 % and torques vanish at the cutoff distance is to include a larger
341 % number of terms in the truncated Taylor expansion, e.g.,
342 % %
343 % \begin{equation}
344 % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
345 % \end{equation}
346 % %
347 % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
348 % Thus, for $f(r)=1/r$, we find
349 % %
350 % \begin{equation}
351 % 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} .
352 % \end{equation}
353 % This function is an approximate electrostatic potential that has
354 % vanishing second derivatives at the cutoff radius, making it suitable
355 % for shifting the forces and torques of charge-dipole interactions.
356
323   The TSF potential for any multipole-multipole interaction can be
324   written
325   \begin{equation}
# Line 368 | Line 334 | performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
334   force, and torques, a Taylor expansion with $n$ terms must be
335   performed at cutoff radius ($r_c$) to obtain $f_n(r)$.
336  
371 % To carry out the same procedure for a damped electrostatic kernel, we
372 % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
373 % Many of the derivatives of the damped kernel are well known from
374 % Smith's early work on multipoles for the Ewald
375 % summation.\cite{Smith82,Smith98}
376
377 % Note that increasing the value of $n$ will add additional terms to the
378 % electrostatic potential, e.g., $f_2(r)$ includes orders up to
379 % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
380 % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
381 % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
382 % for computing multipole energies, forces, and torques, and smooth
383 % cutoffs of these quantities can be guaranteed as long as the number of
384 % terms in the Taylor series exceeds the derivative order required.
385
337   For multipole-multipole interactions, following this procedure results
338   in separate radial functions for each of the distinct orientational
339   contributions to the potential, and ensures that the forces and
340   torques from each of these contributions will vanish at the cutoff
341   radius.  For example, the direct dipole dot product
342 < ($\mathbf{D}_{\bf a}
343 < \cdot \mathbf{D}_{\bf b}$) is treated differently than the dipole-distance
342 > ($\mathbf{D}_{a}
343 > \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
344   dot products:
345   \begin{equation}
346 < U_{D_{\bf a}D_{\bf b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 <  \mathbf{D}_{\bf a} \cdot
348 < \mathbf{D}_{\bf b} \right) v_{21}(r) +
349 < \left( \mathbf{D}_{\bf a} \cdot \hat{r} \right)
350 < \left( \mathbf{D}_{\bf b} \cdot \hat{r} \right) v_{22}(r) \right]
346 > U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
347 >  \mathbf{D}_{a} \cdot
348 > \mathbf{D}_{b} \right) v_{21}(r) +
349 > \left( \mathbf{D}_{a} \cdot \hat{\mathbf{r}} \right)
350 > \left( \mathbf{D}_{b} \cdot \hat{\mathbf{r}} \right) v_{22}(r) \right]
351   \end{equation}
352  
353   For the Taylor shifted (TSF) method with the undamped kernel,
# Line 421 | Line 372 | U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r)
372   which have been projected onto the surface of the cutoff sphere
373   without changing their relative orientation,
374   \begin{equation}
375 < U_{D_{\bf a}D_{\bf b}}(r)  = U_{D_{\bf a}D_{\bf b}}(r) -
376 < U_{D_{\bf a} D_{\bf b}}(r_c)
377 <   - (r_{ab}-r_c) ~~~\hat{r}_{ab} \cdot
378 <  \nabla U_{D_{\bf a}D_{\bf b}}(r_c).
375 > U_{D_{a}D_{b}}(r)  = U_{D_{a}D_{b}}(r) -
376 > U_{D_{a}D_{b}}(r_c)
377 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
378 >  \nabla U_{D_{a}D_{b}}(r_c).
379   \end{equation}
380 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
430 <  a}$ and $\mathbf{D}_{\bf b}$, are retained at the cutoff distance
380 > Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
381   (although the signs are reversed for the dipole that has been
382   projected onto the cutoff sphere).  In many ways, this simpler
383   approach is closer in spirit to the original shifted force method, in
# Line 448 | Line 398 | and any multipolar site inside the cutoff radius is gi
398   In general, the gradient shifted potential between a central multipole
399   and any multipolar site inside the cutoff radius is given by,
400   \begin{equation}
401 <  U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
402 <    U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) - (r-r_c) \hat{\mathbf{r}}
403 <    \cdot \nabla U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
401 > U^{\text{GSF}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
402 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) - (r-r_c)
403 > \hat{\mathbf{r}} \cdot \nabla U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
404   \label{generic2}
405   \end{equation}
406   where the sum describes a separate force-shifting that is applied to
407   each orientational contribution to the energy.  In this expression,
408   $\hat{\mathbf{r}}$ is the unit vector connecting the two multipoles
409 < ($a$ and $b$) in space, and $\hat{\mathbf{a}}$ and $\hat{\mathbf{b}}$
409 > ($a$ and $b$) in space, and $\mathsf{A}$ and $\mathsf{B}$
410   represent the orientations the multipoles.
411  
412   The third term converges more rapidly than the first two terms as a
# Line 483 | Line 433 | U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf
433   interactions with the central multipole and the image. This
434   effectively shifts the total potential to zero at the cutoff radius,
435   \begin{equation}
436 < U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \hat{\mathbf{a}}, \hat{\mathbf{b}}) -
437 < U(r_c \hat{\mathbf{r}},\hat{\mathbf{a}}, \hat{\mathbf{b}}) \right]
436 > U^{\text{SP}} = \sum \left[ U(\mathbf{r}, \mathsf{A}, \mathsf{B}) -
437 > U(r_c \hat{\mathbf{r}},\mathsf{A}, \mathsf{B}) \right]
438   \label{eq:SP}
439   \end{equation}          
440   where the sum describes separate potential shifting that is done for
# Line 529 | Line 479 | in the test cases are given in table~\ref{tab:pars}.
479   used the multipolar Ewald sum as a reference method for comparing
480   energies, forces, and torques for molecular models that mimic
481   disordered and ordered condensed-phase systems.  The parameters used
482 < in the test cases are given in table~\ref{tab:pars}.
482 > in the test cases are given in table~\ref{tab:pars}.
483  
484   \begin{table}
535 \label{tab:pars}
485   \caption{The parameters used in the systems used to evaluate the new
486    real-space methods.  The most comprehensive test was a liquid
487    composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
488    ions).  This test excercises all orders of the multipolar
489 <  interactions developed in the first paper.}
489 >  interactions developed in the first paper.\label{tab:pars}}
490   \begin{tabularx}{\textwidth}{r|cc|YYccc|Yccc} \hline
491               & \multicolumn{2}{c|}{LJ parameters} &
492               \multicolumn{5}{c|}{Electrostatic moments} & & & & \\
# Line 577 | Line 526 | the simulation proceeds. These differences are the mos
526   and have been compared with the values obtained from the multipolar
527   Ewald sum.  In Monte Carlo (MC) simulations, the energy differences
528   between two configurations is the primary quantity that governs how
529 < the simulation proceeds. These differences are the most imporant
529 > the simulation proceeds. These differences are the most important
530   indicators of the reliability of a method even if the absolute
531   energies are not exact.  For each of the multipolar systems listed
532   above, we have compared the change in electrostatic potential energy
# Line 589 | Line 538 | program, OpenMD,\cite{openmd} which was used for all c
538   \subsection{Implementation}
539   The real-space methods developed in the first paper in this series
540   have been implemented in our group's open source molecular simulation
541 < program, OpenMD,\cite{openmd} which was used for all calculations in
541 > program, OpenMD,\cite{Meineke05,openmd} which was used for all calculations in
542   this work.  The complementary error function can be a relatively slow
543   function on some processors, so all of the radial functions are
544   precomputed on a fine grid and are spline-interpolated to provide
# Line 599 | Line 548 | approximations.\cite{Smith82,Smith98} In all cases, th
548   with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
549   the Ewald sum is a re-implementation of the algorithm originally
550   proposed by Smith that does not use the particle mesh or smoothing
551 < approximations.\cite{Smith82,Smith98} In all cases, the quantities
552 < being compared are the electrostatic contributions to energies, force,
553 < and torques.  All other contributions to these quantities (i.e. from
554 < Lennard-Jones interactions) are removed prior to the comparisons.
551 > approximations.\cite{Smith82,Smith98} This implementation was tested
552 > extensively against the analytic energy constants for the multipolar
553 > lattices that are discussed in reference \onlinecite{PaperI}.  In all
554 > cases discussed below, the quantities being compared are the
555 > electrostatic contributions to energies, force, and torques.  All
556 > other contributions to these quantities (i.e. from Lennard-Jones
557 > interactions) are removed prior to the comparisons.
558  
559   The convergence parameter ($\alpha$) also plays a role in the balance
560   of the real-space and reciprocal-space portions of the Ewald
# Line 741 | Line 693 | model must allow for long simulation times with minima
693  
694   \section{\label{sec:result}RESULTS}
695   \subsection{Configurational energy differences}
744 %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.
745 %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.
746 %\begin{figure}[h!]
747 %        \centering
748 %        \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
749 %        \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.  }
750 %        \label{fig:rcutConvergence_dipolarLiquid}
751 %    \end{figure}
752 %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.
753 %\begin{figure}
754 %        \centering
755 %        \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
756 %        \label{fig:barGraph1}
757 %        \end{figure}
758 %        \begin{figure}
759 %        \centering
760 %       \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
761 %        \caption{}
762      
763 %        \label{fig:barGraph2}
764 %      \end{figure}
765 %The correlation coefficient ($R^2$) and slope of the linear
766 %regression plots for the energy differences for all six different
767 %molecular systems is shown in figure 4a and 4b.The plot shows that
768 %the correlation coefficient improves for the SP cutoff method as
769 %compared to the undamped hard cutoff method in the case of SSDQC,
770 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
771 %crystal and liquid, the correlation coefficient is almost unchanged
772 %and close to 1.  The correlation coefficient is smallest (0.696276
773 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
774 %charge-charge and charge-multipole interactions. Since the
775 %charge-charge and charge-multipole interaction is long ranged, there
776 %is huge deviation of correlation coefficient from 1. Similarly, the
777 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
778 %compared to interactions in the other multipolar systems, thus the
779 %correlation coefficient very close to 1 even for hard cutoff
780 %method. The idea of placing image multipole on the surface of the
781 %cutoff sphere improves the correlation coefficient and makes it close
782 %to 1 for all types of multipolar systems. Similarly the slope is
783 %hugely deviated from the correct value for the lower order
784 %multipole-multipole interaction and slightly deviated for higher
785 %order multipole – multipole interaction. The SP method improves both
786 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
787 %dipolar systems.  The Slope is found to be deviated more in dipolar
788 %crystal as compared to liquid which is associated with the large
789 %fluctuation in the electrostatic energy in crystal. The GSF also
790 %produced better values of correlation coefficient and slope with the
791 %proper selection of the damping alpha (Interested reader can consult
792 %accompanying supporting material). The TSF method gives good value of
793 %correlation coefficient for the dipolar crystal, dipolar liquid,
794 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
795 %regression slopes are significantly deviated.
696  
697   \begin{figure}
698    \centering
699 <  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined-crop.pdf}
699 >  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
700    \caption{Statistical analysis of the quality of configurational
701      energy differences for the real-space electrostatic methods
702      compared with the reference Ewald sum.  Results with a value equal
# Line 804 | Line 704 | model must allow for long simulation times with minima
704      from those obtained using the multipolar Ewald sum.  Different
705      values of the cutoff radius are indicated with different symbols
706      (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
707 <    triangles).}
808 <  \label{fig:slopeCorr_energy}
707 >    triangles).\label{fig:slopeCorr_energy}}
708   \end{figure}
709  
710   The combined correlation coefficient and slope for all six systems is
# Line 867 | Line 766 | forces is desired.
766   systematic error in the forces is concerning if replication of Ewald
767   forces is desired.
768  
769 + It is important to note that the forces and torques from the SP and
770 + the Hard cutoffs are not identical. The SP method shifts each
771 + orientational contribution separately (e.g. the dipole-dipole dot
772 + product is shifted by a different function than the dipole-distance
773 + products), while the hard cutoff contains no orientation-dependent
774 + shifting.  The forces and torques for these methods therefore diverge
775 + for multipoles even though the forces for point charges are identical.
776 +
777   \begin{figure}
778    \centering
779 <  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined-crop.pdf}
779 >  \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
780    \caption{Statistical analysis of the quality of the force vector
781      magnitudes for the real-space electrostatic methods compared with
782      the reference Ewald sum. Results with a value equal to 1 (dashed
783      line) indicate force magnitude values indistinguishable from those
784      obtained using the multipolar Ewald sum.  Different values of the
785      cutoff radius are indicated with different symbols (9\AA\ =
786 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
787 <  \label{fig:slopeCorr_force}
786 >    circles, 12\AA\ = squares, and 15\AA\ = inverted
787 >    triangles).\label{fig:slopeCorr_force}}
788   \end{figure}
789  
790  
791   \begin{figure}
792    \centering
793 <  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined-crop.pdf}
793 >  \includegraphics[width=0.6\linewidth]{torquePlot_slopeCorrelation_combined.eps}
794    \caption{Statistical analysis of the quality of the torque vector
795      magnitudes for the real-space electrostatic methods compared with
796      the reference Ewald sum. Results with a value equal to 1 (dashed
797      line) indicate force magnitude values indistinguishable from those
798      obtained using the multipolar Ewald sum.  Different values of the
799      cutoff radius are indicated with different symbols (9\AA\ =
800 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
801 <  \label{fig:slopeCorr_torque}
800 >    circles, 12\AA\ = squares, and 15\AA\ = inverted
801 >    triangles).\label{fig:slopeCorr_torque}}
802   \end{figure}
803  
804   The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
# Line 941 | Line 848 | systematically improved by varying $\alpha$ and $r_c$.
848  
849   \begin{figure}
850    \centering
851 <  \includegraphics[width=0.6 \linewidth]{Variance_forceNtorque_modified-crop.pdf}
851 >  \includegraphics[width=0.65\linewidth]{Variance_forceNtorque_modified.eps}
852    \caption{The circular variance of the direction of the force and
853      torque vectors obtained from the real-space methods around the
854      reference Ewald vectors. A variance equal to 0 (dashed line)
855      indicates direction of the force or torque vectors are
856      indistinguishable from those obtained from the Ewald sum. Here
857      different symbols represent different values of the cutoff radius
858 <    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
952 <  \label{fig:slopeCorr_circularVariance}
858 >    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)\label{fig:slopeCorr_circularVariance}}
859   \end{figure}
860  
861   \subsection{Energy conservation\label{sec:conservation}}
# Line 961 | Line 867 | temperature of 300K.  After equilibration, this liquid
867   in this series and provides the most comprehensive test of the new
868   methods.  A liquid-phase system was created with 2000 water molecules
869   and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
870 < temperature of 300K.  After equilibration, this liquid-phase system
871 < was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
872 < a cutoff radius of 12\AA.  The value of the damping coefficient was
873 < also varied from the undamped case ($\alpha = 0$) to a heavily damped
874 < case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
875 < sample was also run using the multipolar Ewald sum with the same
876 < real-space cutoff.
870 > temperature of 300K.  After equilibration in the canonical (NVT)
871 > ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
872 > was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
873 > Hard, SP, GSF, and TSF methods with a cutoff radius of 12\AA.  The
874 > value of the damping coefficient was also varied from the undamped
875 > case ($\alpha = 0$) to a heavily damped case ($\alpha = 0.3$
876 > \AA$^{-1}$) for all of the real space methods.  A sample was also run
877 > using the multipolar Ewald sum with the same real-space cutoff.
878  
879   In figure~\ref{fig:energyDrift} we show the both the linear drift in
880   energy over time, $\delta E_1$, and the standard deviation of energy
# Line 983 | Line 890 | than the multipolar Ewald sum, even when utilizing a r
890  
891   We note that for all tested values of the cutoff radius, the new
892   real-space methods can provide better energy conservation behavior
893 < than the multipolar Ewald sum, even when utilizing a relatively large
894 < $k$-space cutoff values.
893 > than the multipolar Ewald sum, even when relatively large $k$-space
894 > cutoff values are utilized.
895  
896   \begin{figure}
897    \centering
898 <  \includegraphics[width=\textwidth]{newDrift_12.pdf}
899 < \label{fig:energyDrift}        
900 < \caption{Analysis of the energy conservation of the real-space
901 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
995 <  energy over time (in kcal / mol / particle / ns) and $\delta
898 >  \includegraphics[width=\textwidth]{newDrift_12.eps}
899 > \caption{Analysis of the energy conservation of the real-space methods
900 >  for the SSDQ water/ion system. $\delta \mathrm{E}_1$ is the linear
901 >  drift in energy over time (in kcal/mol/particle/ns) and $\delta
902    \mathrm{E}_0$ is the standard deviation of energy fluctuations
903 <  around this drift (in kcal / mol / particle).  All simulations were
904 <  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
905 <  300 K starting from the same initial configuration. All runs
906 <  utilized the same real-space cutoff, $r_c = 12$\AA.}
903 >  around this drift (in kcal/mol/particle).  Points that appear in the
904 >  green region at the bottom exhibit better energy conservation than
905 >  would be obtained using common parameters for Ewald-based
906 >  electrostatics.\label{fig:energyDrift}}
907 > \end{figure}
908 >
909 > \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
910 > The most important test of the modified interaction potentials is the
911 > fidelity with which they can reproduce structural features and
912 > dynamical properties in a liquid.  One commonly-utilized measure of
913 > structural ordering is the pair distribution function, $g(r)$, which
914 > measures local density deviations in relation to the bulk density.  In
915 > the electrostatic approaches studied here, the short-range repulsion
916 > from the Lennard-Jones potential is identical for the various
917 > electrostatic methods, and since short range repulsion determines much
918 > of the local liquid ordering, one would not expect to see many
919 > differences in $g(r)$.  Indeed, the pair distributions are essentially
920 > identical for all of the electrostatic methods studied (for each of
921 > the different systems under investigation).  An example of this
922 > agreement for the SSDQ water/ion system is shown in
923 > Fig. \ref{fig:gofr}.
924 >
925 > \begin{figure}
926 >  \centering
927 >  \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
928 > \caption{The pair distribution functions, $g(r)$, for the SSDQ
929 >  water/ion system obtained using the different real-space methods are
930 >  essentially identical with the result from the Ewald
931 >  treatment.\label{fig:gofr}}
932   \end{figure}
933 +
934 + There is a very slight overstructuring of the first solvation shell
935 + when using when using TSF at lower values of the damping coefficient
936 + ($\alpha \le 0.1$) or when using undamped GSF.  With moderate damping,
937 + GSF and SP produce pair distributions that are identical (within
938 + numerical noise) to their Ewald counterparts.
939 +
940 + A structural property that is a more demanding test of modified
941 + electrostatics is the mean value of the electrostatic energy $\langle
942 + U_\mathrm{elect} \rangle / N$ which is obtained by sampling the
943 + liquid-state configurations experienced by a liquid evolving entirely
944 + under the influence of each of the methods.  In table \ref{tab:Props}
945 + we demonstrate how $\langle U_\mathrm{elect} \rangle / N$ varies with
946 + the damping parameter, $\alpha$, for each of the methods.
947 +
948 + As in the crystals studied in the first paper, damping is important
949 + for converging the mean electrostatic energy values, particularly for
950 + the two shifted force methods (GSF and TSF).  A value of $\alpha
951 + \approx 0.2$ \AA$^{-1}$ is sufficient to converge the SP and GSF
952 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
953 + dramatic damping ($\alpha \approx 0.3$ \AA$^{-1}$ for $r_c = 9$ \AA).
954 + Overdamping the real-space electrostatic methods occurs with $\alpha >
955 + 0.4$, causing the estimate of the energy to drop below the Ewald
956 + results.
957 +
958 + These ``optimal'' values of the damping coefficient are slightly
959 + larger than what were observed for DSF electrostatics for purely
960 + point-charge systems, although a value of $\alpha=0.18$ \AA$^{-1}$ for
961 + $r_c = 12$\AA\ appears to be an excellent compromise for mixed
962 + charge/multipolar systems.
963 +
964 + To test the fidelity of the electrostatic methods at reproducing
965 + dynamics in a multipolar liquid, it is also useful to look at
966 + transport properties, particularly the diffusion constant,
967 + \begin{equation}
968 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
969 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
970 + \label{eq:diff}
971 + \end{equation}
972 + which measures long-time behavior and is sensitive to the forces on
973 + the multipoles.  For the soft dipolar fluid and the SSDQ liquid
974 + systems, the self-diffusion constants (D) were calculated from linear
975 + fits to the long-time portion of the mean square displacement,
976 + $\langle r^{2}(t) \rangle$.\cite{Allen87}
977 +
978 + In addition to translational diffusion, orientational relaxation times
979 + were calculated for comparisons with the Ewald simulations and with
980 + experiments. These values were determined from the same 1~ns
981 + microcanonical trajectories used for translational diffusion by
982 + calculating the orientational time correlation function,
983 + \begin{equation}
984 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
985 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
986 + \label{eq:OrientCorr}
987 + \end{equation}
988 + where $P_l$ is the Legendre polynomial of order $l$ and
989 + $\hat{\mathbf{A}}_\gamma$ is the unit vector for body axis $\gamma$.
990 + The reference frame used for our sample dipolar systems has the
991 + $z$-axis running along the dipoles, and for the SSDQ water model, the
992 + $y$-axis connects the two implied hydrogen atom positions.  From the
993 + orientation autocorrelation functions, we can obtain time constants
994 + for rotational relaxation either by fitting an exponential function or
995 + by integrating the entire correlation function.  In a good water
996 + model, these decay times would be comparable to water orientational
997 + relaxation times from nuclear magnetic resonance (NMR). The relaxation
998 + constant obtained from $C_2^y(t)$ is normally of experimental interest
999 + because it describes the relaxation of the principle axis connecting
1000 + the hydrogen atoms. Thus, $C_2^y(t)$ can be compared to the
1001 + intermolecular portion of the dipole-dipole relaxation from a proton
1002 + NMR signal and should provide an estimate of the NMR relaxation time
1003 + constant.\cite{Impey82}
1004  
1005 + Results for the diffusion constants and orientational relaxation times
1006 + are shown in figure \ref{tab:Props}. From this data, it is apparent
1007 + that the values for both $D$ and $\tau_2$ using the Ewald sum are
1008 + reproduced with reasonable fidelity by the GSF method.
1009  
1010 + The $\tau_2$ results in \ref{tab:Props} show a much greater difference
1011 + between the real-space and the Ewald results.
1012 +
1013 + \begin{table}
1014 + \caption{Comparison of the structural and dynamic properties for the
1015 +  soft dipolar liquid test for all of the real-space methods.\label{tab:Props}}
1016 + \begin{tabular}{l|c|cccc|cccc|cccc}
1017 +         & Ewald & \multicolumn{4}{c|}{SP} & \multicolumn{4}{c|}{GSF} & \multicolumn{4}{c|}{TSF} \\
1018 + $\alpha$ (\AA$^{-1}$) & &      
1019 + 0.0 & 0.1 & 0.2 & 0.3 &
1020 + 0.0 & 0.1 & 0.2 & 0.3 &
1021 + 0.0 & 0.1 & 0.2 & 0.3 \\ \cline{2-6}\cline{6-10}\cline{10-14}
1022 +
1023 + $\langle U_\mathrm{elect} \rangle /N$ &&&&&&&&&&&&&\\
1024 + D ($10^{-4}~\mathrm{cm}^2/\mathrm{s}$)&
1025 + 470.2(6) &
1026 + 416.6(5) &
1027 + 379.6(5) &
1028 + 438.6(5) &
1029 + 476.0(6) &
1030 + 412.8(5) &
1031 + 421.1(5) &
1032 + 400.5(5) &
1033 + 437.5(6) &
1034 + 434.6(5) &
1035 + 411.4(5) &
1036 + 545.3(7) &
1037 + 459.6(6) \\
1038 + $\tau_2$ (fs) &
1039 + 1.136 &
1040 + 1.041 &
1041 + 1.064 &
1042 + 1.109 &
1043 + 1.211 &
1044 + 1.119 &
1045 + 1.039 &
1046 + 1.058 &
1047 + 1.21  &
1048 + 1.15  &
1049 + 1.172 &
1050 + 1.153 &
1051 + 1.125 \\
1052 + \end{tabular}
1053 + \end{table}
1054 +
1055 +
1056   \section{CONCLUSION}
1057   In the first paper in this series, we generalized the
1058   charge-neutralized electrostatic energy originally developed by Wolf
# Line 1013 | Line 1065 | We also developed two natural extensions of the damped
1065   distance that prevents its use in molecular dynamics.
1066  
1067   We also developed two natural extensions of the damped shifted-force
1068 < (DSF) model originally proposed by Fennel and
1069 < Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1070 < smooth truncation of energies, forces, and torques at the real-space
1071 < cutoff, and both converge to DSF electrostatics for point-charge
1072 < interactions.  The TSF model is based on a high-order truncated Taylor
1073 < expansion which can be relatively perturbative inside the cutoff
1074 < sphere.  The GSF model takes the gradient from an images of the
1075 < interacting multipole that has been projected onto the cutoff sphere
1076 < to derive shifted force and torque expressions, and is a significantly
1077 < more gentle approach.
1068 > (DSF) model originally proposed by Zahn {\it et al.} and extended by
1069 > Fennel and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
1070 > approaches provide smooth truncation of energies, forces, and torques
1071 > at the real-space cutoff, and both converge to DSF electrostatics for
1072 > point-charge interactions.  The TSF model is based on a high-order
1073 > truncated Taylor expansion which can be relatively perturbative inside
1074 > the cutoff sphere.  The GSF model takes the gradient from an images of
1075 > the interacting multipole that has been projected onto the cutoff
1076 > sphere to derive shifted force and torque expressions, and is a
1077 > significantly more gentle approach.
1078  
1079 < Of the two newly-developed shifted force models, the GSF method
1080 < produced quantitative agreement with Ewald energy, force, and torques.
1081 < It also performs well in conserving energy in MD simulations.  The
1082 < Taylor-shifted (TSF) model provides smooth dynamics, but these take
1083 < place on a potential energy surface that is significantly perturbed
1084 < from Ewald-based electrostatics.  
1079 > The GSF method produced quantitative agreement with Ewald energy,
1080 > force, and torques.  It also performs well in conserving energy in MD
1081 > simulations.  The Taylor-shifted (TSF) model provides smooth dynamics,
1082 > but these take place on a potential energy surface that is
1083 > significantly perturbed from Ewald-based electrostatics.  Because it
1084 > performs relatively poorly compared with GSF, it may seem odd that
1085 > that the TSF model was included in this work.  However, the functional
1086 > forms derived for the SP and GSF methods depend on the separation of
1087 > orientational contributions that were made visible by the Taylor
1088 > series of the electrostatic kernel at the cutoff radius. The TSF
1089 > method also has the unique property that a large number of derivatives
1090 > can be made to vanish at the cutoff radius.  This property has proven
1091 > useful in past treatments of the corrections to the fluctuation
1092 > formula for dielectric constants.\cite{Izvekov:2008wo}
1093  
1094 < % The direct truncation of any electrostatic potential energy without
1095 < % multipole neutralization creates large fluctuations in molecular
1096 < % simulations.  This fluctuation in the energy is very large for the case
1097 < % of crystal because of long range of multipole ordering (Refer paper
1098 < % I).\cite{PaperI} This is also significant in the case of the liquid
1099 < % because of the local multipole ordering in the molecules. If the net
1100 < % multipole within cutoff radius neutralized within cutoff sphere by
1101 < % placing image multiples on the surface of the sphere, this fluctuation
1102 < % in the energy reduced significantly. Also, the multipole
1103 < % neutralization in the generalized SP method showed very good agreement
1104 < % with the Ewald as compared to direct truncation for the evaluation of
1045 < % the $\triangle E$ between the configurations.  In MD simulations, the
1046 < % energy conservation is very important. The conservation of the total
1047 < % energy can be ensured by i) enforcing the smooth truncation of the
1048 < % energy, force and torque in the cutoff radius and ii) making the
1049 < % energy, force and torque consistent with each other. The GSF and TSF
1050 < % methods ensure the consistency and smooth truncation of the energy,
1051 < % force and torque at the cutoff radius, as a result show very good
1052 < % total energy conservation. But the TSF method does not show good
1053 < % agreement in the absolute value of the electrostatic energy, force and
1054 < % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1055 < % energy and torque accurately and also conserved energy.
1094 > Reproduction of both structural and dynamical features in the liquid
1095 > systems is remarkably good for both the SP and GSF models.  Pair
1096 > distribution functions are essentially equivalent to the same
1097 > functions produced using Ewald-based electrostatics, and with moderate
1098 > damping, a structural feature that directly probes the electrostatic
1099 > interaction (e.g. the mean electrostatic potential energy) can also be
1100 > made quantitative.  Dynamical features are sensitive probes of the
1101 > forces and torques produced by these methods, and even though the
1102 > smooth behavior of forces is produced by perturbing the overall
1103 > potential, the diffusion constants and orientational correlation times
1104 > are quite close to the Ewald-based results.
1105  
1106   The only cases we have found where the new GSF and SP real-space
1107   methods can be problematic are those which retain a bulk dipole moment
# Line 1063 | Line 1112 | Based on the results of this work, the GSF method is a
1112   replaced by the bare electrostatic kernel, and the energies return to
1113   the expected converged values.
1114  
1115 < Based on the results of this work, the GSF method is a suitable and
1116 < efficient replacement for the Ewald sum for evaluating electrostatic
1117 < interactions in MD simulations.  Both methods retain excellent
1118 < fidelity to the Ewald energies, forces and torques.  Additionally, the
1119 < energy drift and fluctuations from the GSF electrostatics are better
1120 < than a multipolar Ewald sum for finite-sized reciprocal spaces.
1121 < Because they use real-space cutoffs with moderate cutoff radii, the
1122 < GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1123 < increases.  Additionally, they can be made extremely efficient using
1075 < spline interpolations of the radial functions.  They require no
1076 < Fourier transforms or $k$-space sums, and guarantee the smooth
1077 < handling of energies, forces, and torques as multipoles cross the
1078 < real-space cutoff boundary.
1115 > Based on the results of this work, we can conclude that the GSF method
1116 > is a suitable and efficient replacement for the Ewald sum for
1117 > evaluating electrostatic interactions in modern MD simulations, and
1118 > the SP meethod would be an excellent choice for Monte Carlo
1119 > simulations where smooth forces and energy conservation are not
1120 > important.  Both the SP and GSF methods retain excellent fidelity to
1121 > the Ewald energies, forces and torques.  Additionally, the energy
1122 > drift and fluctuations from the GSF electrostatics are significantly
1123 > better than a multipolar Ewald sum for finite-sized reciprocal spaces.
1124  
1125 + As in all purely pairwise cutoff methods, the SP, GSF and TSF methods
1126 + are expected to scale approximately {\it linearly} with system size,
1127 + and are easily parallelizable.  This should result in substantial
1128 + reductions in the computational cost of performing large simulations.
1129 + With the proper use of pre-computation and spline interpolation of the
1130 + radial functions, the real-space methods are essentially the same cost
1131 + as a simple real-space cutoff.  They require no Fourier transforms or
1132 + $k$-space sums, and guarantee the smooth handling of energies, forces,
1133 + and torques as multipoles cross the real-space cutoff boundary.
1134 +
1135 + We are not suggesting that there is any flaw with the Ewald sum; in
1136 + fact, it is the standard by which the SP, GSF, and TSF methods have
1137 + been judged in this work.  However, these results provide evidence
1138 + that in the typical simulations performed today, the Ewald summation
1139 + may no longer be required to obtain the level of accuracy most
1140 + researchers have come to expect.
1141 +
1142   \begin{acknowledgments}
1143    JDG acknowledges helpful discussions with Christopher
1144    Fennell. Support for this project was provided by the National

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines