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 4191 by gezelter, Tue Jun 17 16:08:03 2014 UTC vs.
Revision 4214 by gezelter, Wed Sep 10 21:06:56 2014 UTC

# Line 47 | Line 47 | preprint,
47  
48   %\preprint{AIP/123-QED}
49  
50 < \title{Real space alternatives to the Ewald Sum. II. Comparison of Methods}
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 of Notre Dame, Notre Dame, IN 46556}
# Line 72 | Line 73 | preprint,
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 properties of the new
77 <  methods in molecular dynamics simulations. The SP method shows
78 <  excellent agreement with configurational energy differences, forces,
79 <  and torques, and would be suitable for use in Monte Carlo
80 <  calculations.  Of the two new shifted-force methods, the GSF
81 <  approach shows the best agreement with Ewald-derived energies,
82 <  forces, and torques and also exhibits energy conservation properties
83 <  that make it an excellent choice for efficient computation of
84 <  electrostatic interactions in molecular dynamics simulations.
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 dynamical 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 161 | Line 165 | simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, a
165   formulation, the total energy for the charge and image were not equal
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
168 > simulations.\cite{Zahn:2002hc} Zahn \textit{et al.}, and Fennell 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{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 successfuly
174 > and by Wu and Brooks.\cite{Wu:044107} Recently, Fukuda has successfully
175   used additional neutralization of higher order moments for systems of
176   point charges.\cite{Fukuda:2013sf}
177  
# Line 184 | Line 188 | point charges.\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.}
188 <  \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
# Line 219 | Line 222 | negaitve ions and vice versa.  The reversed-charge ima
222   limited to perfect crystals, but can also apply in disordered fluids.
223   Even at elevated temperatures, there is local charge balance in an
224   ionic liquid, where each positive ion has surroundings dominated by
225 < negaitve 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.
225 > negative ions and vice versa.  The reversed-charge images on the
226 > cutoff sphere that are integral to the Wolf and damped shifted force
227 > (DSF) approaches retain the effective multipolar interactions as the
228 > charges traverse the cutoff boundary.
229  
230   In multipolar fluids (see Fig. \ref{fig:schematic}) there is
231   significant orientational averaging that additionally reduces the
232   effect of long-range multipolar interactions.  The image multipoles
233 < that are introduced in the TSF, GSF, and SP methods mimic this effect
233 > that are introduced in the Taylor shifted force (TSF), gradient
234 > shifted force (GSF), and shifted potential (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.
233
234 % Because of this reason, although the nature of electrostatic
235 % interaction short ranged, the hard cutoff sphere creates very large
236 % fluctuation in the electrostatic energy for the perfect crystal. In
237 % addition, the charge neutralized potential proposed by Wolf et
238 % al. converged to correct Madelung constant but still holds oscillation
239 % in the energy about correct Madelung energy.\cite{Wolf:1999dn}.  This
240 % oscillation in the energy around its fully converged value can be due
241 % to the non-neutralized value of the higher order moments within the
242 % cutoff sphere.
237  
238   Forces and torques acting on atomic sites are fundamental in driving
239 < dynamics in molecular simulations, and the damped shifted force (DSF)
240 < energy kernel provides consistent energies and forces on charged atoms
241 < within the cutoff sphere. Both the energy and the force go smoothly to
242 < zero as an atom aproaches the cutoff radius. The comparisons of the
243 < accuracy these quantities between the DSF kernel and SPME was
244 < surprisingly good.\cite{Fennell:2006lq} As a result, the DSF method
245 < has seen increasing use in molecular systems with relatively uniform
252 < charge
239 > dynamics in molecular simulations, and the DSF energy kernel provides
240 > consistent energies and forces on charged atoms within the cutoff
241 > sphere. Both the energy and the force go smoothly to zero as an atom
242 > approaches the cutoff radius. The comparisons of the accuracy these
243 > quantities between the DSF kernel and SPME was surprisingly
244 > good.\cite{Fennell:2006lq} As a result, the DSF method has seen
245 > increasing use in molecular systems with relatively uniform charge
246   densities.\cite{English08,Kannam:2012rr,Space12,Lawrence13,Acevedo13,Shi:2013ij,Vergne13}
247  
248   \subsection{The damping function}
# Line 264 | Line 257 | With moderate damping coefficients, $\alpha \sim 0.2$,
257   produce complementary error functions when truncated at a finite
258   distance.
259  
260 < With moderate damping coefficients, $\alpha \sim 0.2$, the DSF method
260 > With moderate damping coefficients, $\alpha \sim 0.2$ \AA$^{-1}$, the DSF method
261   produced very good agreement with SPME for interaction energies,
262   forces and torques for charge-charge
263   interactions.\cite{Fennell:2006lq}
# Line 303 | Line 296 | reference method, a full multipolar Ewald treatment.\c
296   reference method, a full multipolar Ewald treatment.\cite{Smith82,Smith98}
297  
298  
306 %\subsection{Conservation of total energy }
307 %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.
308
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 324 | 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
332 < $\bf a$, etc.
333 <
334 < % Interactions between multipoles can be expressed as higher derivatives
335 < % of the bare Coulomb potential, so one way of ensuring that the forces
336 < % and torques vanish at the cutoff distance is to include a larger
337 < % number of terms in the truncated Taylor expansion, e.g.,
338 < % %
339 < % \begin{equation}
340 < % f_n^{\text{shift}}(r)=\sum_{m=0}^{n+1} \frac {(r-r_c)^m}{m!} f^{(m)} \Big \lvert  _{r_c}  .
341 < % \end{equation}
342 < % %
343 < % The combination of $f(r)$ with the shifted function is denoted $f_n(r)=f(r)-f_n^{\text{shift}}(r)$.
344 < % Thus, for $f(r)=1/r$, we find
345 < % %
346 < % \begin{equation}
347 < % 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} .
348 < % \end{equation}
349 < % This function is an approximate electrostatic potential that has
350 < % vanishing second derivatives at the cutoff radius, making it suitable
351 < % for shifting the forces and torques of charge-dipole interactions.
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  
323   The TSF potential for any multipole-multipole interaction can be
324   written
# Line 364 | 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  
367 % To carry out the same procedure for a damped electrostatic kernel, we
368 % replace $1/r$ in the Coulomb potential with $\text{erfc}(\alpha r)/r$.
369 % Many of the derivatives of the damped kernel are well known from
370 % Smith's early work on multipoles for the Ewald
371 % summation.\cite{Smith82,Smith98}
372
373 % Note that increasing the value of $n$ will add additional terms to the
374 % electrostatic potential, e.g., $f_2(r)$ includes orders up to
375 % $(r-r_c)^3/r_c^4$, and so on.  Successive derivatives of the $f_n(r)$
376 % functions are denoted $g_2(r) = f^\prime_2(r)$, $h_2(r) =
377 % f^{\prime\prime}_2(r)$, etc.  These higher derivatives are required
378 % for computing multipole energies, forces, and torques, and smooth
379 % cutoffs of these quantities can be guaranteed as long as the number of
380 % terms in the Taylor series exceeds the derivative order required.
381
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_{\mathbf{D}_{a}\mathbf{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 417 | 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_{\mathbf{D}_{a}\mathbf{D}_{b}}(r)  = U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r) -
376 > U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c)
377 >   - (r_{ab}-r_c) ~~~\hat{\mathbf{r}}_{ab} \cdot
378 >  \nabla U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c).
379   \end{equation}
380 < Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{\bf
426 <  a}$ 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 435 | Line 389 | $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \fra
389   approach than they are in the Taylor-shifted method.
390  
391   For the gradient shifted (GSF) method with the undamped kernel,
392 < $v_{21}(r) = -\frac{1}{r^3} + \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
392 > $v_{21}(r) = -\frac{1}{r^3} - \frac{3 r}{r_c^4} + \frac{4}{r_c^3}$ and
393   $v_{22}(r) = \frac{3}{r^3} + \frac{9 r}{r_c^4} - \frac{12}{r_c^3}$.
394   Again, these functions go smoothly to zero as $r\rightarrow r_c$, and
395   because the Taylor expansion retains only one term, they are
# Line 444 | 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 479 | 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 525 | 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}
485 < \label{tab:pars}
486 < \caption{The parameters used in the systems used to evaluate the new
487 <  real-space methods.  The most comprehensive test was a liquid
488 <  composed of 2000 SSDQ molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
489 <  ions).  This test excercises all orders of the multipolar
536 <  interactions developed in the first paper.}
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 soft DQ liquid molecules with 48 dissolved ions (24 \ce{Na+} and 24 \ce{Cl-}
488 >    ions).  This test exercises all orders of the multipolar
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 546 | Line 499 | Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-
499      Soft Dipolar solid & 2.837 & 1.0   &  & 2.35 & & & & $10^4$  & 17.6 &17.6 & 0 \\
500   Soft Quadrupolar fluid & 3.051 & 0.152 &  &  & -1&-1&-2.5 & 18.0153 & 1.77&0.6145&1.155  \\
501   Soft Quadrupolar solid & 2.837 & 1.0   &  &  & -1&-1&-2.5 & $10^4$  & 17.6&17.6&0 \\
502 <      SSDQ water  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
502 >      Soft DQ liquid  & 3.051 & 0.152 &  & 2.35 & -1.35&0&-0.68 & 18.0153 & 1.77&0.6145&1.155 \\
503                \ce{Na+} & 2.579 & 0.118 & +1& & & & & 22.99 & & &\\
504                \ce{Cl-} & 4.445 & 0.1   & -1& & & & & 35.4527& & & \\ \hline
505   \end{tabularx}
# Line 558 | Line 511 | relatively strict translational order.  The SSDQ model
511   charges in addition to the multipolar fluid.  The solid-phase
512   parameters were chosen so that the systems can explore some
513   orientational freedom for the multipolar sites, while maintaining
514 < relatively strict translational order.  The SSDQ model used here is
515 < not a particularly accurate water model, but it does test
516 < dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
517 < interactions at roughly the same magnitudes. The last test case, SSDQ
518 < water with dissolved ions, exercises \textit{all} levels of the
519 < multipole-multipole interactions we have derived so far and represents
520 < the most complete test of the new methods.
514 > relatively strict translational order.  The soft DQ liquid model used
515 > here based loosely on the SSDQO water
516 > model,\cite{Ichiye10_1,Ichiye10_2,Ichiye10_3} but is not itself a
517 > particularly accurate water model.  However, the soft DQ model does
518 > test dipole-dipole, dipole-quadrupole, and quadrupole-quadrupole
519 > interactions at roughly the same magnitudes. The last test case, a
520 > soft DQ liquid with dissolved ions, exercises \textit{all} levels of
521 > the multipole-multipole interactions we have derived so far and
522 > represents the most complete test of the new methods.
523  
524   In the following section, we present results for the total
525   electrostatic energy, as well as the electrostatic contributions to
# Line 595 | Line 550 | approximations.\cite{Smith82,Smith98} In all cases, th
550   with a reciprocal space cutoff, $k_\mathrm{max} = 7$.  Our version of
551   the Ewald sum is a re-implementation of the algorithm originally
552   proposed by Smith that does not use the particle mesh or smoothing
553 < approximations.\cite{Smith82,Smith98} In all cases, the quantities
554 < being compared are the electrostatic contributions to energies, force,
555 < and torques.  All other contributions to these quantities (i.e. from
556 < Lennard-Jones interactions) are removed prior to the comparisons.
553 > approximations.\cite{Smith82,Smith98} This implementation was tested
554 > extensively against the analytic energy constants for the multipolar
555 > lattices that are discussed in reference \onlinecite{PaperI}.  In all
556 > cases discussed below, the quantities being compared are the
557 > electrostatic contributions to energies, force, and torques.  All
558 > other contributions to these quantities (i.e. from Lennard-Jones
559 > interactions) are removed prior to the comparisons.
560  
561   The convergence parameter ($\alpha$) also plays a role in the balance
562   of the real-space and reciprocal-space portions of the Ewald
# Line 639 | Line 597 | ensemble.  We collected 250 different configurations a
597   simulations, each system was created with 2,048 randomly-oriented
598   molecules.  These were equilibrated at a temperature of 300K for 1 ns.
599   Each system was then simulated for 1 ns in the microcanonical (NVE)
600 < ensemble.  We collected 250 different configurations at equal time
601 < intervals. For the liquid system that included ionic species, we
602 < converted 48 randomly-distributed molecules into 24 \ce{Na+} and 24
603 < \ce{Cl-} ions and re-equilibrated. After equilibration, the system was
604 < run under the same conditions for 1 ns. A total of 250 configurations
605 < were collected. In the following comparisons of energies, forces, and
606 < torques, the Lennard-Jones potentials were turned off and only the
607 < purely electrostatic quantities were compared with the same values
608 < obtained via the Ewald sum.
600 > ensemble with the Dullweber, Leimkuhler, and McLachlan (DLM)
601 > symplectic splitting integrator using 1 fs
602 > timesteps.\cite{Dullweber1997} We collected 250 different
603 > configurations at equal time intervals. For the liquid system that
604 > included ionic species, we converted 48 randomly-distributed molecules
605 > into 24 \ce{Na+} and 24 \ce{Cl-} ions and re-equilibrated. After
606 > equilibration, the system was run under the same conditions for 1
607 > ns. A total of 250 configurations were collected. In the following
608 > comparisons of energies, forces, and torques, the Lennard-Jones
609 > potentials were turned off and only the purely electrostatic
610 > quantities were compared with the same values obtained via the Ewald
611 > sum.
612  
613   \subsection{Accuracy of Energy Differences, Forces and Torques}
614   The pairwise summation techniques (outlined above) were evaluated for
# Line 663 | Line 624 | with the multipolar Ewald reference method.  Unitary r
624   Since none of the real-space methods provide exact energy differences,
625   we used least square regressions analysis for the six different
626   molecular systems to compare $\Delta E$ from Hard, SP, GSF, and TSF
627 < with the multipolar Ewald reference method.  Unitary results for both
628 < the correlation (slope) and correlation coefficient for these
629 < regressions indicate perfect agreement between the real-space method
630 < and the multipolar Ewald sum.
627 > with the multipolar Ewald reference method.  A result of unity for
628 > both the correlation (slope) and coefficient of determination ($R^2$)
629 > for these regressions would indicate perfect agreement between the
630 > real-space method and the multipolar Ewald sum.
631  
632   Molecular systems were run long enough to explore independent
633   configurations and 250 configurations were recorded for comparison.
# Line 685 | Line 646 | force and torque vectors. Fisher developed a probablit
646   simulations.  Because the real space methods reweight the different
647   orientational contributions to the energies, it is also important to
648   understand how the methods impact the \textit{directionality} of the
649 < force and torque vectors. Fisher developed a probablity density
649 > force and torque vectors. Fisher developed a probability density
650   function to analyse directional data sets,
651   \begin{equation}
652   p_f(\theta) = \frac{\kappa}{2 \sinh\kappa}\sin\theta e^{\kappa \cos\theta}
# Line 699 | Line 660 | evaluated,
660   the forces obtained via the Ewald sum and the real-space methods were
661   evaluated,
662   \begin{equation}
663 < \cos\theta_i =  \frac{\vec{f}_i^\mathrm{~Ewald} \cdot
664 <  \vec{f}_i^\mathrm{~GSF}}{\left|\vec{f}_i^\mathrm{~Ewald}\right| \left|\vec{f}_i^\mathrm{~GSF}\right|}
663 >  \cos\theta_i =  \frac{\mathbf{f}_i^\mathrm{~Ewald} \cdot
664 >    \mathbf{f}_i^\mathrm{~GSF}}{\left|\mathbf{f}_i^\mathrm{~Ewald}\right| \left|\mathbf{f}_i^\mathrm{~GSF}\right|}
665   \end{equation}
666   The total angular displacement of the vectors was calculated as,
667   \begin{equation}
# Line 723 | Line 684 | system of 2000 SSDQ water molecules with 24 \ce{Na+} a
684  
685   \subsection{Energy conservation}
686   To test conservation the energy for the methods, the mixed molecular
687 < system of 2000 SSDQ water molecules with 24 \ce{Na+} and 24 \ce{Cl-}
688 < ions was run for 1 ns in the microcanonical ensemble at an average
689 < temperature of 300K.  Each of the different electrostatic methods
690 < (Ewald, Hard, SP, GSF, and TSF) was tested for a range of different
691 < damping values. The molecular system was started with same initial
692 < positions and velocities for all cutoff methods. The energy drift
693 < ($\delta E_1$) and standard deviation of the energy about the slope
694 < ($\delta E_0$) were evaluated from the total energy of the system as a
695 < function of time.  Although both measures are valuable at
687 > system of 2000 soft DQ liquid molecules with 24 \ce{Na+} and 24
688 > \ce{Cl-} ions was run for 1 ns in the microcanonical ensemble at an
689 > average temperature of 300K.  Each of the different electrostatic
690 > methods (Ewald, Hard, SP, GSF, and TSF) was tested for a range of
691 > different damping values. The molecular system was started with same
692 > initial positions and velocities for all cutoff methods. The energy
693 > drift ($\delta E_1$) and standard deviation of the energy about the
694 > slope ($\delta E_0$) were evaluated from the total energy of the
695 > system as a function of time.  Although both measures are valuable at
696   investigating new methods for molecular dynamics, a useful interaction
697   model must allow for long simulation times with minimal energy drift.
698  
699   \section{\label{sec:result}RESULTS}
700   \subsection{Configurational energy differences}
740 %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.
741 %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.
742 %\begin{figure}[h!]
743 %        \centering
744 %        \includegraphics[width=0.50 \textwidth]{rcutConvergence_dipolarLiquid-crop.pdf}
745 %        \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.  }
746 %        \label{fig:rcutConvergence_dipolarLiquid}
747 %    \end{figure}
748 %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.
749 %\begin{figure}
750 %        \centering
751 %        \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
752 %        \label{fig:barGraph1}
753 %        \end{figure}
754 %        \begin{figure}
755 %        \centering
756 %       \includegraphics[width=0.45 \textwidth]{slopeComparision_undamped.pdf}
757 %        \caption{}
758      
759 %        \label{fig:barGraph2}
760 %      \end{figure}
761 %The correlation coefficient ($R^2$) and slope of the linear
762 %regression plots for the energy differences for all six different
763 %molecular systems is shown in figure 4a and 4b.The plot shows that
764 %the correlation coefficient improves for the SP cutoff method as
765 %compared to the undamped hard cutoff method in the case of SSDQC,
766 %SSDQ, dipolar crystal, and dipolar liquid. For the quadrupolar
767 %crystal and liquid, the correlation coefficient is almost unchanged
768 %and close to 1.  The correlation coefficient is smallest (0.696276
769 %for $r_c$ = 9 $A^\circ$) for the SSDQC liquid because of the presence of
770 %charge-charge and charge-multipole interactions. Since the
771 %charge-charge and charge-multipole interaction is long ranged, there
772 %is huge deviation of correlation coefficient from 1. Similarly, the
773 %quarupole–quadrupole interaction is short ranged ($\sim 1/r^6$) with
774 %compared to interactions in the other multipolar systems, thus the
775 %correlation coefficient very close to 1 even for hard cutoff
776 %method. The idea of placing image multipole on the surface of the
777 %cutoff sphere improves the correlation coefficient and makes it close
778 %to 1 for all types of multipolar systems. Similarly the slope is
779 %hugely deviated from the correct value for the lower order
780 %multipole-multipole interaction and slightly deviated for higher
781 %order multipole – multipole interaction. The SP method improves both
782 %correlation coefficient ($R^2$) and slope significantly in SSDQC and
783 %dipolar systems.  The Slope is found to be deviated more in dipolar
784 %crystal as compared to liquid which is associated with the large
785 %fluctuation in the electrostatic energy in crystal. The GSF also
786 %produced better values of correlation coefficient and slope with the
787 %proper selection of the damping alpha (Interested reader can consult
788 %accompanying supporting material). The TSF method gives good value of
789 %correlation coefficient for the dipolar crystal, dipolar liquid,
790 %SSDQ, and SSDQC (not for the quadrupolar crystal and liquid) but the
791 %regression slopes are significantly deviated.
701  
702   \begin{figure}
703    \centering
704 <  \includegraphics[width=0.85\linewidth]{energyPlot_slopeCorrelation_combined.eps}
704 >  \includegraphics[width=0.6\linewidth]{energyPlot_slopeCorrelation_combined.eps}
705    \caption{Statistical analysis of the quality of configurational
706      energy differences for the real-space electrostatic methods
707      compared with the reference Ewald sum.  Results with a value equal
708      to 1 (dashed line) indicate $\Delta E$ values indistinguishable
709      from those obtained using the multipolar Ewald sum.  Different
710      values of the cutoff radius are indicated with different symbols
711 <    (9\AA\ = circles, 12\AA\ = squares, and 15\AA\ = inverted
712 <    triangles).}
804 <  \label{fig:slopeCorr_energy}
711 >    (9~\AA\ = circles, 12~\AA\ = squares, and 15~\AA\ = inverted
712 >    triangles).\label{fig:slopeCorr_energy}}
713   \end{figure}
714  
715 < The combined correlation coefficient and slope for all six systems is
716 < shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the methods
717 < reproduce the Ewald configurational energy differences with remarkable
718 < fidelity.  Undamped hard cutoffs introduce a significant amount of
719 < random scatter in the energy differences which is apparent in the
720 < reduced value of the correlation coefficient for this method.  This
721 < can be easily understood as configurations which exhibit small
722 < traversals of a few dipoles or quadrupoles out of the cutoff sphere
723 < will see large energy jumps when hard cutoffs are used.  The
724 < orientations of the multipoles (particularly in the ordered crystals)
725 < mean that these energy jumps can go in either direction, producing a
726 < significant amount of random scatter, but no systematic error.
715 > The combined coefficient of determination and slope for all six
716 > systems is shown in Figure ~\ref{fig:slopeCorr_energy}.  Most of the
717 > methods reproduce the Ewald configurational energy differences with
718 > remarkable fidelity.  Undamped hard cutoffs introduce a significant
719 > amount of random scatter in the energy differences which is apparent
720 > in the reduced value of $R^2$ for this method.  This can be easily
721 > understood as configurations which exhibit small traversals of a few
722 > dipoles or quadrupoles out of the cutoff sphere will see large energy
723 > jumps when hard cutoffs are used.  The orientations of the multipoles
724 > (particularly in the ordered crystals) mean that these energy jumps
725 > can go in either direction, producing a significant amount of random
726 > scatter, but no systematic error.
727  
728   The TSF method produces energy differences that are highly correlated
729   with the Ewald results, but it also introduces a significant
# Line 826 | Line 734 | excellent fidelity, particularly for moderate damping
734   effect, particularly for the crystalline systems.
735  
736   Both the SP and GSF methods appear to reproduce the Ewald results with
737 < excellent fidelity, particularly for moderate damping ($\alpha =
738 < 0.1-0.2$\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
739 < 12$\AA).  With the exception of the undamped hard cutoff, and the TSF
737 > excellent fidelity, particularly for moderate damping ($\alpha \approx
738 > 0.2$~\AA$^{-1}$) and with a commonly-used cutoff value ($r_c =
739 > 12$~\AA).  With the exception of the undamped hard cutoff, and the TSF
740   method with short cutoffs, all of the methods would be appropriate for
741   use in Monte Carlo simulations.
742  
# Line 858 | Line 766 | commonly-used cutoff values ($r_c = 12$\AA).  The TSF
766   energy conservation issues, and this perturbation is evident in the
767   statistics accumulated for the molecular forces.  The GSF
768   perturbations are minimal, particularly for moderate damping and
769 < commonly-used cutoff values ($r_c = 12$\AA).  The TSF method shows
770 < reasonable agreement in the correlation coefficient but again the
771 < systematic error in the forces is concerning if replication of Ewald
864 < forces is desired.
769 > commonly-used cutoff values ($r_c = 12$~\AA).  The TSF method shows
770 > reasonable agreement in $R^2$, but again the systematic error in the
771 > forces is concerning if replication of Ewald forces is desired.
772  
773 + It is important to note that the forces and torques from the SP and
774 + the Hard cutoffs are not identical. The SP method shifts each
775 + orientational contribution separately (e.g. the dipole-dipole dot
776 + product is shifted by a different function than the dipole-distance
777 + products), while the hard cutoff contains no orientation-dependent
778 + shifting.  The forces and torques for these methods therefore diverge
779 + for multipoles even though the forces for point charges are identical.
780 +
781   \begin{figure}
782    \centering
783    \includegraphics[width=0.6\linewidth]{forcePlot_slopeCorrelation_combined.eps}
# Line 871 | Line 786 | forces is desired.
786      the reference Ewald sum. Results with a value equal to 1 (dashed
787      line) indicate force magnitude values indistinguishable from those
788      obtained using the multipolar Ewald sum.  Different values of the
789 <    cutoff radius are indicated with different symbols (9\AA\ =
790 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles). }
791 <  \label{fig:slopeCorr_force}
789 >    cutoff radius are indicated with different symbols (9~\AA\ =
790 >    circles, 12~\AA\ = squares, and 15~\AA\ = inverted
791 >    triangles).\label{fig:slopeCorr_force}}
792   \end{figure}
793  
794  
# Line 885 | Line 800 | forces is desired.
800      the reference Ewald sum. Results with a value equal to 1 (dashed
801      line) indicate force magnitude values indistinguishable from those
802      obtained using the multipolar Ewald sum.  Different values of the
803 <    cutoff radius are indicated with different symbols (9\AA\ =
804 <    circles, 12\AA\ = squares, and 15\AA\ = inverted triangles).}
805 <  \label{fig:slopeCorr_torque}
803 >    cutoff radius are indicated with different symbols (9~\AA\ =
804 >    circles, 12~\AA\ = squares, and 15~\AA\ = inverted
805 >    triangles).\label{fig:slopeCorr_torque}}
806   \end{figure}
807  
808   The torques (Fig. \ref{fig:slopeCorr_torque}) appear to be
# Line 901 | Line 816 | of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1
816   reproduces the torques in quite good agreement with the Ewald sum.
817   The other real-space methods can cause some deviations, but excellent
818   agreement with the Ewald sum torques is recovered at moderate values
819 < of the damping coefficient ($\alpha = 0.1-0.2$\AA$^{-1}$) and cutoff
820 < radius ($r_c \ge 12$\AA).  The TSF method exhibits only fair agreement
819 > of the damping coefficient ($\alpha \approx 0.2$~\AA$^{-1}$) and cutoff
820 > radius ($r_c \ge 12$~\AA).  The TSF method exhibits only fair agreement
821   in the slope when compared with the Ewald torques even for larger
822   cutoff radii.  It appears that the severity of the perturbations in
823   the TSF method are most in evidence for the torques.
# Line 914 | Line 829 | directionality is shown in terms of circular variance
829   these quantities. Force and torque vectors for all six systems were
830   analyzed using Fisher statistics, and the quality of the vector
831   directionality is shown in terms of circular variance
832 < ($\mathrm{Var}(\theta)$) in figure
833 < \ref{fig:slopeCorr_circularVariance}. The force and torque vectors
834 < from the new real-space methods exhibit nearly-ideal Fisher probability
835 < distributions (Eq.~\ref{eq:pdf}). Both the hard and SP cutoff methods
836 < exhibit the best vectorial agreement with the Ewald sum. The force and
837 < torque vectors from GSF method also show good agreement with the Ewald
838 < method, which can also be systematically improved by using moderate
839 < damping and a reasonable cutoff radius. For $\alpha = 0.2$ and $r_c =
840 < 12$\AA, we observe $\mathrm{Var}(\theta) = 0.00206$, which corresponds
841 < to a distribution with 95\% of force vectors within $6.37^\circ$ of
842 < the corresponding Ewald forces. The TSF method produces the poorest
843 < agreement with the Ewald force directions.
832 > ($\mathrm{Var}(\theta)$) in
833 > Fig. \ref{fig:slopeCorr_circularVariance}. The force and torque
834 > vectors from the new real-space methods exhibit nearly-ideal Fisher
835 > probability distributions (Eq.~\ref{eq:pdf}). Both the hard and SP
836 > cutoff methods exhibit the best vectorial agreement with the Ewald
837 > sum. The force and torque vectors from GSF method also show good
838 > agreement with the Ewald method, which can also be systematically
839 > improved by using moderate damping and a reasonable cutoff radius. For
840 > $\alpha = 0.2$~\AA$^{-1}$ and $r_c = 12$~\AA, we observe
841 > $\mathrm{Var}(\theta) = 0.00206$, which corresponds to a distribution
842 > with 95\% of force vectors within $6.37^\circ$ of the corresponding
843 > Ewald forces. The TSF method produces the poorest agreement with the
844 > Ewald force directions.
845  
846   Torques are again more perturbed than the forces by the new real-space
847   methods, but even here the variance is reasonably small.  For the same
848 < method (GSF) with the same parameters ($\alpha = 0.2, r_c = 12$\AA),
849 < the circular variance was 0.01415, corresponds to a distribution which
850 < has 95\% of torque vectors are within $16.75^\circ$ of the Ewald
851 < results. Again, the direction of the force and torque vectors can be
852 < systematically improved by varying $\alpha$ and $r_c$.
848 > method (GSF) with the same parameters ($\alpha = 0.2$~\AA$^{-1}$, $r_c
849 > = 12$~\AA), the circular variance was 0.01415, corresponds to a
850 > distribution which has 95\% of torque vectors are within $16.75^\circ$
851 > of the Ewald results. Again, the direction of the force and torque
852 > vectors can be systematically improved by varying $\alpha$ and $r_c$.
853  
854   \begin{figure}
855    \centering
# Line 944 | Line 860 | systematically improved by varying $\alpha$ and $r_c$.
860      indicates direction of the force or torque vectors are
861      indistinguishable from those obtained from the Ewald sum. Here
862      different symbols represent different values of the cutoff radius
863 <    (9 \AA\ = circle, 12 \AA\ = square, 15 \AA\ = inverted triangle)}
948 <  \label{fig:slopeCorr_circularVariance}
863 >    (9~\AA\ = circle, 12~\AA\ = square, 15~\AA\ = inverted triangle)\label{fig:slopeCorr_circularVariance}}
864   \end{figure}
865  
866   \subsection{Energy conservation\label{sec:conservation}}
867  
868   We have tested the conservation of energy one can expect to see with
869 < the new real-space methods using the SSDQ water model with a small
869 > the new real-space methods using the soft DQ liquid model with a small
870   fraction of solvated ions. This is a test system which exercises all
871   orders of multipole-multipole interactions derived in the first paper
872   in this series and provides the most comprehensive test of the new
873 < methods.  A liquid-phase system was created with 2000 water molecules
874 < and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
875 < temperature of 300K.  After equilibration, this liquid-phase system
876 < was run for 1 ns under the Ewald, Hard, SP, GSF, and TSF methods with
877 < a cutoff radius of 12\AA.  The value of the damping coefficient was
878 < also varied from the undamped case ($\alpha = 0$) to a heavily damped
879 < case ($\alpha = 0.3$ \AA$^{-1}$) for all of the real space methods.  A
880 < sample was also run using the multipolar Ewald sum with the same
881 < real-space cutoff.
873 > methods.  A liquid-phase system was created with 2000 liquid-phase
874 > molecules and 48 dissolved ions at a density of 0.98 g cm$^{-3}$ and a
875 > temperature of 300K.  After equilibration in the canonical (NVT)
876 > ensemble using a Nos\'e-Hoover thermostat, this liquid-phase system
877 > was run for 1 ns in the microcanonical (NVE) ensemble under the Ewald,
878 > Hard, SP, GSF, and TSF methods with a cutoff radius of 12~\AA.  The
879 > value of the damping coefficient was also varied from the undamped
880 > case ($\alpha = 0$) to a heavily damped case ($\alpha =
881 > 0.3$~\AA$^{-1}$) for all of the real space methods.  A sample was also
882 > run using the multipolar Ewald sum with the same real-space cutoff.
883  
884   In figure~\ref{fig:energyDrift} we show the both the linear drift in
885   energy over time, $\delta E_1$, and the standard deviation of energy
# Line 979 | Line 895 | than the multipolar Ewald sum, even when utilizing a r
895  
896   We note that for all tested values of the cutoff radius, the new
897   real-space methods can provide better energy conservation behavior
898 < than the multipolar Ewald sum, even when utilizing a relatively large
899 < $k$-space cutoff values.
898 > than the multipolar Ewald sum, even when relatively large $k$-space
899 > cutoff values are utilized.
900  
901   \begin{figure}
902    \centering
903    \includegraphics[width=\textwidth]{newDrift_12.eps}
904 < \label{fig:energyDrift}        
905 < \caption{Analysis of the energy conservation of the real-space
906 <  electrostatic methods. $\delta \mathrm{E}_1$ is the linear drift in
907 <  energy over time (in kcal / mol / particle / ns) and $\delta
908 <  \mathrm{E}_0$ is the standard deviation of energy fluctuations
909 <  around this drift (in kcal / mol / particle).  All simulations were
910 <  of a 2000-molecule simulation of SSDQ water with 48 ionic charges at
911 <  300 K starting from the same initial configuration. All runs
996 <  utilized the same real-space cutoff, $r_c = 12$\AA.}
904 >  \caption{Energy conservation of the real-space methods for the soft
905 >    DQ liauid / ion system. $\delta \mathrm{E}_1$ is the linear drift
906 >    in energy over time (in kcal/mol/particle/ns) and $\delta
907 >    \mathrm{E}_0$ is the standard deviation of energy fluctuations
908 >    around this drift (in kcal/mol/particle).  Points that appear in
909 >    the green region at the bottom exhibit better energy conservation
910 >    than would be obtained using common parameters for Ewald-based
911 >    electrostatics.\label{fig:energyDrift}}
912   \end{figure}
913  
914 + \subsection{Reproduction of Structural \& Dynamical Features\label{sec:structure}}
915 + The most important test of the modified interaction potentials is the
916 + fidelity with which they can reproduce structural features and
917 + dynamical properties in a liquid.  One commonly-utilized measure of
918 + structural ordering is the pair distribution function, $g(r)$, which
919 + measures local density deviations in relation to the bulk density.  In
920 + the electrostatic approaches studied here, the short-range repulsion
921 + from the Lennard-Jones potential is identical for the various
922 + electrostatic methods, and since short range repulsion determines much
923 + of the local liquid ordering, one would not expect to see many
924 + differences in $g(r)$.  Indeed, the pair distributions are essentially
925 + identical for all of the electrostatic methods studied (for each of
926 + the different systems under investigation).
927  
928 + % An example of this agreement for the soft DQ liquid/ion system is
929 + % shown in Fig. \ref{fig:gofr}.
930 +
931 + % \begin{figure}
932 + %   \centering
933 + %   \includegraphics[width=\textwidth]{gofr_ssdqc.eps}
934 + % \caption{The pair distribution functions, $g(r)$, for the SSDQ
935 + %   water/ion system obtained using the different real-space methods are
936 + %   essentially identical with the result from the Ewald
937 + %   treatment.\label{fig:gofr}}
938 + % \end{figure}
939 +
940 + There is a minor over-structuring of the first solvation shell when
941 + using TSF or when overdamping with any of the real-space methods.
942 + With moderate damping, GSF and SP produce pair distributions that are
943 + identical (within numerical noise) to their Ewald counterparts.  The
944 + degree of over-structuring can be measured most easily using the
945 + coordination number,
946 + \begin{equation}
947 + n_C = 4\pi\rho \int_{0}^{a}r^2\text{g}(r)dr,
948 + \end{equation}
949 + where $\rho$ is the number density of the site-site pair interactions,
950 + and $a$ is the radial location of the minima following the first peak
951 + in $g(r)$ ($a = 4.2$~\AA\  for the soft DQ liquid / ion system).  The
952 + coordination number is shown as a function of the damping coefficient
953 + for all of the real space methods in Fig. \ref{fig:Props}.
954 +
955 + A more demanding test of modified electrostatics is the average value
956 + of the electrostatic energy $\langle U_\mathrm{elect} \rangle / N$
957 + which is obtained by sampling the liquid-state configurations
958 + experienced by a liquid evolving entirely under the influence of each
959 + of the methods.  In Fig. \ref{fig:Props} we demonstrate how $\langle
960 + U_\mathrm{elect} \rangle / N$ varies with the damping parameter,
961 + $\alpha$, for each of the methods.
962 +
963 + As in the crystals studied in the first paper, damping is important
964 + for converging the mean electrostatic energy values, particularly for
965 + the two shifted force methods (GSF and TSF).  A value of $\alpha
966 + \approx 0.2$~\AA$^{-1}$ is sufficient to converge the SP and GSF
967 + energies with a cutoff of 12 \AA, while shorter cutoffs require more
968 + dramatic damping ($\alpha \approx 0.28$~\AA$^{-1}$ for $r_c = 9$~\AA).
969 + Overdamping the real-space electrostatic methods occurs with $\alpha >
970 + 0.3$~\AA$^{-1}$, causing the estimate of the electrostatic energy to
971 + drop below the Ewald results.
972 +
973 + These ``optimal'' values of the damping coefficient for structural
974 + features are similar to those observed for DSF electrostatics for
975 + purely point-charge systems, and the range $\alpha= 0.175 \rightarrow
976 + 0.225$~\AA$^{-1}$ for $r_c = 12$~\AA\ appears to be an excellent
977 + compromise for mixed charge/multipolar systems.
978 +
979 + To test the fidelity of the electrostatic methods at reproducing
980 + \textit{dynamics} in a multipolar liquid, it is also useful to look at
981 + transport properties, particularly the diffusion constant,
982 + \begin{equation}
983 + D = \lim_{t \rightarrow \infty} \frac{1}{6 t} \langle \left|
984 +  \mathbf{r}(t) -\mathbf{r}(0) \right|^2 \rangle
985 + \label{eq:diff}
986 + \end{equation}
987 + which measures long-time behavior and is sensitive to the forces on
988 + the multipoles. The self-diffusion constants (D) were calculated from
989 + linear fits to the long-time portion of the mean square displacement,
990 + $\langle r^{2}(t) \rangle$.\cite{Allen87} In Fig. \ref{fig:Props} we
991 + demonstrate how the diffusion constant depends on the choice of
992 + real-space methods and the damping coefficient.  Both the SP and GSF
993 + methods can obtain excellent agreement with Ewald again using moderate
994 + damping.
995 +
996 + In addition to translational diffusion, orientational relaxation times
997 + were calculated for comparisons with the Ewald simulations and with
998 + experiments. These values were determined by calculating the
999 + orientational time correlation function,
1000 + \begin{equation}
1001 + C_l^\gamma(t) = \left\langle P_l\left[\hat{\mathbf{A}}_\gamma(t)
1002 +                \cdot\hat{\mathbf{A}}_\gamma(0)\right]\right\rangle,
1003 + \label{eq:OrientCorr}
1004 + \end{equation}
1005 + from the same 350 ps microcanonical trajectories that were used for
1006 + translational diffusion.  Here, $P_l$ is the Legendre polynomial of
1007 + order $l$ and $\hat{\mathbf{A}}_\gamma$ is the unit vector for body
1008 + axis $\gamma$.  The reference frame used for our sample dipolar
1009 + systems has the $z$-axis running along the dipoles, and for the soft
1010 + DQ liquid model, the $y$-axis connects the two implied hydrogen-like
1011 + positions.  From the orientation autocorrelation functions, we can
1012 + obtain time constants for rotational relaxation either by fitting to a
1013 + multi-exponential model for the orientational relaxation, or by
1014 + integrating the correlation functions.
1015 +
1016 + In a good model for water, the orientational decay times would be
1017 + comparable to water orientational relaxation times from nuclear
1018 + magnetic resonance (NMR). The relaxation constant obtained from
1019 + $C_2^y(t)$ is normally of experimental interest because it describes
1020 + the relaxation of the principle axis connecting the hydrogen
1021 + atoms. Thus, $C_2^y(t)$ can be compared to the intermolecular portion
1022 + of the dipole-dipole relaxation from a proton NMR signal and can
1023 + provide an estimate of the NMR relaxation time constant.\cite{Impey82}
1024 + In Fig. \ref{fig:Props} we compare the $\tau_2^y$ and $\tau_2^z$
1025 + values for the various real-space methods over a range of different
1026 + damping coefficients.  The rotational relaxation for the $z$ axis
1027 + primarily probes the torques on the dipoles, while the relaxation for
1028 + the $y$ axis is sensitive primarily to the quadrupolar torques.
1029 +
1030 + \begin{figure}
1031 +  \includegraphics[width=\textwidth]{properties.eps}
1032 +  \caption{Comparison of the structural and dynamic properties for the
1033 +    combined multipolar liquid (soft DQ liquid + ions) for all of the
1034 +    real-space methods with $r_c = 12$~\AA. Electrostatic energies,
1035 +    $\langle U_\mathrm{elect} \rangle / N$ (in kcal / mol),
1036 +    coordination numbers, $n_C$, diffusion constants (in $10^{-5}
1037 +    \mathrm{cm}^2\mathrm{s}^{-1}$), and rotational correlation times
1038 +    (in ps) all show excellent agreement with Ewald results for
1039 +    damping coefficients in the range $\alpha= 0.175 \rightarrow
1040 +    0.225$~\AA$^{-1}$. \label{fig:Props}}
1041 + \end{figure}
1042 +
1043 + In Fig. \ref{fig:Props} it appears that values for $D$, $\tau_2^y$,
1044 + and $\tau_2^z$ using the Ewald sum are reproduced with excellent
1045 + fidelity by the GSF and SP methods.  All of the real space methods can
1046 + be \textit{overdamped}, which reduces the effective range of multipole
1047 + interactions, causing structural and dynamical changes from the
1048 + correct behavior.  Because overdamping weakens orientational
1049 + preferences between adjacent molecules, it manifests as too-rapid
1050 + orientational decay coupled with faster diffusion and
1051 + over-coordination of the liquid.  Underdamping is less problematic for
1052 + the SP and GSF methods, as their structural and dynamical properties
1053 + still reproduce the Ewald results even in the completely undamped
1054 + ($\alpha = 0$) case.  An optimal range for the electrostatic damping
1055 + parameter appears to be $\alpha= 0.175 \rightarrow 0.225$~\AA$^{-1}$
1056 + for $r_c = 12$~\AA, which similar to the optimal range found for the
1057 + damped shifted force potential for point charges.\cite{Fennell:2006lq}
1058 +
1059   \section{CONCLUSION}
1060   In the first paper in this series, we generalized the
1061   charge-neutralized electrostatic energy originally developed by Wolf
# Line 1009 | Line 1068 | We also developed two natural extensions of the damped
1068   distance that prevents its use in molecular dynamics.
1069  
1070   We also developed two natural extensions of the damped shifted-force
1071 < (DSF) model originally proposed by Fennel and
1072 < Gezelter.\cite{Fennell:2006lq} The GSF and TSF approaches provide
1073 < smooth truncation of energies, forces, and torques at the real-space
1074 < cutoff, and both converge to DSF electrostatics for point-charge
1075 < interactions.  The TSF model is based on a high-order truncated Taylor
1076 < expansion which can be relatively perturbative inside the cutoff
1077 < sphere.  The GSF model takes the gradient from an images of the
1078 < interacting multipole that has been projected onto the cutoff sphere
1079 < to derive shifted force and torque expressions, and is a significantly
1080 < more gentle approach.
1071 > (DSF) model originally proposed by Zahn {\it et al.} and extended by
1072 > Fennell and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
1073 > approaches provide smooth truncation of energies, forces, and torques
1074 > at the real-space cutoff, and both converge to DSF electrostatics for
1075 > point-charge interactions.  The TSF model is based on a high-order
1076 > truncated Taylor expansion which can be relatively perturbative inside
1077 > the cutoff sphere.  The GSF model takes the gradient from an images of
1078 > the interacting multipole that has been projected onto the cutoff
1079 > sphere to derive shifted force and torque expressions, and is a
1080 > significantly more gentle approach.
1081  
1082 < Of the two newly-developed shifted force models, the GSF method
1083 < produced quantitative agreement with Ewald energy, force, and torques.
1084 < It also performs well in conserving energy in MD simulations.  The
1085 < Taylor-shifted (TSF) model provides smooth dynamics, but these take
1086 < place on a potential energy surface that is significantly perturbed
1087 < from Ewald-based electrostatics.  
1082 > The GSF method produces quantitative agreement with Ewald energies,
1083 > forces, and torques.  It also performs well in conserving energy in MD
1084 > simulations.  The Taylor-shifted (TSF) model provides smooth dynamics,
1085 > but these take place on a potential energy surface that is
1086 > significantly perturbed from Ewald-based electrostatics.  Because it
1087 > performs relatively poorly compared with GSF, it may seem odd that
1088 > that the TSF model was included in this work.  However, the functional
1089 > forms derived for the SP and GSF methods depend on the separation of
1090 > orientational contributions that were made visible by the Taylor
1091 > series of the electrostatic kernel at the cutoff radius. The TSF
1092 > method also has the unique property that a large number of derivatives
1093 > can be made to vanish at the cutoff radius.  This property has proven
1094 > useful in past treatments of the corrections to the Clausius-Mossotti
1095 > fluctuation formula for dielectric constants.\cite{Izvekov:2008wo}
1096  
1097 < % The direct truncation of any electrostatic potential energy without
1098 < % multipole neutralization creates large fluctuations in molecular
1099 < % simulations.  This fluctuation in the energy is very large for the case
1100 < % of crystal because of long range of multipole ordering (Refer paper
1101 < % I).\cite{PaperI} This is also significant in the case of the liquid
1102 < % because of the local multipole ordering in the molecules. If the net
1103 < % multipole within cutoff radius neutralized within cutoff sphere by
1104 < % placing image multiples on the surface of the sphere, this fluctuation
1105 < % in the energy reduced significantly. Also, the multipole
1106 < % neutralization in the generalized SP method showed very good agreement
1107 < % with the Ewald as compared to direct truncation for the evaluation of
1041 < % the $\triangle E$ between the configurations.  In MD simulations, the
1042 < % energy conservation is very important. The conservation of the total
1043 < % energy can be ensured by i) enforcing the smooth truncation of the
1044 < % energy, force and torque in the cutoff radius and ii) making the
1045 < % energy, force and torque consistent with each other. The GSF and TSF
1046 < % methods ensure the consistency and smooth truncation of the energy,
1047 < % force and torque at the cutoff radius, as a result show very good
1048 < % total energy conservation. But the TSF method does not show good
1049 < % agreement in the absolute value of the electrostatic energy, force and
1050 < % torque with the Ewald.  The GSF method has mimicked Ewald’s force,
1051 < % energy and torque accurately and also conserved energy.
1097 > Reproduction of both structural and dynamical features in the liquid
1098 > systems is remarkably good for both the SP and GSF models.  Pair
1099 > distribution functions are essentially equivalent to the same
1100 > functions produced using Ewald-based electrostatics, and with moderate
1101 > damping, a structural feature that directly probes the electrostatic
1102 > interaction (e.g. the mean electrostatic potential energy) can also be
1103 > made quantitative.  Dynamical features are sensitive probes of the
1104 > forces and torques produced by these methods, and even though the
1105 > smooth behavior of forces is produced by perturbing the overall
1106 > potential, the diffusion constants and orientational correlation times
1107 > are quite close to the Ewald-based results.
1108  
1109   The only cases we have found where the new GSF and SP real-space
1110   methods can be problematic are those which retain a bulk dipole moment
# Line 1059 | Line 1115 | Based on the results of this work, the GSF method is a
1115   replaced by the bare electrostatic kernel, and the energies return to
1116   the expected converged values.
1117  
1118 < Based on the results of this work, the GSF method is a suitable and
1119 < efficient replacement for the Ewald sum for evaluating electrostatic
1120 < interactions in MD simulations.  Both methods retain excellent
1121 < fidelity to the Ewald energies, forces and torques.  Additionally, the
1122 < energy drift and fluctuations from the GSF electrostatics are better
1123 < than a multipolar Ewald sum for finite-sized reciprocal spaces.
1124 < Because they use real-space cutoffs with moderate cutoff radii, the
1125 < GSF and SP models approach $\mathcal{O}(N)$ scaling as the system size
1126 < increases.  Additionally, they can be made extremely efficient using
1127 < spline interpolations of the radial functions.  They require no
1072 < Fourier transforms or $k$-space sums, and guarantee the smooth
1073 < handling of energies, forces, and torques as multipoles cross the
1074 < real-space cutoff boundary.
1118 > Based on the results of this work, we can conclude that the GSF method
1119 > is a suitable and efficient replacement for the Ewald sum for
1120 > evaluating electrostatic interactions in modern MD simulations, and
1121 > the SP method would be an excellent choice for Monte Carlo
1122 > simulations where smooth forces and energy conservation are not
1123 > important.  Both the SP and GSF methods retain excellent fidelity to
1124 > the Ewald energies, forces and torques.  Additionally, the energy
1125 > drift and fluctuations from the GSF electrostatics are significantly
1126 > better than a multipolar Ewald sum for finite-sized reciprocal spaces,
1127 > and physical properties are reproduced accurately.
1128  
1129 + As in all purely pairwise cutoff methods, the SP, GSF and TSF methods
1130 + are expected to scale approximately {\it linearly} with system size,
1131 + and are easily parallelizable.  This should result in substantial
1132 + reductions in the computational cost of performing large simulations.
1133 + With the proper use of pre-computation and spline interpolation of the
1134 + radial functions, the real-space methods are essentially the same cost
1135 + as a simple real-space cutoff.  They require no Fourier transforms or
1136 + $k$-space sums, and guarantee the smooth handling of energies, forces,
1137 + and torques as multipoles cross the real-space cutoff boundary.
1138 +
1139 + We are not suggesting that there is any flaw with the Ewald sum; in
1140 + fact, it is the standard by which the SP, GSF, and TSF methods have
1141 + been judged in this work.  However, these results provide evidence
1142 + that in the typical simulations performed today, the Ewald summation
1143 + may no longer be required to obtain the level of accuracy most
1144 + researchers have come to expect.
1145 +
1146   \begin{acknowledgments}
1147    JDG acknowledges helpful discussions with Christopher
1148    Fennell. Support for this project was provided by the National

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines