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 4212 by gezelter, Mon Aug 18 14:54:19 2014 UTC vs.
Revision 4214 by gezelter, Wed Sep 10 21:06:56 2014 UTC

# Line 165 | 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
# Line 223 | Line 223 | cutoff sphere that are integral to the Wolf and DSF ap
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.
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.
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 approaches 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
245 < 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 257 | 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 343 | Line 343 | U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \
343   \cdot \mathbf{D}_{b}$) is treated differently than the dipole-distance
344   dot products:
345   \begin{equation}
346 < U_{D_{a}D_{b}}(r)= -\frac{1}{4\pi \epsilon_0} \left[ \left(
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)
# Line 372 | Line 372 | U_{D_{a}D_{b}}(r)  = U_{D_{a}D_{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_{a}D_{b}}(r)  = U_{D_{a}D_{b}}(r) -
376 < U_{D_{a}D_{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_{D_{a}D_{b}}(r_c).
378 >  \nabla U_{\mathbf{D}_{a}\mathbf{D}_{b}}(r_c).
379   \end{equation}
380   Here the lab-frame orientations of the two dipoles, $\mathbf{D}_{a}$ and $\mathbf{D}_{b}$, are retained at the cutoff distance
381   (although the signs are reversed for the dipole that has been
# Line 482 | Line 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 < \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 exercises all orders of the multipolar
489 <  interactions developed in the first paper.\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 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 499 | 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 511 | 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 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 619 | 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 655 | 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 679 | 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  
# Line 696 | Line 701 | model must allow for long simulation times with minima
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
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 729 | 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 761 | 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
767 < 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
# Line 782 | Line 786 | for multipoles even though the forces for point charge
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
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  
# Line 796 | Line 800 | for multipoles even though the forces for point charge
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
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  
# Line 812 | 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$~\AA).  The TSF method exhibits only fair agreement
821   in the slope when compared with the Ewald torques even for larger
822   cutoff radii.  It appears that the severity of the perturbations in
823   the TSF method are most in evidence for the torques.
# Line 825 | Line 829 | directionality is shown in terms of circular variance
829   these quantities. Force and torque vectors for all six systems were
830   analyzed using Fisher statistics, and the quality of the vector
831   directionality is shown in terms of circular variance
832 < ($\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 855 | 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)\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
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
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 ($\alpha = 0.3$
881 < \AA$^{-1}$) for all of the real space methods.  A sample was also run
882 < using the multipolar Ewald sum with the same real-space cutoff.
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 896 | Line 901 | cutoff values are utilized.
901   \begin{figure}
902    \centering
903    \includegraphics[width=\textwidth]{newDrift_12.eps}
904 <  \caption{Energy conservation of the real-space methods for the SSDQ
905 <    water/ion system. $\delta \mathrm{E}_1$ is the linear drift in
906 <    energy over time (in kcal/mol/particle/ns) and $\delta
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
# Line 918 | Line 923 | the different systems under investigation).  An exampl
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).  An example of this
922 < agreement for the SSDQ water/ion system is shown in
923 < Fig. \ref{fig:gofr}.
926 > the different systems under investigation).
927  
928 < \begin{figure}
929 <  \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}
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
# Line 942 | Line 948 | in $g(r)$ ($a = 4.2$ \AA\  for the SSDQ water/ion syst
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 SSDQ water/ion system).  The
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  
# Line 950 | Line 956 | of the methods.  In fig \ref{fig:Props} we demonstrate
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
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
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).
968 > dramatic damping ($\alpha \approx 0.28$~\AA$^{-1}$ for $r_c = 9$~\AA).
969   Overdamping the real-space electrostatic methods occurs with $\alpha >
970 < 0.3$, causing the estimate of the electrostatic energy to drop below
971 < the Ewald results.
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
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
# Line 981 | Line 987 | $\langle r^{2}(t) \rangle$.\cite{Allen87} In fig. \ref
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
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
# Line 1000 | Line 1006 | systems has the $z$-axis running along the dipoles, an
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 SSDQ
1010 < water model, the $y$-axis connects the two implied hydrogen atom
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 water model, the orientational decay times would be
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
# Line 1015 | Line 1021 | In fig. \ref{fig:Props} we compare the $\tau_2^y$ and
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 $\tau_2^z$
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 (SSDQ water + ions) for all of the
1034 <    real-space methods with $r_c = 12$\AA. Electrostatic energies,
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}}
1034 <  \includegraphics[width=\textwidth]{properties.eps}
1040 >    0.225$~\AA$^{-1}$. \label{fig:Props}}
1041   \end{figure}
1042  
1043   In Fig. \ref{fig:Props} it appears that values for $D$, $\tau_2^y$,
# Line 1046 | Line 1052 | parameter appears to be $\alpha= 0.175 \rightarrow 0.2
1052   the SP and GSF methods, as their structural and dynamical properties
1053   still reproduce the Ewald results even in the completely undamped
1054   ($\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
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}
# Line 1063 | Line 1069 | Fennel and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq}
1069  
1070   We also developed two natural extensions of the damped shifted-force
1071   (DSF) model originally proposed by Zahn {\it et al.} and extended by
1072 < Fennel and Gezelter.\cite{Zahn:2002hc,Fennell:2006lq} The GSF and TSF
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
# Line 1085 | Line 1091 | useful in past treatments of the corrections to the fl
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 fluctuation
1095 < formula for dielectric constants.\cite{Izvekov:2008wo}
1094 > useful in past treatments of the corrections to the Clausius-Mossotti
1095 > fluctuation formula for dielectric constants.\cite{Izvekov:2008wo}
1096  
1097   Reproduction of both structural and dynamical features in the liquid
1098   systems is remarkably good for both the SP and GSF models.  Pair

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines