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

Comparing trunk/ssdePaper/nptSSD.tex (file contents):
Revision 921 by gezelter, Mon Jan 12 16:20:53 2004 UTC vs.
Revision 1022 by chrisfen, Wed Feb 4 22:13:36 2004 UTC

# Line 139 | Line 139 | + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i
139   \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)
140   + s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf
141   \Omega}_j)]\ .
142 + \label{stickyfunction}
143   \end{equation}
144   Here, $\nu_0$ is a strength parameter for the sticky potential, and
145   $s$ and $s^\prime$ are cubic switching functions which turn off the
# Line 151 | Line 152 | w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
152   while the $w^\prime$ function counters the normal aligned and
153   anti-aligned structures favored by point dipoles:
154   \begin{equation}
155 < w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^0,
155 > w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) = (\cos\theta_{ij}-0.6)^2(\cos\theta_{ij}+0.8)^2-w^\circ,
156   \end{equation}
157   It should be noted that $w$ is proportional to the sum of the $Y_3^2$
158   and $Y_3^{-2}$ spherical harmonics (a linear combination which
# Line 208 | Line 209 | cubic switching function at a cutoff radius.  Under th
209  
210   Long-range dipole-dipole interactions were accounted for in this study
211   by using either the reaction field method or by resorting to a simple
212 < cubic switching function at a cutoff radius.  Under the first method,
213 < the magnitude of the reaction field acting on dipole $i$ is
212 > cubic switching function at a cutoff radius.  The reaction field
213 > method was actually first used in Monte Carlo simulations of liquid
214 > water.\cite{Barker73} Under this method, the magnitude of the reaction
215 > field acting on dipole $i$ is
216   \begin{equation}
217   \mathcal{E}_{i} = \frac{2(\varepsilon_{s} - 1)}{2\varepsilon_{s} + 1}
218   \frac{1}{r_{c}^{3}} \sum_{j\in{\mathcal{R}}} {\bf \mu}_{j} f(r_{ij})\  ,
# Line 232 | Line 235 | at the cutoff radius) and as a result we have two repa
235  
236   We have also performed a companion set of simulations {\it without} a
237   surrounding dielectric (i.e. using a simple cubic switching function
238 < at the cutoff radius) and as a result we have two reparamaterizations
239 < of SSD which could be used either with or without the Reaction Field
238 > at the cutoff radius), and as a result we have two reparamaterizations
239 > of SSD which could be used either with or without the reaction field
240   turned on.
241  
242   Simulations to obtain the preferred density were performed in the
# Line 251 | Line 254 | traditional quaternion integration.\cite{Evans77,Evans
254   symplectic splitting method proposed by Dullweber {\it et
255   al.}\cite{Dullweber1997} Our reason for selecting this integrator
256   centers on poor energy conservation of rigid body dynamics using
257 < traditional quaternion integration.\cite{Evans77,Evans77b} While quaternions
258 < may work well for orientational motion under NVT or NPT integrators,
259 < our limits on energy drift in the microcanonical ensemble were quite
260 < strict, and the drift under quaternions was substantially greater than
261 < in the symplectic splitting method.  This steady drift in the total
259 < energy has also been observed by Kol {\it et al.}\cite{Laird97}
257 > traditional quaternion integration.\cite{Evans77,Evans77b} In typical
258 > microcanonical ensemble simulations, the energy drift when using
259 > quaternions was substantially greater than when using the symplectic
260 > splitting method (fig. \ref{timestep}).  This steady drift in the
261 > total energy has also been observed by Kol {\it et al.}\cite{Laird97}
262  
263   The key difference in the integration method proposed by Dullweber
264   \emph{et al.} is that the entire rotation matrix is propagated from
# Line 446 | Line 448 | results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
448   mean-square displacement as a function of time. The averaged results
449   from five sets of NVE simulations are displayed in figure
450   \ref{diffuse}, alongside experimental, SPC/E, and TIP5P
451 < results.\cite{Gillen72,Mills73,Clancy94,Jorgensen01}
451 > results.\cite{Gillen72,Holz00,Clancy94,Jorgensen01}
452  
453   \begin{figure}
454   \begin{center}
# Line 454 | Line 456 | and Experimental data [Refs. \citen{Gillen72} and \cit
456   \epsfbox{betterDiffuse.epsi}
457   \caption{Average self-diffusion constant as a function of temperature for
458   SSD, SPC/E [Ref. \citen{Clancy94}], TIP5P [Ref. \citen{Jorgensen01}],
459 < and Experimental data [Refs. \citen{Gillen72} and \citen{Mills73}]. Of
459 > and Experimental data [Refs. \citen{Gillen72} and \citen{Holz00}]. Of
460   the three water models shown, SSD has the least deviation from the
461   experimental values. The rapidly increasing diffusion constants for
462   TIP5P and SSD correspond to significant decrease in density at the
# Line 591 | Line 593 | The parameters available for tuning include the $\sigm
593   important properties. In this case, it would be ideal to correct the
594   densities while maintaining the accurate transport behavior.
595  
596 < The parameters available for tuning include the $\sigma$ and $\epsilon$
597 < Lennard-Jones parameters, the dipole strength ($\mu$), and the sticky
598 < attractive and dipole repulsive terms with their respective
599 < cutoffs. To alter the attractive and repulsive terms of the sticky
600 < potential independently, it is necessary to separate the terms as
601 < follows:
602 < \begin{equation}
601 < u_{ij}^{sp}
602 < ({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j) =
603 < \frac{\nu_0}{2}[s(r_{ij})w({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)] + \frac{\nu_0^\prime}{2} [s^\prime(r_{ij})w^\prime({\bf r}_{ij},{\bf \Omega}_i,{\bf \Omega}_j)],
604 < \end{equation}
605 < where $\nu_0$ scales the strength of the tetrahedral attraction and
606 < $\nu_0^\prime$ scales the dipole repulsion term independently. The
607 < separation was performed for purposes of the reparameterization, but
608 < the final parameters were adjusted so that it is not necessary to
609 < separate the terms when implementing the adjusted water
610 < potentials. The results of the reparameterizations are shown in table
611 < \ref{params}. Note that the tetrahedral attractive and dipolar
612 < repulsive terms do not share the same lower cutoff ($r_l$) in the
613 < newly parameterized potentials.  We are calling these
614 < reparameterizations the Soft Sticky Dipole / Reaction Field
596 > The parameters available for tuning include the $\sigma$ and
597 > $\epsilon$ Lennard-Jones parameters, the dipole strength ($\mu$), the
598 > strength of the sticky potential ($\nu_0$), and the sticky attractive
599 > and dipole repulsive cubic switching function cutoffs ($r_l$, $r_u$
600 > and $r_l^\prime$, $r_u^\prime$ respectively). The results of the
601 > reparameterizations are shown in table \ref{params}. We are calling
602 > these reparameterizations the Soft Sticky Dipole / Reaction Field
603   (SSD/RF - for use with a reaction field) and Soft Sticky Dipole
604 < Enhanced (SSD/E - an attempt to improve the liquid structure in
604 > Extended (SSD/E - an attempt to improve the liquid structure in
605   simulations without a long-range correction).
606  
607   \begin{table}
# Line 628 | Line 616 | simulations without a long-range correction).
616   \ \ \ $\epsilon$ (kcal/mol) & 0.152 & 0.152 & 0.152 & 0.152\\
617   \ \ \ $\mu$ (D) & 2.35 & 2.35 & 2.42 & 2.48\\
618   \ \ \ $\nu_0$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
619 + \ \ \ $\omega^\circ$ & 0.07715 & 0.07715 & 0.07715 & 0.07715\\
620   \ \ \ $r_l$ (\AA) & 2.75 & 2.75 & 2.40 & 2.40\\
621   \ \ \ $r_u$ (\AA) & 3.35 & 3.35 & 3.80 & 3.80\\
633 \ \ \ $\nu_0^\prime$ (kcal/mol) & 3.7284 & 3.6613 & 3.90 & 3.90\\
622   \ \ \ $r_l^\prime$ (\AA) & 2.75 & 2.75 & 2.75 & 2.75\\
623   \ \ \ $r_u^\prime$ (\AA) & 4.00 & 4.00 & 3.35 & 3.35\\
624   \end{tabular}
# Line 806 | Line 794 | K, shown by SSD and SSD1 respectively.
794   \begin{center}
795   \epsfxsize=6in
796   \epsfbox{ssdeDiffuse.epsi}
797 < \caption{Plots of the diffusion constants calculated from SSD/E and SSD1,
798 < both without a reaction field, along with experimental results
799 < [Refs. \citen{Gillen72} and \citen{Mills73}]. The NVE calculations were
800 < performed at the average densities observed in the 1 atm NPT
801 < simulations for the respective models. SSD/E is slightly more fluid
802 < than experiment at all of the temperatures, but it is closer than SSD1
803 < without a long-range correction.}
797 > \caption{The diffusion constants calculated from SSD/E and SSD1,
798 > both without a reaction field, along with experimental results
799 > [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
800 > were performed at the average densities observed in the 1 atm NPT
801 > simulations for the respective models. SSD/E is slightly more mobile
802 > than experiment at all of the temperatures, but it is closer to
803 > experiment at biologically relavent temperatures than SSD1 without a
804 > long-range correction.}
805   \label{ssdediffuse}
806   \end{center}
807   \end{figure}
# Line 823 | Line 812 | without an active reaction field, both at the densitie
812   the densities, it is important that the excellent diffusive behavior
813   of SSD be maintained or improved. Figure \ref{ssdediffuse} compares
814   the temperature dependence of the diffusion constant of SSD/E to SSD1
815 < without an active reaction field, both at the densities calculated at
816 < 1 atm and at the experimentally calculated densities for super-cooled
817 < and liquid water. The diffusion constant for SSD/E is consistently
818 < higher than experiment, while SSD1 remains lower than experiment until
819 < relatively high temperatures (greater than 330 K). Both models follow
820 < the shape of the experimental curve well below 300 K but tend to
821 < diffuse too rapidly at higher temperatures, something that is
822 < especially apparent with SSD1.  This increasing diffusion relative to
823 < the experimental values is caused by the rapidly decreasing system
824 < density with increasing temperature.  The densities of SSD1 decay more
825 < rapidly with temperature than do those of SSD/E, leading to more
826 < visible deviation from the experimental diffusion trend.  Thus, the
827 < changes made to improve the liquid structure may have had an adverse
828 < affect on the density maximum, but they improve the transport behavior
829 < of SSD/E relative to SSD1.
815 > without an active reaction field at the densities calculated from the
816 > NPT simulations at 1 atm. The diffusion constant for SSD/E is
817 > consistently higher than experiment, while SSD1 remains lower than
818 > experiment until relatively high temperatures (around 360 K). Both
819 > models follow the shape of the experimental curve well below 300 K but
820 > tend to diffuse too rapidly at higher temperatures, as seen in SSD1's
821 > crossing above 360 K.  This increasing diffusion relative to the
822 > experimental values is caused by the rapidly decreasing system density
823 > with increasing temperature.  Both SSD1 and SSD/E show this deviation
824 > in diffusive behavior, but this trend has different implications on
825 > the diffusive behavior of the models.  While SSD1 shows more
826 > experimentally accurate diffusive behavior in the high temperature
827 > regimes, SSD/E shows more accurate behavior in the supercooled and
828 > biologically relavent temperature ranges.  Thus, the changes made to
829 > improve the liquid structure may have had an adverse affect on the
830 > density maximum, but they improve the transport behavior of SSD/E
831 > relative to SSD1 under the most commonly simulated conditions.
832  
833   \begin{figure}
834   \begin{center}
835   \epsfxsize=6in
836   \epsfbox{ssdrfDiffuse.epsi}
837 < \caption{Plots of the diffusion constants calculated from SSD/RF and SSD1,
837 > \caption{The diffusion constants calculated from SSD/RF and SSD1,
838   both with an active reaction field, along with experimental results
839 < [Refs. \citen{Gillen72} and \citen{Mills73}]. The NVE calculations
839 > [Refs. \citen{Gillen72} and \citen{Holz00}]. The NVE calculations
840   were performed at the average densities observed in the 1 atm NPT
841   simulations for both of the models. Note how accurately SSD/RF
842   simulates the diffusion of water throughout this temperature
843   range. The more rapidly increasing diffusion constants at high
844 < temperatures for both models is attributed to the significantly lower
845 < densities than observed in experiment.}
844 > temperatures for both models is attributed to lower calculated
845 > densities than those observed in experiment.}
846   \label{ssdrfdiffuse}
847   \end{center}
848   \end{figure}
# Line 859 | Line 850 | throughout the temperature range shown and with only a
850   In figure \ref{ssdrfdiffuse}, the diffusion constants for SSD/RF are
851   compared to SSD1 with an active reaction field. Note that SSD/RF
852   tracks the experimental results quantitatively, identical within error
853 < throughout the temperature range shown and with only a slight
854 < increasing trend at higher temperatures. SSD1 tends to diffuse more
855 < slowly at low temperatures and deviates to diffuse too rapidly at
853 > throughout most of the temperature range shown and exhibiting only a
854 > slight increasing trend at higher temperatures. SSD1 tends to diffuse
855 > more slowly at low temperatures and deviates to diffuse too rapidly at
856   temperatures greater than 330 K.  As stated above, this deviation away
857   from the ideal trend is due to a rapid decrease in density at higher
858   temperatures. SSD/RF does not suffer from this problem as much as SSD1
# Line 869 | Line 860 | reparameterization when using an altered long-range co
860   values. These results again emphasize the importance of careful
861   reparameterization when using an altered long-range correction.
862  
863 + \begin{table}
864 + \begin{center}
865 + \caption{Calculated and experimental properties of the single point waters and liquid water at 298 K and 1 atm. (a) Ref. [\citen{Mills73}]. (b) Calculated by integrating the data in ref. \citen{Head-Gordon00_1}. (c) Calculated by integrating the data in ref. \citen{Soper86}. (d) Ref. [\citen{Eisenberg69}]. (e) Calculated for 298 K from data in ref. \citen{Krynicki66}.}
866 + \begin{tabular}{ l  c  c  c  c  c }
867 + \hline \\[-3mm]
868 + \ \ \ \ \ \  & \ \ \ SSD1 \ \ \ & \ SSD/E \ \ \ & \ SSD1 (RF) \ \
869 + \ & \ SSD/RF \ \ \ & \ Expt. \\
870 + \hline \\[-3mm]
871 + \ \ \ $\rho$ (g/cm$^3$) & 0.999 $\pm$0.001 & 0.996 $\pm$0.001 & 0.972 $\pm$0.002 & 0.997 $\pm$0.001 & 0.997 \\
872 + \ \ \ $C_p$ (cal/mol K) & 28.80 $\pm$0.11 & 25.45 $\pm$0.09 & 28.28 $\pm$0.06 & 23.83 $\pm$0.16 & 17.98 \\
873 + \ \ \ $D$ ($10^{-5}$ cm$^2$/s) & 1.78 $\pm$0.07 & 2.51 $\pm$0.18 & 2.00 $\pm$0.17 & 2.32 $\pm$0.06 & 2.299$^\text{a}$ \\
874 + \ \ \ Coordination Number & 3.9 & 4.3 & 3.8 & 4.4 & 4.7$^\text{b}$ \\
875 + \ \ \ H-bonds per particle & 3.7 & 3.6 & 3.7 & 3.7 & 3.4$^\text{c}$ \\
876 + \ \ \ $\tau_1^\mu$ (ps) & 10.9 $\pm$0.6 & 7.3 $\pm$0.4 & 7.5 $\pm$0.7 & 7.2 $\pm$0.4 & 4.76$^\text{d}$ \\
877 + \ \ \ $\tau_2^\mu$ (ps) & 4.7 $\pm$0.4 & 3.1 $\pm$0.2 & 3.5 $\pm$0.3 & 3.2 $\pm$0.2 & 2.3$^\text{e}$ \\
878 + \end{tabular}
879 + \label{liquidproperties}
880 + \end{center}
881 + \end{table}
882 +
883 + Table \ref{liquidproperties} gives a synopsis of the liquid state
884 + properties of the water models compared in this study along with the
885 + experimental values for liquid water at ambient conditions. The
886 + coordination number and hydrogen bonds per particle were calculated by
887 + integrating the following relation:
888 + \begin{equation}
889 + 4\pi\rho\int_{0}^{a}r^2\text{g}(r)dr,
890 + \end{equation}
891 + where $\rho$ is the number density of pair interactions, $a$ is the
892 + radial location of the minima following the first solvation shell
893 + peak, and g$(r)$ is either g$_\text{OO}(r)$ or g$_\text{OH}(r)$ for
894 + calculation of the coordination number or hydrogen bonds per particle
895 + respectively. The number of hydrogen bonds stays relatively constant
896 + across all of the models, but the coordination numbers of SSD/E and
897 + SSD/RF show an improvement over SSD1. This improvement is primarily
898 + due to the widening of the first solvation shell peak, allowing the
899 + first minima to push outward. Comparing the coordination number with
900 + the number of hydrogen bonds can lead to more insight into the
901 + structural character of the liquid.  Because of the near identical
902 + values for SSD1, it appears to be a little too exclusive, in that all
903 + molecules in the first solvation shell are involved in forming ideal
904 + hydrogen bonds.  The differing numbers for the newly parameterized
905 + models indicate the allowance of more fluid configurations in addition
906 + to the formation of an acceptable number of ideal hydrogen bonds.
907 +
908 + The time constants for the self orientational autocorrelation function
909 + are also displayed in Table \ref{liquidproperties}. The dipolar
910 + orientational time correlation function ($\Gamma_{l}$) is described
911 + by:
912 + \begin{equation}
913 + \Gamma_{l}(t) = \langle P_l[\mathbf{u}_j(0)\cdot\mathbf{u}_j(t)]\rangle,
914 + \end{equation}
915 + where $P_l$ is a Legendre polynomial of order $l$ and $\mathbf{u}_j$
916 + is the unit vector of the particle dipole.\cite{Rahman71} From these
917 + correlation functions, the orientational relaxation time of the dipole
918 + vector can be calculated from an exponential fit in the long-time
919 + regime ($t > \tau_l^\mu$).\cite{Rothschild84} Calculation of these
920 + time constants were averaged from five detailed NVE simulations
921 + performed at the STP density for each of the respective models. It
922 + should be noted that the commonly cited value for $\tau_2$ of 1.9 ps
923 + was determined from the NMR data in reference \citen{Krynicki66} at a
924 + temperature near 34$^\circ$C.\cite{Rahman73} Because of the strong
925 + temperature dependence of $\tau_2$, it is necessary to recalculate it
926 + at 298 K to make proper comparisons. The value shown in Table
927 + \ref{liquidproperties} was calculated from the same NMR data in the
928 + fashion described in reference \citen{Krynicki66}. Again, SSD/E and
929 + SSD/RF show improved behavior over SSD1, both with and without an
930 + active reaction field. Turning on the reaction field leads to much
931 + improved time constants for SSD1; however, these results also include
932 + a corresponding decrease in system density. Numbers published from the
933 + original SSD dynamics studies appear closer to the experimental
934 + values, and this difference can be attributed to the use of the Ewald
935 + sum technique versus a reaction field.\cite{Ichiye99}
936 +
937   \subsection{Additional Observations}
938  
939   \begin{figure}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines