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

Comparing trunk/multipole/dielectric_new.tex (file contents):
Revision 4404 by gezelter, Tue Mar 29 15:27:46 2016 UTC vs.
Revision 4408 by gezelter, Wed Mar 30 20:12:46 2016 UTC

# Line 12 | Line 12
12   \usepackage{braket}
13   \usepackage{tabularx}
14   \usepackage{rotating}
15 + \usepackage{multirow}
16 + \usepackage{booktabs}
17  
18   \newcolumntype{Y}{>{\centering\arraybackslash}X}
19  
# Line 165 | Line 167 | Under the influence of weak external fields, the bulk
167   \item potentials of mean force between solvated ions,
168   \end{enumerate}
169  
170 + \begin{figure}
171 + \includegraphics[width=\linewidth]{Schematic}
172 + \caption{Dielectric properties of a fluid measure the response to
173 +  external electric fields and gradients (left panel), or internal
174 +  fields and gradients generated by the molecules themselves (central
175 +  panel), or fields produced by embedded ions (right panel). The
176 +  dielectric constant measures all three responses in dipolar fluids.
177 +  In quadrupolar liquids, the relevant bulk property is the
178 +  quadrupolar susceptibility, and the geometry of the perturbation
179 +  determines the effective dielectric screening.}
180 + \label{fig:schematic}
181 + \end{figure}
182 +
183   Under the influence of weak external fields, the bulk polarization of
184   the system is primarily a linear response to the perturbation, where
185   proportionality constant depends on the electrostatic interactions
# Line 385 | Line 400 | $ \textbf{E} \rightarrow 0 $.
400   $ \textbf{E} \rightarrow 0 $.
401  
402   \subsection{Correction Factors}
403 <
403 > \label{sec:corrFactor}
404   In the presence of a uniform external field $ \mathbf{E}^\circ$, the
405   total electric field at $\mathbf{r}$ depends on the polarization
406   density at all other points in the system,\cite{NeumannI83}
# Line 401 | Line 416 | point dipoles (see Fig. \ref{fig:stockmayer}).
416  
417   In simulations of dipolar fluids, the molecular dipoles may be
418   represented either by closely-spaced point charges or by
419 < point dipoles (see Fig. \ref{fig:stockmayer}).
419 > point dipoles (see Fig. \ref{fig:tensor}).
420   \begin{figure}
421 < \includegraphics[width=3in]{DielectricFigure}
422 < \caption{With the real-space electrostatic methods, the effective
423 <  dipole tensor, $\mathbf{T}$, governing interactions between
424 <  molecular dipoles is not the same for charge-charge interactions as
425 <  for point dipoles.}
426 < \label{fig:stockmayer}
421 > \includegraphics[width=\linewidth]{Tensors}
422 > \caption{In the real-space electrostatic methods, the molecular dipole
423 >  tensor, $\mathbf{T}_{\alpha\beta}(r)$, is not the same for
424 >  charge-charge interactions as for point dipoles (left panel). The
425 >  same holds true for the molecular quadrupole tensor (right panel),
426 >  $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$, which can have distinct
427 >  forms if the molecule is represented by charges, dipoles, or point
428 >  quadrupoles.}
429 > \label{fig:tensor}
430   \end{figure}
431   In the case where point charges are interacting via an electrostatic
432   kernel, $v(r)$, the effective {\it molecular} dipole tensor,
# Line 473 | Line 491 | expression for the dielectric constant,
491   \begin{equation}
492   \epsilon_{CB} = 1 + \frac{\braket{\bf{M}^2}-{\braket{\bf{M}}}^2}{3
493    \epsilon_o V k_B T}
494 < \label{correctionFormula}
494 > \label{conductingBoundary}
495   \end{equation}
496 < Equation (\ref{correctionFormula}) allows estimation of the static
497 < dielectric constant from fluctuations easily computed directly from
498 < simulations.
481 <
496 > Eqs. (\ref{correctionFormula}) and (\ref{conductingBoundary}) allows
497 > estimation of the static dielectric constant from fluctuations easily
498 > computed directly from simulations.
499  
500   % Here $\chi^*_D$ is a dipolar susceptibility can be expressed in terms of dielectric constant as $ \chi^*_D = \epsilon - 1$ which different than macroscopic dipolar polarizability $\alpha_D$ in the sections \ref{subsec:perturbation} and \ref{subsec:fluctuation}. We can split integral into two parts: singular part i.e $|\textbf{r}-\textbf{r}'|\rightarrow 0 $ and non-singular part i.e $|\textbf{r}-\textbf{r}'| > 0 $ . The singular part of the integral can be written as,\cite{NeumannI83, Jackson98}
501   % \begin{equation}
# Line 522 | Line 539 | listed in Table \ref{tab:A}.
539      for comparison using the Ewald convergence parameter ($\kappa$)
540      and the real-space cutoff value ($r_c$). }
541   \label{tab:A}
542 < {%
543 < \begin{tabular}{l|c|c|c}
527 <      
542 > \begin{tabular}{l|c|c}
543 > \toprule      
544   Method & point charges & point dipoles  \\
545 < \hline
545 > \colrule
546   Spherical Cutoff (SC) & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
547   Shifted Potental (SP) & $ \mathrm{erf}(r_c \alpha) - \frac{2 \alpha r_c}{\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $\mathrm{erf}(r_c \alpha) -\frac{2 \alpha r_c}{\sqrt{\pi}}\left(1+\frac{2\alpha^2 {r_c}^2}{3} \right)e^{-\alpha^2{r_c}^2} $\\
548   Gradient-shifted  (GSF) & 1 & $\mathrm{erf}(\alpha  r_c)-\frac{2 \alpha  r_c}{\sqrt{\pi}}  \left(1 + \frac{2 \alpha^2 r_c^2}{3} + \frac{\alpha^4 r_c^4}{3}\right)e^{-\alpha^2 r_c^2} $ \\
549 < Taylor-shifted  (TSF) & 1 & 1 \\
550 < Ewald-Kornfeld (EK) & $\mathrm{erf}(r_c \kappa) - \frac{2 \kappa
551 <                      r_c}{\sqrt{\pi}} e^{-\kappa^2 r_c^2}$ & \\
552 < \end{tabular}%
553 < }
549 > Taylor-shifted  (TSF) & 1 & 1 \\ \colrule
550 > Ewald-Kornfeld (EK) & \multicolumn{2}{c}{$\mathrm{erf}(r_c \kappa) -
551 >                      \frac{2 \kappa r_c}{\sqrt{\pi}} e^{-\kappa^2
552 >                      r_c^2}$}  \\
553 > \botrule
554 > \end{tabular}
555   \end{table}
556   Note that for point charges, the GSF and TSF methods produce estimates
557   of the dielectric that need no correction, and the TSF method likewise
# Line 629 | Line 646 | interest ($\mathbf{r}$).
646   tensor connecting quadrupoles at $\mathbf{r}^\prime$ with the point of
647   interest ($\mathbf{r}$).
648  
649 < \begin{figure}
650 < \includegraphics[width=3in]{QuadrupoleFigure}
651 < \caption{With the real-space electrostatic methods, the molecular
652 <  quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
653 <  governing interactions between molecules is not the same for
654 <  quadrupoles represented via sets of charges, point dipoles, or by
655 <  single point quadrupoles.}
656 < \label{fig:quadrupolarFluid}
657 < \end{figure}
649 > % \begin{figure}
650 > % \includegraphics[width=3in]{QuadrupoleFigure}
651 > % \caption{With the real-space electrostatic methods, the molecular
652 > %   quadrupole tensor, $\mathbf{T}_{\alpha\beta\gamma\delta}(r)$,
653 > %   governing interactions between molecules is not the same for
654 > %   quadrupoles represented via sets of charges, point dipoles, or by
655 > %   single point quadrupoles.}
656 > % \label{fig:quadrupolarFluid}
657 > % \end{figure}
658  
659   In simulations of quadrupolar fluids, the molecular quadrupoles may be
660   represented by closely-spaced point charges, by multiple point
661   dipoles, or by a single point quadrupole (see
662 < Fig. \ref{fig:quadrupolarFluid}).  In the case where point charges are
662 > Fig. \ref{fig:tensor}).  In the case where point charges are
663   interacting via an electrostatic kernel, $v(r)$, the effective
664   molecular quadrupole tensor can obtained from four successive
665   applications of the gradient operator to the electrostatic kernel,
# Line 717 | Line 734 | the field gradient vanishes at the singularity (see Ap
734   The integral in equation (\ref{gradMaxwell}) can be divided into two
735   parts, $|\mathbf{r}-\mathbf{r}^\prime|\rightarrow 0 $ and
736   $|\mathbf{r}-\mathbf{r}^\prime|> 0$. Since the self-contribution to
737 < the field gradient vanishes at the singularity (see Appendix
738 < \ref{singularQuad}), equation (\ref{gradMaxwell}) can be written as,
737 > the field gradient vanishes at the singularity (see the supporting
738 > information), equation (\ref{gradMaxwell}) can be written as,
739   \begin{equation}
740   \partial_\alpha E_\beta(\mathbf{r}) = \partial_\alpha {E}^\circ_\beta(\mathbf{r}) +
741    \frac{1}{4\pi \epsilon_o}\int\limits_{|\mathbf{r}-\mathbf{r}^\prime|> 0 }
# Line 835 | Line 852 | susceptibility,
852   susceptibility,
853   \begin{equation}
854   \chi_Q = \frac{\alpha_Q}{1 + B \alpha_Q}.
855 + \label{eq:quadrupolarSusceptiblity}
856   \end{equation}
857   If an electrostatic method produces $B \rightarrow 0$, the computed
858   quadrupole polarizability and quadrupole susceptibility converge to
# Line 847 | Line 865 | the same value.
865      of the correction factor are $ \mathrm{length}^{-2}$ for quadrupolar
866      fluids.}
867   \label{tab:B}
868 < \begin{tabular}{l|c|c|c|c}
869 <      
870 < Method & charges & dipoles & quadrupoles \\\hline
868 > \begin{tabular}{l|c|c|c}
869 > \toprule      
870 > Method & charges & dipoles & quadrupoles \\\colrule
871   Spherical Cutoff (SC) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ &  $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & $ -\frac{8 {\alpha}^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ \\
872   Shifted Potental (SP) & $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ &  $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$& $ -\frac{16 \alpha^7 {r_c}^5}{9\sqrt{\pi}} e^{-\alpha^2 r_c^2}$  \\
873   Gradient-shifted  (GSF) & $- \frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}} e^{-\alpha^2 r_c^2}$ & 0 &  $-\frac{4{\alpha}^7{r_c}^5 }{9\sqrt{\pi}}e^{-\alpha^2 r_c^2}(-1+2\alpha ^2 r_c^2)$\\
874   Taylor-shifted  (TSF) &  $ -\frac{8 \alpha^5 {r_c}^3}{3\sqrt{\pi}}
875                          e^{-\alpha^2 r_c^2}$ &
876 <                                               $\frac{4\;\mathrm{erfc(\alpha r_c)}}{{r_c}^2} + \frac{8 \alpha}{3\sqrt{\pi}r_c}e^{-\alpha^2 {r_c}^2}\left(3+ 2 \alpha^2 {r_c}^2 + \alpha^4 {r_c}^4\right)  $ & $\frac{10\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}}{9\sqrt{\pi}{r_c}}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)$ \\
877 < Ewald-Kornfeld (EK) & & &
876 >                                               $\frac{4\;\mathrm{erfc(\alpha
877 >                                               r_c)}}{{r_c}^2} +
878 >                                               \frac{8
879 >                                               \alpha}{3\sqrt{\pi}r_c}e^{-\alpha^2
880 >                                               {r_c}^2}\left(3+ 2
881 >                                               \alpha^2 {r_c}^2 +
882 >                                               \alpha^4 {r_c}^4\right)
883 >                                               $ &
884 >                                                   $\frac{10\;\mathrm{erfc}(\alpha r_c )}{{r_c}^2} + \frac{4{\alpha}}{9\sqrt{\pi}{r_c}}e^{-\alpha^2 r_c^2}\left(45 + 30\alpha ^2 {r_c}^2 + 12\alpha^4 {r_c}^4 + 3\alpha^6 {r_c}^6 + 2 \alpha^8 {r_c}^8\right)$ \\
885 > \colrule
886 > Ewald-Kornfeld (EK) & \multicolumn{3}{c}{ Nothing Yet} \\
887 > \botrule
888   \end{tabular}
889   \end{sidewaystable}
890  
# Line 1017 | Line 1045 | perturbations, field strengths ranging from
1045   0.5~ns in the canonical (NVT) ensemble.  Data collection was carried
1046   out over a 1~ns simulation in the microcanonical (NVE) ensemble.  Box
1047   dipole moments were sampled every fs.  For simulations with external
1048 < perturbations, field strengths ranging from
1049 < $0 - 10 \times 10^{-4} \mathrm{~V/\r{A}}$ with increments of
1050 < $ 10^{-4} \mathrm{~V/\r{A}}$ were carried out for each system.
1048 > perturbations, field strengths ranging from $0 - 10 \times
1049 > 10^{-4}$~V/\AA with increments of $ 10^{-4}$~V/\AA were carried out
1050 > for each system.
1051  
1052   Quadrupolar systems contained 4000 linear point quadrupoles with a
1053 < density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K.  These
1053 > density $2.338 \mathrm{~g/cm}^3$ at a temperature of 500~K. These
1054   systems were equilibrated for 200~ps in a canonical (NVT) ensemble.
1055   Data collection was carried out over a 500~ps simulation in the
1056 < microcanonical (NVE) ensemble.  Components of box quadrupole moments
1057 < were sampled every 100 fs.  For quadrupolar simulations with external
1056 > microcanonical (NVE) ensemble. Components of box quadrupole moments
1057 > were sampled every 100 fs. For quadrupolar simulations with external
1058   field gradients, field strengths ranging from
1059 < $0 - 9 \times 10^{-2} \mathrm{~V/\r{A}}^2$ with increments of
1060 < $10^{-2} \mathrm{~V/\r{A}^2}$ were carried out for each system.
1059 > $0 - 9 \times 10^{-2}$~V/\AA$^2$ with increments of
1060 > $10^{-2}$~V/\AA$^2$ were carried out for each system.
1061  
1062   To carry out the PMF simulations, two of the multipolar molecules in
1063   the test system were converted into \ce{q+} and \ce{q-} ions and
1064   constrained to remain at a fixed distance for the duration of the
1065 < simulation. The constrained distance was then varied from
1066 < $5 - 12 \mathrm{~\r{A}}$. In the PMF calculations, all simulations were
1067 < equilibrated for 500 ps in the NVT ensemble and run for 5 ns in the
1068 < microcanonical (NVE) ensemble.  Constraint forces were sampled every
1041 < 20~fs.
1065 > simulation. The constrained distance was then varied from 5--12~\AA.
1066 > In the PMF calculations, all simulations were equilibrated for 500 ps
1067 > in the NVT ensemble and run for 5 ns in the microcanonical (NVE)
1068 > ensemble.  Constraint forces were sampled every 20~fs.
1069        
1070   \section{Results}
1071   \subsection{Dipolar fluid}
# Line 1056 | Line 1083 | shown in the upper panels in Fig.  \ref{fig:dielectric
1083   \label{fig:dielectricDipole}
1084   \end{figure}
1085   The macroscopic polarizability ($\alpha_D$) for the dipolar fluid is
1086 < shown in the upper panels in Fig.  \ref{fig:dielectricDipole}.  The
1086 > shown in the upper panels in Fig. \ref{fig:dielectricDipole}.  The
1087   polarizability obtained from the both perturbation and fluctuation
1088   approaches are in excellent agreement with each other.  The data also
1089   show a stong dependence on the damping parameter for both the Shifted
# Line 1065 | Line 1092 | The calculated correction factors discussed in section
1092   parameter.
1093  
1094   The calculated correction factors discussed in section
1095 < \ref{sec:corrFactor} are shown in the middle panels.  Because the TSF
1095 > \ref{sec:corrFactor} are shown in the middle panels. Because the TSF
1096   method has $A = 1$ for all values of the damping parameter, the
1097   computed polarizabilities need no correction for the dielectric
1098   calculation. The value of $A$ varies with the damping parameter in
1099   both the SP and GSF methods, and inclusion of the correction yields
1100   dielectric estimates (shown in the lower panel) that are generally too
1101 < large until the damping reaches $\sim 0.25 \mathrm{~\r{A}}^{-1}$.
1102 < Above this value, the dielectric constants are generally in good
1103 < agreement with previous simulation results.\cite{NeumannI83}
1101 > large until the damping reaches $\sim$~0.25~\AA$^{-1}$. Above this
1102 > value, the dielectric constants are generally in good agreement with
1103 > previous simulation results.\cite{NeumannI83}
1104  
1105   Figure \ref{fig:dielectricDipole} also contains back-calculations of
1106   the polarizability using the reference (Ewald) simulation
# Line 1085 | Line 1112 | of ($\alpha \sim 0.2-0.3 \mathrm{\r{A}}^{-1}$ when eva
1112   the value of $\mathrm{A}$ deviates significantly from unity.
1113  
1114   These results also suggest an optimal value for the damping parameter
1115 < of ($\alpha \sim 0.2-0.3 \mathrm{\r{A}}^{-1}$ when evaluating
1116 < dielectric constants of point dipolar fluids using either the
1117 < perturbation and fluctuation approaches for the new real-space
1091 < methods.
1115 > of ($\alpha \sim 0.2-0.3$~\AA$^{-1}$ when evaluating dielectric
1116 > constants of point dipolar fluids using either the perturbation and
1117 > fluctuation approaches for the new real-space methods.
1118  
1119   \begin{figure}
1120   \includegraphics[width=4in]{ScreeningFactor_Dipole.pdf}
# Line 1115 | Line 1141 | $\alpha = 0.2 \AA^{-1}$ and $0.3 \AA^{-1}$ and large s
1141   occurs when the ions are separated (or when the damping parameter is
1142   large). In Fig. \ref{fig:ScreeningFactor_Dipole} we observe that for
1143   the higher value of damping alpha \textit{i.e.}
1144 < $\alpha = 0.2 \AA^{-1}$ and $0.3 \AA^{-1}$ and large separation
1144 > $\alpha = 0.2$~\AA$^{-1}$ and $0.3$~\AA$^{-1}$ and large separation
1145   between ions, the screening factor does indeed approach the correct
1146   dielectric constant.
1147  
# Line 1244 | Line 1270 | parameters in the range 0.25 - 0.3$\AA^{-1}$.
1270   The dielectric constant evaluated using the computed polarizability
1271   and correction factors agrees well with the previous Ewald-based
1272   simulation results \cite{Adams81,NeumannI83} for moderate damping
1273 < parameters in the range 0.25 - 0.3$\AA^{-1}$.
1273 > parameters in the range 0.25--0.3~\AA~$^{-1}$.
1274  
1275   Although the TSF method alters many dynamic and structural features in
1276   multipolar liquids,\cite{PaperII} it is surprisingly good at computing

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines