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

Comparing trunk/fennellDissertation/dissertation.tex (file contents):
Revision 2957 by chrisfen, Fri Jul 21 22:18:24 2006 UTC vs.
Revision 2971 by chrisfen, Sun Aug 6 22:29:27 2006 UTC

# Line 1 | Line 1
1   \documentclass[11pt]{ndthesis}
2  
3   % some packages for things like equations and graphics
4 + \usepackage[tbtags]{amsmath}
5   \usepackage{amsmath,bm}
6   \usepackage{amssymb}
7   \usepackage{mathrsfs}
# Line 8 | Line 9
9   \usepackage{graphicx}
10   \usepackage{booktabs}
11   \usepackage{cite}
12 + \usepackage{enumitem}
13  
14   \begin{document}
15  
# Line 293 | Line 295 | tricks:
295   al.} are constructed using two different (and separable) computational
296   tricks:
297  
298 < \begin{enumerate}
298 > \begin{enumerate}[itemsep=0pt]
299   \item shifting through the use of image charges, and
300   \item damping the electrostatic interaction.
301   \end{enumerate}  
302 < Wolf \textit{et al.} treated the
303 < development of their summation method as a progressive application of
304 < these techniques,\cite{Wolf99} while Zahn \textit{et al.} founded
305 < their damped Coulomb modification (Eq. (\ref{eq:ZahnPot})) on the
306 < post-limit forces (Eq. (\ref{eq:WolfForces})) which were derived using
307 < both techniques.  It is possible, however, to separate these
308 < tricks and study their effects independently.
302 > Wolf \textit{et al.} treated the development of their summation method
303 > as a progressive application of these techniques,\cite{Wolf99} while
304 > Zahn \textit{et al.} founded their damped Coulomb modification
305 > (Eq. (\ref{eq:ZahnPot})) on the post-limit forces
306 > (Eq. (\ref{eq:WolfForces})) which were derived using both techniques.
307 > It is possible, however, to separate these tricks and study their
308 > effects independently.
309  
310   Starting with the original observation that the effective range of the
311   electrostatic interaction in condensed phases is considerably less
# Line 644 | Line 646 | crystals), so the systems studied were:
646   simulations (i.e. from liquids of neutral molecules to ionic
647   crystals), so the systems studied were:
648  
649 < \begin{enumerate}
649 > \begin{enumerate}[itemsep=0pt]
650   \item liquid water (SPC/E),\cite{Berendsen87}
651   \item crystalline water (Ice I$_\textrm{c}$ crystals of SPC/E),
652   \item NaCl crystals,
# Line 689 | Line 691 | from the reference method ({\sc spme}):
691   We compared the following alternative summation methods with results
692   from the reference method ({\sc spme}):
693  
694 < \begin{enumerate}
694 > \begin{enumerate}[itemsep=0pt]
695   \item {\sc sp} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
696   and 0.3\AA$^{-1}$,
697   \item {\sc sf} with damping parameters ($\alpha$) of 0.0, 0.1, 0.2,
# Line 933 | Line 935 | THE REFERENCE EWALD SUMMATION}
935  
936   \footnotesize
937   \begin{center}
938 < \begin{tabular}{@{} ccrrrrrrrr @{}} \\
938 > \begin{tabular}{@{} ccrrrrrrrr @{}}
939   \toprule
940   \toprule
939
941   & &\multicolumn{4}{c}{Shifted Potential} & \multicolumn{4}{c}{Shifted
942   Force} \\
943   \cmidrule(lr){3-6} \cmidrule(l){7-10} $R_\textrm{c}$ & Groups &
# Line 1009 | Line 1010 | abbreviations are as follows:
1010   investigated.  In all of the individual results table, the method
1011   abbreviations are as follows:
1012  
1013 < \begin{itemize}
1013 > \begin{itemize}[itemsep=0pt]
1014   \item PC = Pure Cutoff,
1015   \item SP = Shifted Potential,
1016   \item SF = Shifted Force,
# Line 1027 | Line 1028 | AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1028  
1029   \footnotesize
1030   \begin{tabular}{@{} ccrrrrrr @{}}
1030 \\
1031   \toprule
1032   \toprule
1033   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1084 | Line 1084 | SYSTEM}
1084  
1085   \footnotesize
1086   \begin{tabular}{@{} ccrrrrrr @{}}
1087 \\
1087   \toprule
1088   \toprule
1089   & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
# Line 1181 | Line 1180 | middle}) AND TORQUE VECTOR MAGNITUDES ({\it lower})}
1180  
1181   \footnotesize
1182   \begin{tabular}{@{} ccrrrrrr @{}}
1184 \\
1183   \toprule
1184   \toprule
1185   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1237 | Line 1235 | OF THE FORCE AND TORQUE VECTORS IN THE ICE I$_\textrm{
1235  
1236   \footnotesize
1237   \begin{tabular}{@{} ccrrrrrr @{}}
1240 \\
1238   \toprule
1239   \toprule
1240   & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque
# Line 1307 | Line 1304 | lower})}
1304  
1305   \footnotesize
1306   \begin{tabular}{@{} ccrrrrrr @{}}
1310 \\
1307   \toprule
1308   \toprule
1309   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1347 | Line 1343 | OF THE FORCE VECTORS IN THE MOLTEN SODIUM CHLORIDE SYS
1343  
1344   \footnotesize
1345   \begin{tabular}{@{} ccrrrrrr @{}}
1350 \\
1346   \toprule
1347   \toprule
1348   & & \multicolumn{3}{c}{Force $\sigma^2$} \\
# Line 1399 | Line 1394 | SYSTEM FOR $\Delta E$ VALUES ({\it upper}) AND FORCE V
1394  
1395   \footnotesize
1396   \begin{tabular}{@{} ccrrrrrr @{}}
1402 \\
1397   \toprule
1398   \toprule
1399   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1440 | Line 1434 | SYSTEM}
1434  
1435   \footnotesize
1436   \begin{tabular}{@{} ccrrrrrr @{}}
1443 \\
1437   \toprule
1438   \toprule
1439   & & \multicolumn{3}{c}{Force $\sigma^2$} \\
# Line 1499 | Line 1492 | SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECT
1492  
1493   \footnotesize
1494   \begin{tabular}{@{} ccrrrrrr @{}}
1502 \\
1495   \toprule
1496   \toprule
1497   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1556 | Line 1548 | CHLORIDE SOLUTION SYSTEM}
1548  
1549   \footnotesize
1550   \begin{tabular}{@{} ccrrrrrr @{}}
1559 \\
1551   \toprule
1552   \toprule
1553   & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
# Line 1620 | Line 1611 | SYSTEM FOR $\Delta E$ VALUES ({\it upper}), FORCE VECT
1611  
1612   \footnotesize
1613   \begin{tabular}{@{} ccrrrrrr @{}}
1623 \\
1614   \toprule
1615   \toprule
1616   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1677 | Line 1667 | SYSTEM}
1667  
1668   \footnotesize
1669   \begin{tabular}{@{} ccrrrrrr @{}}
1680 \\
1670   \toprule
1671   \toprule
1672   & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
# Line 1742 | Line 1731 | MAGNITUDES ({\it middle}) AND TORQUE VECTOR MAGNITUDES
1731  
1732   \footnotesize
1733   \begin{tabular}{@{} ccrrrrrr @{}}
1745 \\
1734   \toprule
1735   \toprule
1736   & & \multicolumn{2}{c}{9\AA} & \multicolumn{2}{c}{12\AA} & \multicolumn{2}{c}{15\AA}\\
# Line 1799 | Line 1787 | ARGON IN LIQUID WATER SYSTEM}  
1787  
1788   \footnotesize
1789   \begin{tabular}{@{} ccrrrrrr @{}}
1802 \\
1790   \toprule
1791   \toprule
1792   & & \multicolumn{3}{c}{Force $\sigma^2$} & \multicolumn{3}{c}{Torque $\sigma^2$} \\
# Line 2010 | Line 1997 | kinetic energy and the virial. More specifically, the
1997   internal pressure of the system in the $NPT$ ensemble, and the
1998   calculation of the pressure includes a components from both the
1999   kinetic energy and the virial. More specifically, the instantaneous
2000 < molecular pressure ($P(t)$) is given by
2000 > molecular pressure ($p(t)$) is given by
2001   \begin{equation}
2002 < P(t) =  \frac{1}{\textrm{d}V}\sum_\mu
2002 > p(t) =  \frac{1}{\textrm{d}V}\sum_\mu
2003          \left[\frac{\mathbf{P}_{\mu}^2}{M_{\mu}}
2004          + \mathbf{R}_{\mu}\cdot\sum_i\mathbf{F}_{\mu i}\right],
2005   \label{eq:MolecularPressure}
# Line 2112 | Line 2099 | S_\textrm{OO}(k) = 1 + 4\pi\rho
2099          \int_0^\infty r^2\frac{\sin kr}{kr}g_\textrm{OO}(r)\textrm{d}r.
2100   \label{eq:SOOofK}
2101   \end{equation}
2102 < Thus, $S_\textrm{OO}(k)$ is related to the Fourier transform of
2103 < $g_\textrm{OO}(r)$.
2102 > Thus, $S_\textrm{OO}(k)$ is related to the Fourier-Laplace transform
2103 > of $g_\textrm{OO}(r)$.
2104  
2105   The expermentally determined $g_\textrm{OO}(r)$ for liquid water has
2106   been compared in great detail with the various common water models,
# Line 2143 | Line 2130 | electrostatic correction.
2130   that the $g_\textrm{OO}(r)$ is insensitive to the choice of
2131   electrostatic correction.
2132  
2133 < \subsection{Thermodynamic Properties}
2133 > \subsection{Thermodynamic Properties}\label{sec:t5peThermo}
2134  
2135 < \subsection{Dynamic Properties}
2135 > In addition to the density, there are a variety of thermodynamic
2136 > quantities that can be calculated for water and compared directly to
2137 > experimental values. Some of these additional quatities include the
2138 > latent heat of vaporization ($\Delta H_\textrm{vap}$), the constant
2139 > pressure heat capacity ($C_p$), the isothermal compressibility
2140 > ($\kappa_T$), the thermal expansivity ($\alpha_p$), and the static
2141 > dielectric constant ($\epsilon$). All of these properties were
2142 > calculated for TIP5P-E with the Ewald summation, so they provide a
2143 > good set for comparisons involving the {\sc sf} technique.
2144  
2145 + The $\Delta H_\textrm{vap}$ is the enthalpy change required to
2146 + transform one mol of substance from the liquid phase to the gas
2147 + phase.\cite{Berry00} In molecular simulations, this quantity can be
2148 + determined via
2149 + \begin{equation}
2150 + \begin{split}
2151 + \Delta H_\textrm{vap} &= H_\textrm{gas} - H_\textrm{liq.} \\
2152 +                      &= E_\textrm{gas} - E_\textrm{liq.}
2153 +                        + p(V_\textrm{gas} - V_\textrm{liq.}) \\
2154 +                      &\approx -\frac{\langle U_\textrm{liq.}\rangle}{N}+ RT,
2155 + \end{split}
2156 + \label{eq:DeltaHVap}
2157 + \end{equation}
2158 + where $E$ is the total energy, $U$ is the potential energy, $p$ is the
2159 + pressure, $V$ is the volume, $N$ is the number of molecules, $R$ is
2160 + the gas constant, and $T$ is the temperature.\cite{Jorgensen98b} As
2161 + seen in the last line of equation \ref{eq:DeltaHVap}, we can
2162 + approximate $\Delta H_\textrm{vap}$ by using the ideal gas for the gas
2163 + state. This allows us to cancel the kinetic energy terms, leaving only
2164 + the $U_\textrm{liq.}$ term. Additionally, the $pV$ term for the gas is
2165 + several orders of magnitude larger than that of the liquid, so we can
2166 + neglect the liquid $pV$ term.
2167 +
2168 + The remaining thermodynamic properties can all be calculated from
2169 + fluctuations of the enthalpy, volume, and system dipole
2170 + moment.\cite{Allen87} $C_p$ can be calculated from fluctuations in the
2171 + enthalpy in constant pressure simulations via
2172 + \begin{equation}
2173 + \begin{split}
2174 + C_p     = \left(\frac{\partial H}{\partial T}\right)_{N,p}
2175 +        = \frac{(\langle H^2\rangle - \langle H\rangle^2)}{Nk_BT^2},
2176 + \end{split}
2177 + \label{eq:Cp}
2178 + \end{equation}
2179 + where $k_B$ is Boltzmann's constant.  $\kappa_T$ can be calculated via
2180 + \begin{equation}
2181 + \begin{split}
2182 + \kappa_T = \frac{1}{V}\left(\frac{\partial V}{\partial p}\right)_{N,T}
2183 +         = \frac{(\langle V^2\rangle_{N,P,T} - \langle V\rangle^2_{N,P,T})}
2184 +                {k_BT\langle V\rangle_{N,P,T}},
2185 + \end{split}
2186 + \label{eq:kappa}
2187 + \end{equation}
2188 + and $\alpha_p$ can be calculated via
2189 + \begin{equation}
2190 + \begin{split}
2191 + \alpha_p        = \frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_{N,P}
2192 +                = \frac{(\langle VH\rangle_{N,P,T}
2193 +                        - \langle V\rangle_{N,P,T}\langle H\rangle_{N,P,T})}
2194 +                        {k_BT^2\langle V\rangle_{N,P,T}}.
2195 + \end{split}
2196 + \label{eq:alpha}
2197 + \end{equation}
2198 + Using the Ewald sum under tin-foil boundary conditions, $\epsilon$ can
2199 + be calculated for systems of non-polarizable substances via
2200 + \begin{equation}
2201 + \epsilon = 1 + \frac{\langle M^2\rangle}{3\epsilon_0k_\textrm{B}TV},
2202 + \label{eq:staticDielectric}
2203 + \end{equation}
2204 + where $\epsilon_0$ is the permittivity of free space and $\langle
2205 + M^2\rangle$ is the fluctuation of the system dipole
2206 + moment.\cite{Allen87} The numerator in the fractional term in equation
2207 + \ref{eq:staticDielectric} is the fluctuation of the simulation-box
2208 + dipole moment, identical to the quantity calculated in the
2209 + finite-system Kirkwood $g$ factor ($G_k$):
2210 + \begin{equation}
2211 + G_k = \frac{\langle M^2\rangle}{N\mu^2},
2212 + \label{eq:KirkwoodFactor}
2213 + \end{equation}
2214 + where $\mu$ is the dipole moment of a single molecule of the
2215 + homogeneous system.\cite{Neumann83,Neumann84,Neumann85} The
2216 + fluctuation term in both equation \ref{eq:staticDielectric} and
2217 + \ref{eq:KirkwoodFactor} is calculated as follows,
2218 + \begin{equation}
2219 + \begin{split}
2220 + \langle M^2\rangle &= \langle\bm{M}\cdot\bm{M}\rangle
2221 +                        - \langle\bm{M}\rangle\cdot\langle\bm{M}\rangle \\
2222 +                   &= \langle M_x^2+M_y^2+M_z^2\rangle
2223 +                        - (\langle M_x\rangle^2 + \langle M_x\rangle^2
2224 +                                + \langle M_x\rangle^2).        
2225 + \end{split}
2226 + \label{eq:fluctBoxDipole}
2227 + \end{equation}
2228 + This fluctuation term can be accumulated during the simulation;
2229 + however, it converges rather slowly, thus requiring multi-nanosecond
2230 + simulation times.\cite{Horn04} In the case of tin-foil boundary
2231 + conditions, the dielectric/surface term of equation \ref{eq:EwaldSum}
2232 + is equal to zero. Since the {\sc sf} method also lacks this
2233 + dielectric/surface term, equation \ref{eq:staticDielectric} is still
2234 + valid for determining static dielectric constants.
2235 +
2236 + All of the above properties were calculated from the same trajectories
2237 + used to determine the densities in section \ref{sec:t5peDensity}
2238 + except for the static dielectric constants. The $\epsilon$ values were
2239 + accumulated from 2ns $NVE$ ensemble trajectories with system densities
2240 + fixed at the average values from the $NPT$ simulations at each of the
2241 + temperatures. The resulting values are displayed in figure
2242 + \ref{fig:t5peThermo}.
2243 + \begin{figure}
2244 + \centering
2245 + \includegraphics[width=5.5in]{./figures/t5peThermo.pdf}
2246 + \caption{Thermodynamic properties for TIP5P-E using the Ewald summation
2247 + and the {\sc sf} techniques along with the experimental values. Units
2248 + for the properties are kcal mol$^{-1}$ for $\Delta H_\textrm{vap}$,
2249 + cal mol$^{-1}$ K$^{-1}$ for $C_p$, 10$^6$ atm$^{-1}$ for $\kappa_T$,
2250 + and 10$^5$ K$^{-1}$ for $\alpha_p$. Ewald summation results are from
2251 + reference \cite{Rick04}. Experimental values for $\Delta
2252 + H_\textrm{vap}$, $\kappa_T$, and $\alpha_p$ are from reference
2253 + \cite{Kell75}. Experimental values for $C_p$ are from reference
2254 + \cite{Wagner02}. Experimental values for $\epsilon$ are from reference
2255 + \cite{Malmberg56}.}
2256 + \label{fig:t5peThermo}
2257 + \end{figure}
2258 +
2259 + As observed for the density in section \ref{sec:t5peDensity}, the
2260 + property trends with temperature seen when using the Ewald summation
2261 + are reproduced with the {\sc sf} technique. Differences include the
2262 + calculated values of $\Delta H_\textrm{vap}$ underpredicting the Ewald
2263 + values. This is to be expected due to the direct weakening of the
2264 + electrostatic interaction through forced neutralization in {\sc
2265 + sf}. This results in an increase of the intermolecular potential
2266 + producing lower values from equation \ref{eq:DeltaHVap}. The slopes of
2267 + these values with temperature are similar to that seen using the Ewald
2268 + summation; however, they are both steeper than the experimental trend,
2269 + indirectly resulting in the inflated $C_p$ values at all temperatures.
2270 +
2271 + Above the supercooled regim\'{e}, $C_p$, $\kappa_T$, and $\alpha_p$
2272 + values all overlap within error. As indicated for the $\Delta
2273 + H_\textrm{vap}$ and $C_p$ results discussed in the previous paragraph,
2274 + the deviations between experiment and simulation in this region are
2275 + not the fault of the electrostatic summation methods but are due to
2276 + the TIP5P class model itself. Like most rigid, non-polarizable,
2277 + point-charge water models, the density decreases with temperature at a
2278 + much faster rate than experiment (see figure
2279 + \ref{fig:t5peDensities}). The reduced density leads to the inflated
2280 + compressibility and expansivity values at higher temperatures seen
2281 + here in figure \ref{fig:t5peThermo}. Incorporation of polarizability
2282 + and many-body effects are required in order for simulation to overcome
2283 + these differences with experiment.\cite{Laasonen93,Donchev06}
2284 +
2285 + At temperatures below the freezing point for experimental water, the
2286 + differences between {\sc sf} and the Ewald summation results are more
2287 + apparent. The larger $C_p$ and lower $\alpha_p$ values in this region
2288 + indicate a more pronounced transition in the supercooled regim\'{e},
2289 + particularly in the case of {\sc sf} without damping. This points to
2290 + the onset of a more frustrated or glassy behavior for TIP5P-E at
2291 + temperatures below 250K in these simulations. Because the systems are
2292 + locked in different regions of phase-space, comparisons between
2293 + properties at these temperatures are not exactly fair. This
2294 + observation is explored in more detail in section
2295 + \ref{sec:t5peDynamics}.
2296 +
2297 + The final thermodynamic property displayed in figure
2298 + \ref{fig:t5peThermo}, $\epsilon$, shows the greatest discrepancy
2299 + between the Ewald summation and the {\sc sf} technique (and experiment
2300 + for that matter). It is known that the dielectric constant is
2301 + dependent upon and quite sensitive to the imposed boundary
2302 + conditions.\cite{Neumann80,Neumann83} This is readily apparent in the
2303 + converged $\epsilon$ values accumulated for the {\sc sf}
2304 + simulations. Lack of a damping function results in dielectric
2305 + constants significantly smaller than that obtained using the Ewald
2306 + sum. Increasing the damping coefficient to 0.2\AA$^{-1}$ improves the
2307 + agreement considerably. It should be noted that the choice of the
2308 + ``Ewald coefficient'' value also has a significant effect on the
2309 + calculated value when using the Ewald summation. In the simulations of
2310 + TIP5P-E with the Ewald sum, this screening parameter was tethered to
2311 + the simulation box size (as was the $R_\textrm{c}$).\cite{Rick04}
2312 + Systems with larger screening parameters reported larger dielectric
2313 + constant values, the same behavior we see here with {\sc sf}. In
2314 + section \ref{sec:dampingDielectric}, this connection is further
2315 + explored as optimal damping coefficients are determined for {\sc
2316 + sf} for capturing the dielectric behavior.
2317 +
2318 + \subsection{Dynamic Properties}\label{sec:t5peDynamics}
2319 +
2320 + To look at the dynamic properties of TIP5P-E when using the {\sc sf}
2321 + method, 200ps $NVE$ simulations were performed for each temperature at
2322 + the average density reported by the $NPT$ simulations. The
2323 + self-diffusion constants ($D$) were calculated with the Einstein
2324 + relation using the mean square displacement (MSD),
2325 + \begin{equation}
2326 + D = \frac{\langle\left|\mathbf{r}_i(t)-\mathbf{r}_i(0)\right|^2\rangle}{6t},
2327 + \label{eq:MSD}
2328 + \end{equation}
2329 + where $t$ is time, and $\mathbf{r}_i$ is the position of particle
2330 + $i$.\cite{Allen87} Figure \ref{fig:ExampleMSD} shows an example MSD
2331 + plot. As labeled in the figure, MSD plots consist of three distinct
2332 + regions:
2333 +
2334 + \begin{enumerate}[itemsep=0pt]
2335 + \item parabolic short-time ballistic motion,
2336 + \item linear diffusive regime, and
2337 + \item poor statistic region at long-time.
2338 + \end{enumerate}
2339 + The slope from the linear region (region 2) is used to calculate $D$.
2340 + \begin{figure}
2341 + \centering
2342 + \includegraphics[width=3.5in]{./figures/ExampleMSD.pdf}
2343 + \caption{Example plot of mean square displacement verses time. The
2344 + left red region is the ballistic motion regime, the middle green
2345 + region is the linear diffusive regime, and the right blue region is
2346 + the region with poor statistics.}
2347 + \label{fig:ExampleMSD}
2348 + \end{figure}
2349 +
2350 + \begin{figure}
2351 + \centering
2352 + \includegraphics[width=3.5in]{./figures/waterFrame.pdf}
2353 + \caption{Body-fixed coordinate frame for a water molecule. The
2354 + respective molecular principle axes point in the direction of the
2355 + labeled frame axes.}
2356 + \label{fig:waterFrame}
2357 + \end{figure}
2358 + In addition to translational diffusion, reorientational time constants
2359 + were calculated for comparisons with the Ewald simulations and with
2360 + experiments. These values were determined from 25ps $NVE$ trajectories
2361 + through calculation of the orientational time correlation function,
2362 + \begin{equation}
2363 + C_l^\alpha(t) = \left\langle P_l\left[\hat{\mathbf{u}}_i^\alpha(t)
2364 +                \cdot\hat{\mathbf{u}}_i^\alpha(0)\right]\right\rangle,
2365 + \label{eq:OrientCorr}
2366 + \end{equation}
2367 + where $P_l$ is the Legendre polynomial of order $l$ and
2368 + $\hat{\mathbf{u}}_i^\alpha$ is the unit vector of molecule $i$ along
2369 + principle axis $\alpha$. The principle axis frame for these water
2370 + molecules is shown in figure \ref{fig:waterFrame}. As an example,
2371 + $C_l^y$ is calculated from the time evolution of the unit vector
2372 + connecting the two hydrogen atoms.
2373 +
2374 + \begin{figure}
2375 + \centering
2376 + \includegraphics[width=3.5in]{./figures/ExampleOrientCorr.pdf}
2377 + \caption{Example plots of the orientational autocorrelation functions
2378 + for the first and second Legendre polynomials. These curves show the
2379 + time decay of the unit vector along the $y$ principle axis.}
2380 + \label{fig:OrientCorr}
2381 + \end{figure}
2382 + From the orientation autocorrelation functions, we can obtain time
2383 + constants for rotational relaxation. Figure \ref{fig:OrientCorr} shows
2384 + some example plots of orientational autocorrelation functions for the
2385 + first and second Legendre polynomials. The relatively short time
2386 + portions (between 1 and 3ps for water) of these curves can be fit to
2387 + an exponential decay to obtain these constants, and they are directly
2388 + comparable to water orientational relaxation times from nuclear
2389 + magnetic resonance (NMR). The relaxation constant obtained from
2390 + $C_2^y(t)$ is of particular interest because it is about the principle
2391 + axis with the minimum moment of inertia and should thereby dominate
2392 + the orientational relaxation of the molecule.\cite{Impey82} This means
2393 + that $C_2^y(t)$ should provide the best comparison to the NMR
2394 + relaxation data.
2395 +
2396 + \begin{figure}
2397 + \centering
2398 + \includegraphics[width=5.5in]{./figures/t5peDynamics.pdf}
2399 + \caption{Diffusion constants ({\it upper}) and reorientational time
2400 + constants ({\it lower}) for TIP5P-E using the Ewald sum and {\sc sf}
2401 + technique compared with experiment. Data at temperatures less that
2402 + 0$^\circ$C were not included in the $\tau_2^y$ plot to allow for
2403 + easier comparisons in the more relavent temperature regime.}
2404 + \label{fig:t5peDynamics}
2405 + \end{figure}
2406 + Results for the diffusion constants and reorientational time constants
2407 + are shown in figure \ref{fig:t5peDynamics}. From this figure, it is
2408 + apparent that the trends for both $D$ and $\tau_2^y$ of TIP5P-E using
2409 + the Ewald sum are reproduced with the {\sc sf} techinque. The enhanced
2410 + diffusion at high temperatures are again the product of the lower
2411 + densities in comparison with experiment and do not provide any special
2412 + insight into differences between the electrostatic summation
2413 + techniques. With the undamped {\sc sf} technique, TIP5P-E tends to
2414 + diffuse a little faster than with the Ewald sum; however, use of light
2415 + to moderate damping results in indistiguishable $D$ values. Though not
2416 + apparent in this figure, {\sc sf} values at the lowest temperature are
2417 + approximately an order of magnitude lower than with Ewald. These
2418 + values support the observation from section \ref{sec:t5peThermo} that
2419 + there appeared to be a change to a more glassy-like phase with the
2420 + {\sc sf} technique at these lower temperatures.
2421 +
2422 + The $\tau_2^y$ results in the lower frame of figure
2423 + \ref{fig:t5peDynamics} show a much greater difference between the {\sc
2424 + sf} results and the Ewald results. At all temperatures shown, TIP5P-E
2425 + relaxes faster than experiment with the Ewald sum while tracking
2426 + experiment fairly well when using the {\sc sf} technique, independent
2427 + of the choice of damping constant. Their are several possible reasons
2428 + for this deviation between techniques. The Ewald results were taken
2429 + shorter (10ps) trajectories than the {\sc sf} results (25ps). A quick
2430 + calculation from a 10ps trajectory with {\sc sf} with an $\alpha$ of
2431 + 0.2\AA$^-1$ at 25$^\circ$C showed a 0.4ps drop in $\tau_2^y$, placing
2432 + the result more in line with that obtained using the Ewald sum. These
2433 + results support this explanation; however, recomputing the results to
2434 + meet a poorer statistical standard is counter-productive. Assuming the
2435 + Ewald results are not the product of poor statistics, differences in
2436 + techniques to integrate the orientational motion could also play a
2437 + role. {\sc shake} is the most commonly used technique for
2438 + approximating rigid-body orientational motion,\cite{Ryckaert77} where
2439 + as in {\sc oopse}, we maintain and integrate the entire rotation
2440 + matrix using the {\sc dlm} method.\cite{Meineke05} Since {\sc shake}
2441 + is an iterative constraint technique, if the convergence tolerances
2442 + are raised for increased performance, error will accumulate in the
2443 + orientational motion. Finally, the Ewald results were calculated using
2444 + the $NVT$ ensemble, while the $NVE$ ensemble was used for {\sc sf}
2445 + calculations. The additional mode of motion due to the thermostat will
2446 + alter the dynamics, resulting in differences between $NVT$ and $NVE$
2447 + results. These differences are increasingly noticable as the
2448 + thermostat time constant decreases.
2449 +
2450 + \section{Damping of Point Multipoles}\label{sec:dampingMultipoles}
2451 +
2452 +
2453 +
2454 + \section{Damping and Dielectric Constants}\label{sec:dampingDielectric}
2455 +
2456   \section{Conclusions}\label{sec:PairwiseConclusions}
2457  
2458   The above investigation of pairwise electrostatic summation techniques
# Line 2201 | Line 2507 | expect.
2507   required to obtain the level of accuracy most researchers have come to
2508   expect.
2509  
2204
2205
2206
2510   \chapter{\label{chap:water}SIMPLE MODELS FOR WATER}
2511  
2512   \chapter{\label{chap:ice}PHASE BEHAVIOR OF WATER IN COMPUTER SIMULATIONS}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines