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

Comparing trunk/nonperiodicVSS/nonperiodicVSS.tex (file contents):
Revision 3932 by kstocke1, Tue Aug 6 21:44:34 2013 UTC vs.
Revision 3947 by kstocke1, Fri Sep 6 13:09:47 2013 UTC

# Line 2 | Line 2
2   \setkeys{acs}{usetitle = true}
3  
4   \usepackage{caption}
5 < \usepackage{float}
5 > \usepackage{endfloat}
6   \usepackage{geometry}
7   \usepackage{natbib}
8   \usepackage{setspace}
# Line 86 | Line 86 | We have adapted the Velocity Shearing and Scaling Reve
86   \section{Methodology}
87  
88   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89 % FORCE FIELD PARAMETERS
90 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91 \subsection{Force field parameters}
92
93 We have chosen the SPC/E water model for these simulations. There are many values for physical properties from previous simulations available for direct comparison.
94
95 Gold-gold interactions are described by the quantum Sutton-Chen (QSC) model. The QSC parameters are tuned to experimental properties such as density, cohesive energy, and elastic moduli and include zero-point quantum corrections.
96
97 Hexane molecules are described by the TraPPE united atom model. This model provides good computational efficiency and reasonable accuracy for bulk thermal conductivity values.
98
99 Metal-nonmetal interactions are governed by parameters derived from Luedtke and Landman.
100
101
102 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89   % NON-PERIODIC DYNAMICS
90   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91   \subsection{Dynamics for non-periodic systems}
92  
93 < We have run all tests using the Langevin Hull nonperiodic simulation methodology. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are subject to standard Newtonian dynamics. Thermal coupling to the bath was turned off to avoid interference with any imposed kinetic flux. Systems containing liquids were run under moderate pressure ($\sim$ 5 atm) to avoid the formation of a substantial vapor phase, which would have a hull created out of a relatively small number of molecules.
93 > We have run all tests using the Langevin Hull nonperiodic simulation methodology.\cite{Vardeman2011} The Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different compressibilities, which are typically problematic for traditional affine transform methods. We have had success applying this method to
94 > several different systems including bare metal nanoparticles, liquid water clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are subject to standard Newtonian dynamics. For these tests, thermal coupling to the bath was turned off to avoid interference with any imposed kinetic flux. Systems containing liquids were run under moderate pressure ($\sim$ 5 atm) to avoid the formation of a substantial vapor phase.
95  
96   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97   % NON-PERIODIC RNEMD
98   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
99   \subsection{VSS-RNEMD for non-periodic systems}
100  
101 < The adaptation of VSS-RNEMD for non-periodic systems is relatively
101 > The most useful RNEMD approach developed so far utilizes a series of
102 > simultaneous velocity shearing and scaling (VSS) exchanges between the two
103 > regions.\cite{Kuang2012} This method provides a set of conservation constraints
104 > while simultaneously creating a desired flux between the two regions. Satisfying
105 > the constraint equations ensures that the new configurations are sampled from the
106 > same NVE ensemble.
107 >
108 > We have implemented this new method in OpenMD, our group molecular dynamics code.\cite{openmd} The adaptation of VSS-RNEMD for non-periodic systems is relatively
109   straightforward. The major modifications to the method are the addition of a rotational shearing term and the use of more versatile hot / cold regions instead
110   of rectangular slabs. A temperature profile along the $r$ coordinate is created by recording the average temperature of concentric spherical shells.
111  
112 + \begin{figure}
113 +        \center{\includegraphics[width=7in]{figures/VSS}}
114 +        \caption{Schematics of periodic (left) and nonperiodic (right) Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum flux is applied from region B to region A. Thermal gradients are depicted by a color gradient. Linear or angular velocity gradients are shown as arrows.}
115 +        \label{fig:VSS}
116 + \end{figure}
117 +
118   At each time interval, the particle velocities ($\mathbf{v}_i$ and
119   $\mathbf{v}_j$) in the cold and hot shells ($C$ and $H$) are modified by a
120   velocity scaling coefficient ($c$ and $h$), an additive linear velocity shearing
# Line 122 | Line 122 | average linear and angular velocities for each shell.
122   shearing term ($\mathbf{b}_c$ and $\mathbf{b}_h$). Here, $\langle
123   \mathbf{v}_{i,j} \rangle$ and $\langle \mathbf{\omega}_{i,j} \rangle$ are the
124   average linear and angular velocities for each shell.
125
125   \begin{displaymath}
126   \begin{array}{rclcl}
127   & \underline{~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~} & &
# Line 142 | Line 141 | The total energy is constrained via two quadratic form
141   \end{eqnarray}
142  
143   The total energy is constrained via two quadratic formulae,
145
144   \begin{eqnarray}
145   K_c - J_r \; \Delta \, t & = & c^2 \, (K_c - \frac{1}{2}\langle \omega_c \rangle \cdot I_c\langle \omega_c \rangle) + \frac{1}{2} \mathbf{b}_c \cdot I_c \, \mathbf{b}_c \label{eq:Kc}\\
146   K_h + J_r \; \Delta \, t & = & h^2 \, (K_h - \frac{1}{2}\langle \omega_h \rangle \cdot I_h\langle \omega_h \rangle) + \frac{1}{2} \mathbf{b}_h \cdot I_h \, \mathbf{b}_h \label{eq:Kh}
# Line 155 | Line 153 | the velocity scaling ($c$ and $h$) and shearing ($\mat
153   the velocity scaling ($c$ and $h$) and shearing ($\mathbf{b}_c,\,$ and $\mathbf{b}_h$) at each time interval.
154  
155   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156 < % **TESTS AND APPLICATIONS**
156 > % **COMPUTATIONAL DETAILS**
157   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158 < \section{Tests and Applications}
158 > \section{Computational Details}
159  
160   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161 < % THERMAL CONDUCTIVITIES
161 > % SIMULATION PROTOCOL
162   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
163 < \subsection{Thermal conductivities}
163 > \subsection{Simulation protocol}
164  
167 Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at different kinetic energy flux values are shown in Table \ref{table:goldconductivity}. For these calculations, the hot and cold slabs were excluded from the linear regression of the thermal gradient. The measured linear slope $\langle dT / dr \rangle$ is linearly dependent on the applied kinetic energy flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 - 1.26 W m$^{-1}$ K$^{-1}$ [cite NIVS paper], though still significantly lower than the experimental value of 320 W m$^{-1}$ K$^{-1}$, as the QSC force field neglects significant electronic contributions to heat conduction. The small increase relative to previous simulated bulk values is due to a slight increase in gold density -- as expected, an increase in density results in higher thermal conductivity values. The increased density is a result of nanoparticle curvature relative to an infinite bulk slab, which introduces surface tension that increases ambient density.
165  
166 < \begin{longtable}{p{2.7cm} p{2.5cm} p{2.5cm}}
167 <        \caption{Calculated thermal conductivity of a crystalline gold nanoparticle of radius 40 \AA. Calculations were performed at 300 K and ambient density. Gold-gold interactions are described by the Quantum Sutton-Chen potential.}
168 <        \\ \hline \hline
169 <                \centering {$J_r$} & \centering\arraybackslash {$\langle dT / dr \rangle$} & \centering\arraybackslash {$\boldsymbol \lambda$}\\
173 <        \centering {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & \centering\arraybackslash {\small(K \AA$^{-1}$)} & \centering\arraybackslash {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
174 < \endhead
175 < \hline
176 < % \endfoot
177 < \centering3.25$\times 10^{-6}$ & \centering\arraybackslash 0.11435 & \centering\arraybackslash 1.9753 \\
178 < \centering6.50$\times 10^{-6}$ & \centering\arraybackslash 0.2324 & \centering\arraybackslash 1.9438 \\
179 < \centering1.30$\times 10^{-5}$ & \centering\arraybackslash 0.44922 & \centering\arraybackslash 2.0113 \\
180 < \centering3.25$\times 10^{-5}$ & \centering\arraybackslash 1.1802 & \centering\arraybackslash 1.9139 \\
181 < \centering6.50$\times 10^{-5}$ & \centering\arraybackslash 2.339 & \centering\arraybackslash 1.9314
182 < \\ \hline \hline
183 < \label{table:goldconductivity}
184 < \end{longtable}
185 <        
186 < SPC/E Water Cluster:
166 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
167 > % FORCE FIELD PARAMETERS
168 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169 > \subsection{Force field parameters}
170  
171 < \begin{longtable}{p{2.7cm} p{2.5cm} p{2.5cm}}
189 <        \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and ambient density.}
190 <        \\ \hline \hline
191 <                \centering {$J_r$} & \centering\arraybackslash {$\langle dT / dr \rangle$} & \centering\arraybackslash {$\boldsymbol \lambda$}\\
192 <        \centering {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & \centering\arraybackslash {\small(K \AA$^{-1}$)} & \centering\arraybackslash {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
193 < \endhead
194 < \hline
195 < % \endfoot
171 > We have chosen the SPC/E water model for these simulations, which is particularly useful for method validation as there are many values for physical properties from previous simulations available for direct comparison.\cite{Bedrov:2000, Kuang2010}
172  
173 < \\ \hline \hline
198 < \label{table:waterconductivity}
199 < \end{longtable}
173 > Gold -- gold interactions are described by the quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} The QSC parameters are tuned to experimental properties such as density, cohesive energy, and elastic moduli and include zero-point quantum corrections.
174  
175 + Hexane molecules are described by the TraPPE united atom model,\cite{TraPPE-UA.alkanes} which provides good computational efficiency and reasonable accuracy for bulk thermal conductivity values. In this model,
176 + sites are located at the carbon centers for alkyl groups. Bonding
177 + interactions, including bond stretches and bends and torsions, were
178 + used for intra-molecular sites closer than 3 bonds. For non-bonded
179 + interactions, Lennard-Jones potentials were used.  We have previously
180 + utilized both united atom (UA) and all-atom (AA) force fields for
181 + thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
182 + atom force fields cannot populate the high-frequency modes that are
183 + present in AA force fields, they appear to work better for modeling
184 + thermal conductivity.
185 +
186 + Gold -- hexane nonbonded interactions are governed by pairwise Lennard-Jones parameters derived from Vlugt \emph{et al}.\cite{vlugt:cpc2007154} They fitted parameters for the interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the effective potential of Hautman and Klein for the Au(111) surface.\cite{hautman:4994}
187 +
188   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189 < % SHEAR VISCOSITY
189 > % THERMAL CONDUCTIVITIES
190   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191 < \subsection{Shear viscosity}
191 > \subsection{Thermal conductivities}
192  
193 < SPC/E Water Cluster:
193 > The thermal conductivity, $\lambda$, can be calculated using the imposed kinetic energy flux, $J_r$, and thermal gradient, $\frac{\partial T}{\partial r}$ measured from the resulting temperature profile
194  
195 + \begin{equation}
196 +        J_r = -\lambda \frac{\partial T}{\partial r}
197 + \end{equation}
198 +
199   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
200   % INTERFACIAL THERMAL CONDUCTANCE
201   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
202   \subsection{Interfacial thermal conductance}
203  
204 < Gold Nanoparticle in Hexane:
204 > A thermal flux is created using VSS-RNEMD moves, and the resulting temperature
205 > profiles are analyzed to yield information about the interfacial thermal
206 > conductance. As the simulation progresses, the VSS moves are performed on a regular basis, and
207 > the system develops a thermal or velocity gradient in response to the applied
208 > flux. A definition of the interfacial conductance in terms of a temperature difference ($\Delta T$) measured at $r_0$,
209 > \begin{equation}
210 >        G = \frac{J_r}{\Delta T_{r_0}}, \label{eq:G}
211 > \end{equation}
212 > is useful once the RNEMD approach has generated a
213 > stable temperature gap across the interface.
214  
215   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
216   % INTERFACIAL FRICTION
217   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
218   \subsection{Interfacial friction}
219  
220 < Table \ref{table:interfacialfriction} gives the calculated interfacial friction coefficients $\kappa$ for a spherical gold nanoparticle and prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied between the A and B regions defined as the gold structure and hexane molecules included in the convex hull, respectively. The resulting angular velocity gradient resulted in the gold structure rotating about the prescribed axis within the solvent.
220 > The slip length, $\delta$, is defined as the ratio of the interfacial friction coefficient, $\kappa$, and dynamic solvent viscosity, $\eta$
221  
222 + \begin{equation}
223 +        \delta = \frac{\eta}{\kappa}
224 + \end{equation}
225 +
226 + and is measured from a radial angular momentum profile generated from a nonperiodic VSS-RNEMD simulation.
227 +
228   Analytical solutions for the rotational friction coefficients for a solvated spherical body of radius $r$ under ``stick'' boundary conditions can be estimated using Stokes' law
229  
230   \begin{equation}
231 <        f_r = 8 \pi \eta r^3 \label{eq:fr}.
231 >        \Xi = 8 \pi \eta r^3 \label{eq:Xi}.
232   \end{equation}
233  
234 < For general ellipsoids with semiaxes $a$, $b$, and $c$, Perrin's extension of Stokes' law provides exact solutions for prolate $(a \geq b = c)$ and oblate $(a < b = c)$ ellipsoids. For a prolate ellipsoidal rod, demonstrated here,
234 > where $\eta$ is the dynamic viscosity of the surrounding solvent and was determined for TraPPE-UA hexane under these condition by applying a traditional VSS-RNEMD linear momentum flux to a periodic box of solvent.
235  
236 < \begin{eqnarray}
231 <        f_a = \label{eq:fa}\\
232 <        f_b = f_c = \label{eq:fb}
233 < \end{eqnarray}
236 > For general ellipsoids with semiaxes $a$, $b$, and $c$, Perrin's extension of Stokes' law provides exact solutions for symmetric prolate $(a \geq b = c)$ and oblate $(a < b = c)$ ellipsoids. For simplicity, we define a Perrin Factor, $S$,
237  
238 < The dynamic viscosity of the solvent, $\eta$, was calculated by applying a linear momentum flux to a periodic box of TraPPE-UA hexane.
238 > \begin{equation}
239 >        S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right]. \label{eq:S}
240 > \end{equation}
241  
242 < \begin{longtable}{p{3.8cm} p{3cm} p{2.8cm} p{2.5cm} p{2.5cm}}
243 <        \caption{Calculated interfacial friction coefficients ($\kappa$) and slip length ($\delta$) of gold nanostructures solvated in TraPPE-UA hexane. The ellipsoid is oriented with the long axis along the $z$ direction.}
244 <        \\ \hline \hline
245 <                {Structure} & \centering{Axis of rotation} & \centering\arraybackslash {$\kappa$} & \centering\arraybackslash {$\delta$} & \centering\arraybackslash Stokes' Law $F$\\
246 <        \centering {} & {} & \centering\arraybackslash {\small($10^4$ Pa s m$^{-1}$)} & \centering\arraybackslash {\small(nm)} & \centering\arraybackslash{\small()}\\ \hline
247 < \endhead
248 < \hline
249 < % \endfoot
250 < Nanoparticle & \centering$x = y = z$ & & & \\
251 < Prolate Ellipsoidal rod & \centering$x = y$ & & & \\
252 < Prolate Ellipsoidal rod & \centering$z$ & & &
242 > For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
243 > \begin{equation}
244 >        \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S}
245 > \label{eq:Xia}
246 > \end{equation}
247 > \begin{equation}
248 >        \Xi^{rr}_{b,c} = \frac{32 \pi}{2} \eta \frac{ \left( a^4 - b^4 \right)}{ \left( 2a^2 - b^2 \right)S - 2a}.                                      \label{eq:Xibc}
249 > \end{equation}
250 >
251 > % However, the friction between hexane solvent and gold more likely falls within ``slip'' boundary conditions. Hu and Zwanzig\cite{Zwanzig} investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained numerial results for a scaling factor as a function of $\tau$, the ratio of the shorter semiaxes and the longer semiaxis of the spheroid. For the sphere and prolate ellipsoid shown here, the values of $\tau$ are $1$ and $0.3939$, respectively. According to the values tabulated by Hu and Zwanzig, the sphere friction coefficient approaches $0$, while the ellipsoidal friction coefficient must be scaled by a factor of $0.880$ to account for the reduced interfacial friction under ``slip'' boundary conditions.
252 >
253 > Another useful quantity is the friction factor $f$, or the friction coefficient of a non-spherical shape divided by the friction coefficient of a sphere of equivalent volume. The nanoparticle and ellipsoidal nanorod dimensions used here were specifically chosen to create structures with equal volumes.
254 >
255 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
256 > % **TESTS AND APPLICATIONS**
257 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
258 > \section{Tests and Applications}
259 >
260 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
261 > % THERMAL CONDUCTIVITIES
262 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
263 > \subsection{Thermal conductivities}
264 >
265 > Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at different kinetic energy flux values are shown in Table \ref{table:goldconductivity}. For these calculations, the hot and cold slabs were excluded from the linear regression of the thermal gradient. The measured linear slope $\langle dT / dr \rangle$ is linearly dependent on the applied kinetic energy flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 -- 1.26 W m$^{-1}$ K$^{-1}$\cite{Kuang2010}, though still significantly lower than the experimental value of 320 W m$^{-1}$ K$^{-1}$, as the QSC force field neglects significant electronic contributions to heat conduction. The small increase relative to previous simulated bulk values is due to a slight increase in gold density -- as expected, an increase in density results in higher thermal conductivity values. The increased density is a result of nanoparticle curvature relative to an infinite bulk slab, which introduces surface tension that increases ambient density.
266 >
267 > \begin{longtable}{ccc}
268 > \caption{Calculated thermal conductivity of a crystalline gold nanoparticle of radius 40 \AA. Calculations were performed at 300 K and ambient density. Gold-gold interactions are described by the Quantum Sutton-Chen potential.}
269   \\ \hline \hline
270 < \label{table:interfacialfriction}
270 > {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
271 > {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
272 > 3.25$\times 10^{-6}$ & 0.11435 & 1.9753 \\
273 > 6.50$\times 10^{-6}$ & 0.2324 & 1.9438 \\
274 > 1.30$\times 10^{-5}$ & 0.44922 & 2.0113 \\
275 > 3.25$\times 10^{-5}$ & 1.1802 & 1.9139 \\
276 > 6.50$\times 10^{-5}$ & 2.339 & 1.9314
277 > \\ \hline \hline
278 > \label{table:goldconductivity}
279   \end{longtable}
280  
281 + Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table \ref{table:waterconductivity}.
282  
283 + \begin{longtable}{ccc}
284 + \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and ambient density.}
285 + \\ \hline \hline
286 + {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
287 + {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
288 + \\ \hline \hline
289 + \label{table:waterconductivity}
290 + \end{longtable}
291 +
292 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
293 + % INTERFACIAL THERMAL CONDUCTANCE
294 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
295 + \subsection{Interfacial thermal conductance}
296 +
297 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
298 + % INTERFACIAL FRICTION
299 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300 + \subsection{Interfacial friction}
301 +
302 + Table \ref{table:interfacialfrictionstick} shows the calculated interfacial friction coefficients $\kappa$ for a spherical gold nanoparticle and prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius, respectively. The resulting angular velocity gradient resulted in the gold structure rotating about the prescribed axis within the solvent.
303 +        
304 + \begin{longtable}{lccccc}
305 + \caption{Calculated ``stick'' interfacial friction coefficients ($\kappa$) and friction factors ($f$) of gold nanostructures solvated in TraPPE-UA hexane. The ellipsoid is oriented with the long axis along the $z$ direction.}
306 + \\ \hline \hline
307 + {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$} & {$f_{VSS}$} & {$f_{calc}$}\\
308 + {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)} & {} & {}\\  \hline
309 + {Sphere} & {$x = y = z$} & {} & {5.37237} & {1} & {1}\\
310 + {Prolate Ellipsoid} & {$x = y$} & {} & {3.59881} & {} & {0.768726}\\
311 + {Prolate Ellipsoid} & {$z$} & {} & {9.01084} & {} & {1.92477}\\  \hline \hline
312 + \label{table:interfacialfrictionstick}
313 + \end{longtable}
314 +
315 + % \begin{longtable}{lccc}
316 + % \caption{Calculated ``slip'' interfacial friction coefficients ($\kappa$) and friction factors ($f$) of gold nanostructures solvated in TraPPE-UA hexane. The ellipsoid is oriented with the long axis along the $z$ direction.}
317 + % \\ \hline \hline
318 + % {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$}\\
319 + % {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)}\\  \hline
320 + % {Sphere} & {$x = y = z$} & {} & {0}\\
321 + % {Prolate Ellipsoid} & {$x = y$} & {} & {0}\\
322 + % {Prolate Ellipsoid} & {$z$} & {} & {7.9295392}\\  \hline \hline
323 + % \label{table:interfacialfrictionslip}
324 + % \end{longtable}
325 +
326   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
327   % **DISCUSSION**
328   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 271 | Line 344 | Computing (CRC) at the University of Notre Dame.
344   \bibliography{nonperiodicVSS}
345  
346   \end{doublespace}
347 < \end{document}
347 > \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines