ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 3973
Committed: Thu Oct 31 20:43:00 2013 UTC (10 years, 8 months ago) by kstocke1
Content type: application/x-tex
File size: 20228 byte(s)
Log Message:

File Contents

# User Rev Content
1 kstocke1 3927 \documentclass[journal = jctcce, manuscript = article]{achemso}
2     \setkeys{acs}{usetitle = true}
3    
4     \usepackage{caption}
5 kstocke1 3944 \usepackage{endfloat}
6 kstocke1 3927 \usepackage{geometry}
7     \usepackage{natbib}
8     \usepackage{setspace}
9     \usepackage{xkeyval}
10     %%%%%%%%%%%%%%%%%%%%%%%
11     \usepackage{amsmath}
12     \usepackage{amssymb}
13     \usepackage{times}
14     \usepackage{mathptm}
15     \usepackage{setspace}
16     \usepackage{endfloat}
17     \usepackage{caption}
18     \usepackage{tabularx}
19     \usepackage{longtable}
20     \usepackage{graphicx}
21     \usepackage{multirow}
22     \usepackage{multicol}
23     \usepackage{achemso}
24     \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25     % \usepackage[square, comma, sort&compress]{natbib}
26     \usepackage{url}
27     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29     9.0in \textwidth 6.5in \brokenpenalty=10000
30    
31     % double space list of tables and figures
32     % \AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33     \setlength{\abovecaptionskip}{20 pt}
34     \setlength{\belowcaptionskip}{30 pt}
35    
36     % \bibpunct{}{}{,}{s}{}{;}
37    
38     % \citestyle{nature}
39     % \bibliographystyle{achemso}
40    
41     \title{A Method for Creating Thermal and Angular Momentum Fluxes in Non-Periodic Systems}
42    
43     \author{Kelsey M. Stocker}
44     \author{J. Daniel Gezelter}
45     \email{gezelter@nd.edu}
46     \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\ Department of Chemistry and Biochemistry\\ University of Notre Dame\\ Notre Dame, Indiana 46556}
47    
48     \begin{document}
49    
50     \newcolumntype{A}{p{1.5in}}
51     \newcolumntype{B}{p{0.75in}}
52    
53     % \author{Kelsey M. Stocker and J. Daniel
54     % Gezelter\footnote{Corresponding author. \ Electronic mail:
55     % gezelter@nd.edu} \\
56     % 251 Nieuwland Science Hall, \\
57     % Department of Chemistry and Biochemistry,\\
58     % University of Notre Dame\\
59     % Notre Dame, Indiana 46556}
60    
61     \date{\today}
62    
63     \maketitle
64    
65     \begin{doublespace}
66    
67     \begin{abstract}
68    
69     We have adapted the Velocity Shearing and Scaling Reverse Non-Equilibium Molecular Dynamics (VSS-RNEMD) method for use with aperiodic system geometries. This new method is capable of creating stable temperature and angular velocity gradients in heterogeneous non-periodic systems while conserving total energy and angular momentum. To demonstrate the method, we have computed the thermal conductivities of a gold nanoparticle and water cluster, the shear viscosity of a water cluster, the interfacial thermal conductivity of a solvated gold nanoparticle and the interfacial friction of solvated gold nanostructures.
70    
71     \end{abstract}
72    
73     \newpage
74    
75     %\narrowtext
76    
77     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
78     % **INTRODUCTION**
79     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
80     \section{Introduction}
81    
82    
83     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
84     % **METHODOLOGY**
85     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86     \section{Methodology}
87    
88     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89     % NON-PERIODIC DYNAMICS
90     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91     \subsection{Dynamics for non-periodic systems}
92    
93 kstocke1 3947 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 kstocke1 3927
96     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97     % NON-PERIODIC RNEMD
98     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
99     \subsection{VSS-RNEMD for non-periodic systems}
100    
101 kstocke1 3947 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 kstocke1 3927 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 kstocke1 3944 \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 kstocke1 3927 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
121     term ($\mathbf{a}_c$ and $\mathbf{a}_h$), and an additive angular velocity
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     \begin{displaymath}
126     \begin{array}{rclcl}
127     & \underline{~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~} & &
128     \underline{\mathrm{rotational \; shearing}} \\ \\
129     \mathbf{v}_i $~~~$\leftarrow &
130     c \, \left(\mathbf{v}_i - \langle \omega_c
131     \rangle \times r_i\right) & + & \mathbf{b}_c \times r_i \\
132     \mathbf{v}_j $~~~$\leftarrow &
133     h \, \left(\mathbf{v}_j - \langle \omega_h
134     \rangle \times r_j\right) & + & \mathbf{b}_h \times r_j
135     \end{array}
136     \end{displaymath}
137    
138     \begin{eqnarray}
139     \mathbf{b}_c & = & - \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_c^{-1} + \langle \omega_c \rangle \label{eq:bc}\\
140     \mathbf{b}_h & = & + \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_h^{-1} + \langle \omega_h \rangle \label{eq:bh}
141     \end{eqnarray}
142    
143     The total energy is constrained via two quadratic formulae,
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}
147     \end{eqnarray}
148    
149     the simultaneous
150     solution of which provide the velocity scaling coefficients $c$ and $h$. Given an
151     imposed angular momentum flux, $\mathbf{j}_{r} \left( \mathbf{L} \right)$, and/or
152     thermal flux, $J_r$, equations \ref{eq:bc} - \ref{eq:Kh} are sufficient to obtain
153     the velocity scaling ($c$ and $h$) and shearing ($\mathbf{b}_c,\,$ and $\mathbf{b}_h$) at each time interval.
154    
155     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156 kstocke1 3947 % **COMPUTATIONAL DETAILS**
157     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158     \section{Computational Details}
159    
160     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161     % SIMULATION PROTOCOL
162     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
163     \subsection{Simulation protocol}
164    
165    
166     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
167     % FORCE FIELD PARAMETERS
168     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169     \subsection{Force field parameters}
170    
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     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     % THERMAL CONDUCTIVITIES
190     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191     \subsection{Thermal conductivities}
192    
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     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     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     \Xi = 8 \pi \eta r^3 \label{eq:Xi}.
232     \end{equation}
233    
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     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     \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     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 kstocke1 3927 % **TESTS AND APPLICATIONS**
257     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
258     \section{Tests and Applications}
259    
260     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
261     % THERMAL CONDUCTIVITIES
262     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
263     \subsection{Thermal conductivities}
264    
265 kstocke1 3947 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 kstocke1 3927
267 kstocke1 3934 \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 kstocke1 3927 \\ \hline \hline
270 kstocke1 3934 {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
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 kstocke1 3927 \label{table:goldconductivity}
279     \end{longtable}
280    
281 kstocke1 3947 Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table \ref{table:waterconductivity}.
282    
283 kstocke1 3934 \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 kstocke1 3927 \\ \hline \hline
286 kstocke1 3934 {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
287     {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
288 kstocke1 3962 1.00$\times 10^{-5}$ & 0.38683 & 1.79665486\\
289     3.00$\times 10^{-5}$ & 1.1643 & 1.79077557\\
290     6.00$\times 10^{-5}$ & 2.2262 & 1.87314707\\
291 kstocke1 3973 \hline \hline
292 kstocke1 3927 \label{table:waterconductivity}
293     \end{longtable}
294    
295     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
296     % INTERFACIAL THERMAL CONDUCTANCE
297     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
298     \subsection{Interfacial thermal conductance}
299    
300 kstocke1 3973 \begin{longtable}{ccc}
301     \caption{Caption.}
302     \\ \hline \hline
303     {Nanoparticle Radius} & {$\boldsymbol \lambda$}\\
304     {\small(\AA)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
305     20 & 59.66\\
306     30 & 57.88\\
307     40 & \\
308     $\infty$ & \\
309     \hline \hline
310     \label{table:waterconductivity}
311     \end{longtable}
312 kstocke1 3962
313 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
314     % INTERFACIAL FRICTION
315     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
316     \subsection{Interfacial friction}
317    
318 kstocke1 3973 Table \ref{table:interfacialfrictionstick} shows the calculated interfacial friction coefficients $\kappa$ for spherical gold nanoparticles and a 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.
319 kstocke1 3934
320     \begin{longtable}{lccccc}
321 kstocke1 3973 \caption{Calculated ``stick'' interfacial friction coefficients ($\kappa$) and friction factors ($f$) of gold nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid is oriented with the long axis along the $z$ direction.}
322 kstocke1 3927 \\ \hline \hline
323 kstocke1 3973 {Structure} & {Axis of Rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$} & {$f_{VSS}$} & {$f_{calc}$}\\
324 kstocke1 3943 {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)} & {} & {}\\ \hline
325 kstocke1 3973 {Sphere (r = 20 \AA)} & {$x = y = z$} & {} & {6.13239} & {1} & {1}\\
326     {Sphere (r = 30 \AA)} & {$x = y = z$} & {} & {20.6968} & {1} & {1}\\
327     {Sphere (r = 40 \AA)} & {$x = y = z$} & {} & {49.0591} & {1} & {1}\\
328     {Prolate Ellipsoid} & {$x = y$} & {} & {8.22846} & {} & {1.92477}\\
329     {Prolate Ellipsoid} & {$z$} & {} & {3.28634} & {} & {0.768726}\\
330     \hline \hline
331 kstocke1 3947 \label{table:interfacialfrictionstick}
332 kstocke1 3927 \end{longtable}
333    
334 kstocke1 3947 % \begin{longtable}{lccc}
335     % \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.}
336     % \\ \hline \hline
337     % {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$}\\
338     % {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)}\\ \hline
339     % {Sphere} & {$x = y = z$} & {} & {0}\\
340     % {Prolate Ellipsoid} & {$x = y$} & {} & {0}\\
341     % {Prolate Ellipsoid} & {$z$} & {} & {7.9295392}\\ \hline \hline
342     % \label{table:interfacialfrictionslip}
343     % \end{longtable}
344 kstocke1 3943
345 kstocke1 3927 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
346     % **DISCUSSION**
347     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
348     \section{Discussion}
349    
350    
351     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352     % **ACKNOWLEDGMENTS**
353     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354     \section*{Acknowledgments}
355    
356     We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
357     this project was provided by the National Science Foundation under grant
358     CHE-0848243. Computational time was provided by the Center for Research
359     Computing (CRC) at the University of Notre Dame.
360    
361     \newpage
362    
363     \bibliography{nonperiodicVSS}
364    
365     \end{doublespace}
366 kstocke1 3934 \end{document}

Properties

Name Value
svn:executable