ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 3991
Committed: Wed Jan 8 22:23:49 2014 UTC (10 years, 6 months ago) by kstocke1
Content type: application/x-tex
File size: 25048 byte(s)
Log Message:

File Contents

# Content
1 \documentclass[journal = jctcce, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3
4 \usepackage{caption}
5 \usepackage{endfloat}
6 \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 Non-equilibrium Molecular Dynamics (NEMD) methods impose a temperature
83 or velocity {\it gradient} on a
84 system,\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
85 and use linear response theory to connect the resulting thermal or
86 momentum flux to transport coefficients of bulk materials. However,
87 for heterogeneous systems, such as phase boundaries or interfaces, it
88 is often unclear what shape of gradient should be imposed at the
89 boundary between materials.
90
91 % \begin{figure}
92 % \includegraphics[width=\linewidth]{figures/VSS}
93 % \caption{Schematics of periodic (left) and non-periodic (right)
94 % Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
95 % flux is applied from region B to region A. Thermal gradients are
96 % depicted by a color gradient. Linear or angular velocity gradients
97 % are shown as arrows.}
98 % \label{fig:VSS}
99 % \end{figure}
100
101 Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods impose an
102 unphysical {\it flux} between different regions or ``slabs'' of the
103 simulation
104 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
105 system responds by developing a temperature or velocity {\it gradient}
106 between the two regions. The gradients which develop in response to
107 the applied flux are then related (via linear response theory) to the
108 transport coefficient of interest. Since the amount of the applied
109 flux is known exactly, and measurement of a gradient is generally less
110 complicated, imposed-flux methods typically take shorter simulation
111 times to obtain converged results. At interfaces, the observed
112 gradients often exhibit near-discontinuities at the boundaries between
113 dissimilar materials. RNEMD methods do not need many trajectories to
114 provide information about transport properties, and they have become
115 widely used to compute thermal and mechanical transport in both
116 homogeneous liquids and
117 solids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
118 well as heterogeneous
119 interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
120
121 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
122 % **METHODOLOGY**
123 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124 \section{Velocity Shearing and Scaling (VSS) for non-periodic systems}
125 The VSS-RNEMD approach uses a series of simultaneous velocity shearing
126 and scaling exchanges between the two slabs.\cite{2012MolPh.110..691K}
127 This method imposes energy and momentum conservation constraints while
128 simultaneously creating a desired flux between the two slabs. These
129 constraints ensure that all configurations are sampled from the same
130 microcanonical (NVE) ensemble.
131
132 We have extended the VSS method for use in {\it non-periodic}
133 simulations, in which the ``slabs'' have been generalized to two
134 separated regions of space. These regions could be defined as
135 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
136 can be defined in terms of a dynamically changing ``hull'' comprising
137 the surface atoms of the cluster. This latter definition is identical
138 to the hull used in the Langevin Hull algorithm.
139
140 We present here a new set of constraints that are more general than
141 the VSS constraints. For the non-periodic variant, the constraints
142 fix both the total energy and total {\it angular} momentum of the
143 system while simultaneously imposing a thermal and angular momentum
144 flux between the two regions.
145
146 After each $\Delta t$ time interval, the particle velocities
147 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
148 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
149 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$).
150 \begin{displaymath}
151 \begin{array}{rclcl}
152 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
153 \underline{\mathrm{rotational~shearing}} \\ \\
154 \mathbf{v}_i $~~~$\leftarrow &
155 a \left(\mathbf{v}_i - \langle \omega_a
156 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
157 \mathbf{v}_j $~~~$\leftarrow &
158 b \left(\mathbf{v}_j - \langle \omega_b
159 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
160 \end{array}
161 \end{displaymath}
162 Here $\langle\mathbf{\omega}_a\rangle$ and
163 $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
164 velocities of each shell, and $\mathbf{r}_i$ is the position of
165 particle $i$ relative to a fixed point in space (usually the center of
166 mass of the cluster). Particles in the shells also receive an
167 additive ``angular shear'' to their velocities. The amount of shear
168 is governed by the imposed angular momentum flux,
169 $\mathbf{j}_r(\mathbf{L})$,
170 \begin{eqnarray}
171 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
172 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
173 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
174 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
175 \end{eqnarray}
176 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia tensor for
177 each of the two shells.
178
179 To simultaneously impose a thermal flux ($J_r$) between the shells we
180 use energy conservation constraints,
181 \begin{eqnarray}
182 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
183 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
184 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
185 \cdot \mathbf{c}_a \label{eq:Kc}\\
186 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
187 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
188 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
189 \end{eqnarray}
190 Simultaneous solution of these quadratic formulae for the scaling
191 coefficients, $a$ and $b$, will ensure that the simulation samples
192 from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$
193 is the instantaneous translational kinetic energy of each shell. At
194 each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
195 $\mathbf{c}_b$, subject to the imposed angular momentum flux,
196 $j_r(\mathbf{L})$, and thermal flux, $J_r$ values. The new particle
197 velocities are computed, and the simulation continues. System
198 configurations after the transformations have exactly the same energy
199 ({\it and} angular momentum) as before the moves.
200
201 As the simulation progresses, the velocity transformations can be
202 performed on a regular basis, and the system will develop a
203 temperature and/or angular velocity gradient in response to the
204 applied flux. Using the slope of the radial temperature or velocity
205 gradients, it is quite simple to obtain both the thermal conductivity
206 ($\lambda$) and shear viscosity ($\eta$),
207 \begin{equation}
208 J_r = -\lambda \frac{\partial T}{\partial
209 r} \hspace{2in} j_r(\mathbf{L}_z) = -\eta \frac{\partial
210 \omega_z}{\partial r}
211 \end{equation}
212 of a liquid cluster.
213
214 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
215 % NON-PERIODIC DYNAMICS
216 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
217 \subsection{Dynamics for non-periodic systems}
218
219 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
220 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 flux. Systems containing liquids were run under moderate pressure ($\sim$ 5 atm) to avoid the formation of a substantial vapor phase.
221
222 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223 % **COMPUTATIONAL DETAILS**
224 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225 \section{Computational Details}
226
227 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
228 % SIMULATION PROTOCOL
229 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
230 \subsection{Simulation protocol}
231
232
233 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
234 % FORCE FIELD PARAMETERS
235 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
236 \subsection{Force field parameters}
237
238 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}
239
240 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.
241
242 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,
243 sites are located at the carbon centers for alkyl groups. Bonding
244 interactions, including bond stretches and bends and torsions, were
245 used for intra-molecular sites closer than 3 bonds. For non-bonded
246 interactions, Lennard-Jones potentials were used. We have previously
247 utilized both united atom (UA) and all-atom (AA) force fields for
248 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
249 atom force fields cannot populate the high-frequency modes that are
250 present in AA force fields, they appear to work better for modeling
251 thermal conductivity.
252
253 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}
254
255 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
256 % THERMAL CONDUCTIVITIES
257 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
258 \subsection{Thermal conductivities}
259
260 Fourier's Law of heat conduction in radial coordinates is
261
262 \begin{equation}
263 q_r = -\lambda A \frac{dT}{dr}
264 \label{eq:fourier}
265 \end{equation}
266
267 Substituting the area of a sphere and integrating between $r = r_1$ and $r_2$ and $T = T_1$ and $T_2$, we arrive at an expression for the heat flow between the concentric spherical RNEMD shells:
268
269 \begin{equation}
270 q_r = - \frac{4 \pi \lambda (T_2 - T_1)}{\frac{1}{r_1} - \frac{1}{r_2}}
271 \label{eq:Q}
272 \end{equation}
273
274 Once a stable thermal gradient has been established between the two regions, the thermal conductivity, $\lambda$, can be calculated using the the temperature difference between the selected RNEMD regions, the radii of the two shells, and the heat, $q_r$, transferred between the regions.
275
276 \begin{equation}
277 \lambda = \frac{q_r (\frac{1}{r_2} - \frac{1}{r_1})}{4 \pi (T_2 - T_1)}
278 \label{eq:lambda}
279 \end{equation}
280
281 The heat transferred between the two RNEMD regions is the amount of transferred kinetic energy over the length of the simulation, t
282
283 \begin{equation}
284 q_r = \frac{KE}{t}
285 \label{eq:heat}
286 \end{equation}
287
288 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
289 % INTERFACIAL THERMAL CONDUCTANCE
290 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
291 \subsection{Interfacial thermal conductance}
292
293 A thermal flux is created using VSS-RNEMD moves, and the resulting temperature
294 profiles are analyzed to yield information about the interfacial thermal
295 conductance. As the simulation progresses, the VSS moves are performed on a regular basis, and
296 the system develops a thermal or velocity gradient in response to the applied
297 flux. A definition of the interfacial conductance in terms of a temperature difference ($\Delta T$) measured at $r_0$,
298 \begin{equation}
299 G = \frac{J_r}{\Delta T_{r_0}}, \label{eq:G}
300 \end{equation}
301 is useful once the RNEMD approach has generated a
302 stable temperature gap across the interface.
303
304 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
305 % INTERFACIAL FRICTION
306 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
307 \subsection{Interfacial friction}
308
309 Analytical solutions for the rotational friction coefficients for a solvated spherical body of radius $r$ under ideal ``stick'' boundary conditions can be estimated using Stokes' law
310
311 \begin{equation}
312 \Xi^{rr} = 8 \pi \eta r^3
313 \label{eq:Xistick}.
314 \end{equation}
315
316 where $\eta$ is the dynamic viscosity of the surrounding solvent and was determined for TraPPE-UA hexane by applying a traditional VSS-RNEMD linear momentum flux to a periodic box of solvent under the same temperature and pressure conditions as the nonperiodic systems.
317
318 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, $S$,
319
320 \begin{equation}
321 S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right]. \label{eq:S}
322 \end{equation}
323
324 For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
325 \begin{equation}
326 \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S}
327 \label{eq:Xia}
328 \end{equation}\vspace{-0.45in}\\
329 \begin{equation}
330 \Xi^{rr}_{b,c} = \frac{32 \pi}{2} \eta \frac{ \left( a^4 - b^4 \right)}{ \left( 2a^2 - b^2 \right)S - 2a}.
331 \label{eq:Xibc}
332 \end{equation}
333
334 The effective rotational friction coefficient at the interface can be extracted from nonperiodic VSS-RNEMD simulations quite easily using the applied torque ($\tau$) and the observed angular velocity of the gold structure ($\omega_{Au}$)
335
336 \begin{equation}
337 \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
338 \label{eq:Xieff}
339 \end{equation}
340
341 The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
342
343 \begin{equation}
344 \tau = \frac{L}{2 t}
345 \label{eq:tau}
346 \end{equation}
347
348 where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
349
350 % 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.
351
352 % 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.
353
354 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
355 % **TESTS AND APPLICATIONS**
356 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357 \section{Tests and Applications}
358
359 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
360 % THERMAL CONDUCTIVITIES
361 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
362 \subsection{Thermal conductivities}
363
364 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:goldTC}. 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.
365
366 % 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.
367
368 \begin{longtable}{ccc}
369 \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.}
370 \\ \hline \hline
371 {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
372 {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
373 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
374 6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
375 1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
376 3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
377 6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
378 \hline
379 This work & & 1.0040
380 \\ \hline \hline
381 \label{table:goldTC}
382 \end{longtable}
383
384 Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were excluded from the $\langle dT / dr \rangle$ fit. Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average calculated thermal conductivity from this work, $0.8841$ W m$^{-1}$ K$^{-1}$, compares very well to previous nonequilibrium molecular dynamics results (0.81 and 0.87 W m$^{-1}$ K$^{-1}$\cite{Romer2012, Zhang2005}) and experimental values (0.607 W m$^{-1}$ K$^{-1}$\cite{WagnerKruse})
385
386 \begin{longtable}{ccc}
387 \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
388 \\ \hline \hline
389 {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
390 {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
391 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
392 3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
393 6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
394 \hline
395 This work & & 0.8841 \\
396 Zhang, et al\cite{Zhang2005} & & 0.81 \\
397 R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
398 Experiment\cite{WagnerKruse} & & 0.61
399 \\ \hline \hline
400 \label{table:waterTC}
401 \end{longtable}
402
403 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
404 % INTERFACIAL THERMAL CONDUCTANCE
405 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
406 \subsection{Interfacial thermal conductance}
407
408 \begin{longtable}{ccc}
409 \caption{Calculated interfacial thermal conductance (G) values for gold nanoparticles of varying radii solvated in explicit TraPPE-UA hexane. The nanoparticle G values are compared to previous results for a gold slab in TraPPE-UA hexane, revealing increased interfacial thermal conductance for non-planar interfaces.}
410 \\ \hline \hline
411 {Nanoparticle Radius} & {G}\\
412 {\small(\AA)} & {\small(W m$^{-2}$ K$^{-1}$)}\\ \hline
413 20 & {49.3} \\
414 30 & {46.9} \\
415 40 & {47.3} \\
416 slab & {30.2} \\
417 \hline \hline
418 \label{table:interfacialconductance}
419 \end{longtable}
420
421 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
422 % INTERFACIAL FRICTION
423 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
424 \subsection{Interfacial friction}
425
426 Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ 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 causes the gold structure to rotate about the prescribed axis.
427
428 \begin{longtable}{lcccc}
429 \caption{Comparison of rotational friction coefficients under ideal ``stick'' conditions ($\Xi^{rr}_{stick}$) calculated via Stokes' and Perrin's laws and effective rotational friction coefficients ($\Xi^{rr}_{\mathit{eff}}$) of gold nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid is oriented with the long axis along the $z$ direction.}
430 \\ \hline \hline
431 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{stick}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{stick}$}\\
432 {} & {} & {\small(amu A$^2$ fs$^{-1}$)} & {\small(amu A$^2$ fs$^{-1}$)} & \\ \hline
433 Sphere (r = 20 \AA) & {$x = y = z$} & {3314} & {2386} & {0.720}\\
434 Sphere (r = 30 \AA) & {$x = y = z$} & {11749} & {8415} & {0.716}\\
435 Sphere (r = 40 \AA) & {$x = y = z$} & {34464} & {47544} & {1.380}\\
436 Prolate Ellipsoid & {$x = y$} & {4991} & {3128} & {0.627}\\
437 Prolate Ellipsoid & {$z$} & {1993} & {1590} & {0.798}\\
438 \hline \hline
439 \label{table:couple}
440 \end{longtable}
441
442 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
443 % **DISCUSSION**
444 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
445 \section{Discussion}
446
447
448 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
449 % **ACKNOWLEDGMENTS**
450 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
451 \section*{Acknowledgments}
452
453 We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
454 this project was provided by the National Science Foundation under grant
455 CHE-0848243. Computational time was provided by the Center for Research
456 Computing (CRC) at the University of Notre Dame.
457
458 \newpage
459
460 \bibliography{nonperiodicVSS}
461
462 \end{doublespace}
463 \end{document}

Properties

Name Value
svn:executable