ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nonperiodicVSS/nonperiodicVSS.tex
Revision: 4003
Committed: Fri Jan 17 19:45:57 2014 UTC (10 years, 5 months ago) by kstocke1
Content type: application/x-tex
File size: 26607 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/npVSS}
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,Kuang2010} 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
126 The VSS-RNEMD approach uses a series of simultaneous velocity shearing
127 and scaling exchanges between the two slabs.\cite{Kuang2012}
128 This method imposes energy and momentum conservation constraints while
129 simultaneously creating a desired flux between the two slabs. These
130 constraints ensure that all configurations are sampled from the same
131 microcanonical (NVE) ensemble.
132
133 We have extended the VSS method for use in {\it non-periodic}
134 simulations, in which the ``slabs'' have been generalized to two
135 separated regions of space. These regions could be defined as
136 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
137 can be defined in terms of a dynamically changing ``hull'' comprising
138 the surface atoms of the cluster. This latter definition is identical
139 to the hull used in the Langevin Hull algorithm.
140
141 We present here a new set of constraints that are more general than
142 the VSS constraints. For the non-periodic variant, the constraints
143 fix both the total energy and total {\it angular} momentum of the
144 system while simultaneously imposing a thermal and angular momentum
145 flux between the two regions.
146
147 After each $\Delta t$ time interval, the particle velocities
148 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
149 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
150 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$).
151 \begin{displaymath}
152 \begin{array}{rclcl}
153 & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
154 \underline{\mathrm{rotational~shearing}} \\ \\
155 \mathbf{v}_i $~~~$\leftarrow &
156 a \left(\mathbf{v}_i - \langle \omega_a
157 \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
158 \mathbf{v}_j $~~~$\leftarrow &
159 b \left(\mathbf{v}_j - \langle \omega_b
160 \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
161 \end{array}
162 \end{displaymath}
163 Here $\langle\mathbf{\omega}_a\rangle$ and
164 $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
165 velocities of each shell, and $\mathbf{r}_i$ is the position of
166 particle $i$ relative to a fixed point in space (usually the center of
167 mass of the cluster). Particles in the shells also receive an
168 additive ``angular shear'' to their velocities. The amount of shear
169 is governed by the imposed angular momentum flux,
170 $\mathbf{j}_r(\mathbf{L})$,
171 \begin{eqnarray}
172 \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
173 \overleftrightarrow{I_a}^{-1} \Delta t + \langle \omega_a \rangle \label{eq:bc}\\
174 \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
175 \overleftrightarrow{I_b}^{-1} \Delta t + \langle \omega_b \rangle \label{eq:bh}
176 \end{eqnarray}
177 where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia tensor for
178 each of the two shells.
179
180 To simultaneously impose a thermal flux ($J_r$) between the shells we
181 use energy conservation constraints,
182 \begin{eqnarray}
183 K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
184 \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
185 \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
186 \cdot \mathbf{c}_a \label{eq:Kc}\\
187 K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
188 \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
189 \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
190 \end{eqnarray}
191 Simultaneous solution of these quadratic formulae for the scaling
192 coefficients, $a$ and $b$, will ensure that the simulation samples
193 from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$
194 is the instantaneous translational kinetic energy of each shell. At
195 each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
196 $\mathbf{c}_b$, subject to the imposed angular momentum flux,
197 $j_r(\mathbf{L})$, and thermal flux, $J_r$, values. The new particle
198 velocities are computed, and the simulation continues. System
199 configurations after the transformations have exactly the same energy
200 ({\it and} angular momentum) as before the moves.
201
202 As the simulation progresses, the velocity transformations can be
203 performed on a regular basis, and the system will develop a
204 temperature and/or angular velocity gradient in response to the
205 applied flux. Using the slope of the radial temperature or velocity
206 gradients, it is quite simple to obtain both the thermal conductivity
207 ($\lambda$), interfacial thermal conductance ($G$), or rotational friction coefficients ($\Xi^{rr}$) of any nonperiodic system.
208
209 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
210 % **COMPUTATIONAL DETAILS**
211 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
212 \section{Computational Details}
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.
221
222 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223 % SIMULATION PROTOCOL
224 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225 \subsection{Simulation protocol}
226
227 Systems containing liquids were run under moderate pressure ($\sim$ 5 atm) to avoid the formation of a substantial vapor phase. Thermal coupling to the Langevin Hull external bath was turned off to avoid interference with any imposed flux.
228
229 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
230 % FORCE FIELD PARAMETERS
231 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
232 \subsection{Force field parameters}
233
234 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}
235
236 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.
237
238 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,
239 sites are located at the carbon centers for alkyl groups. Bonding
240 interactions, including bond stretches and bends and torsions, were
241 used for intra-molecular sites closer than 3 bonds. For non-bonded
242 interactions, Lennard-Jones potentials were used. We have previously
243 utilized both united atom (UA) and all-atom (AA) force fields for
244 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
245 atom force fields cannot populate the high-frequency modes that are
246 present in AA force fields, they appear to work better for modeling
247 thermal conductivity.
248
249 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}
250
251 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
252 % THERMAL CONDUCTIVITIES
253 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254 \subsection{Thermal conductivities}
255
256 Fourier's Law of heat conduction in radial coordinates yields an expression for the heat flow between the concentric spherical RNEMD shells:
257
258 \begin{equation}
259 q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
260 \label{eq:Q}
261 \end{equation}
262
263 where $\lambda$ is the thermal conductivity, and $T_{a,b}$ and $r_{a,b}$ are the temperatures and radii of the two RNEMD regions, respectively.
264
265 Once a stable thermal gradient has been established between the two regions, the thermal conductivity, $\lambda$, can be calculated using a linear regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
266
267 \begin{equation}
268 \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
269 \label{eq:lambda}
270 \end{equation}
271
272 The rate of heat transfer between the two RNEMD regions is the amount of transferred kinetic energy over the length of the simulation, t
273
274 \begin{equation}
275 q_r = \frac{KE}{t}
276 \label{eq:heat}
277 \end{equation}
278
279 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
280 % INTERFACIAL THERMAL CONDUCTANCE
281 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
282 \subsection{Interfacial thermal conductance}
283
284 A thermal flux is created using VSS-RNEMD moves, and the temperature in each of the radial shells is recorded. The resulting temperature
285 profiles are analyzed to yield information about the interfacial thermal
286 conductance. As the simulation progresses, the VSS moves are performed on a regular basis, and
287 the system develops a thermal or velocity gradient in response to the applied
288 flux. We can treat the temperature in each radial shell as discrete, making the thermal conductance, $G$, of each shell the inverse of its Kapitza resistance, $R_K$. To model the thermal conductance across an interface (or multiple interfaces) it is useful to consider the shells as resistors wired in series. The total resistance of the shells is then additive, and the interfacial thermal conductance is the inverse of the total Kapitza resistance. The thermal resistance of each shell is
289
290 \begin{equation}
291 R_K = \frac{1}{q_r} \Delta T 4 \pi r^2
292 \label{eq:RK}
293 \end{equation}
294
295 making the total resistance of two neighboring shells
296
297 \begin{equation}
298 R_{total} = \frac{1}{q_r} \left [ (T_2 - T_1) 4 \pi r^2_1 + (T_3 - T_2) 4 \pi r^2_2 \right ]
299 \label{eq:Rtotal}
300 \end{equation}
301
302 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303 % INTERFACIAL FRICTION
304 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
305 \subsection{Interfacial friction}
306
307 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
308
309 \begin{equation}
310 \Xi^{rr} = 8 \pi \eta r^3
311 \label{eq:Xistick}.
312 \end{equation}
313
314 where $\eta$ is the dynamic viscosity of the surrounding solvent. An $\eta$ value for UA hexane under these particular temperature and pressure conditions was determined by applying a traditional VSS-RNEMD linear momentum flux to a periodic box of solvent.
315
316 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, $S$,
317
318 \begin{equation}
319 S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right].
320 \label{eq:S}
321 \end{equation}
322
323 For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
324 \begin{equation}
325 \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S}
326 \label{eq:Xia}
327 \end{equation}\vspace{-0.45in}\\
328 \begin{equation}
329 \Xi^{rr}_{b,c} = \frac{32 \pi}{2} \eta \frac{ \left( a^4 - b^4 \right)}{ \left( 2a^2 - b^2 \right)S - 2a}.
330 \label{eq:Xibc}
331 \end{equation}
332
333 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 ($\omega_{Au}$)
334
335 \begin{equation}
336 \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
337 \label{eq:Xieff}
338 \end{equation}
339
340 The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
341
342 \begin{equation}
343 \tau = \frac{L}{2 t}
344 \label{eq:tau}
345 \end{equation}
346
347 where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
348
349 Previous VSS-RNEMD simulations of the interfacial friction of the planar Au(111) / hexane interface have shown that the interface exists within ``slip'' boundary conditions.\cite{Kuang2012} Hu and Zwanzig\cite{Zwanzig} investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained numerial results for a scaling factor to be applied to $\Xi^{rr}_{\mathit{stick}}$ 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, $\Xi^{rr}_{\mathit{slip}}$ for any sphere approaches $0$, while the ellipsoidal $\Xi^{rr}_{\mathit{slip}}$ is the analytical $\Xi^{rr}_{\mathit{stick}}$ result scaled by a factor of $0.359$ to account for the reduced interfacial friction under ``slip'' boundary conditions.
350
351 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352 % **TESTS AND APPLICATIONS**
353 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354 \section{Tests and Applications}
355
356 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357 % THERMAL CONDUCTIVITIES
358 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
359 \subsection{Thermal conductivities}
360
361 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 $\langle \frac{dT}{dr} \rangle$ 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 {\footnotesize W / m $\cdot$ K}\cite{Kuang2010}, though still significantly lower than the experimental value of 320 {\footnotesize W / m $\cdot$ K}, as the QSC force field neglects significant electronic contributions to heat conduction.
362
363 % 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.
364
365 \begin{longtable}{ccc}
366 \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.}
367 \\ \hline \hline
368 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
369 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
370 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
371 6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
372 1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
373 3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
374 6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
375 \hline
376 This work & & 1.0040
377 \\ \hline \hline
378 \label{table:goldTC}
379 \end{longtable}
380
381 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 $\langle \frac{dT}{dr} \rangle$ fit. Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average calculated thermal conductivity from this work, $0.8841$ {\footnotesize W / m $\cdot$ K}, compares very well to previous nonequilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental values.\cite{WagnerKruse}
382
383 \begin{longtable}{ccc}
384 \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
385 \\ \hline \hline
386 {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
387 {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
388 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
389 3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
390 6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
391 \hline
392 This work & & 0.8841 \\
393 Zhang, et al\cite{Zhang2005} & & 0.81 \\
394 R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
395 Experiment\cite{WagnerKruse} & & 0.61
396 \\ \hline \hline
397 \label{table:waterTC}
398 \end{longtable}
399
400 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
401 % INTERFACIAL THERMAL CONDUCTANCE
402 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403 \subsection{Interfacial thermal conductance}
404
405 Calculated interfacial thermal conductance ($G$)
406
407 \begin{longtable}{ccc}
408 \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.}
409 \\ \hline \hline
410 {Nanoparticle Radius} & {$G$}\\
411 {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
412 20 & {47.1} \\
413 30 & {45.4} \\
414 40 & {46.5} \\
415 slab & {30.2}
416 \\ \hline \hline
417 \label{table:interfacialconductance}
418 \end{longtable}
419
420 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
421 % INTERFACIAL FRICTION
422 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
423 \subsection{Interfacial friction}
424
425 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.
426
427 \begin{longtable}{lccccc}
428 \caption{Comparison of rotational friction coefficients under ideal ``slip'' ($\Xi^{rr}_{\mathit{slip}}$) and ``stick'' conditions ($\Xi^{rr}_{\mathit{stick}}$) 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.}
429 \\ \hline \hline
430 {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
431 {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\ \hline
432 Sphere (r = 20 \AA) & {$x = y = z$} & 0 & {2386} & {3314} & {0.720}\\
433 Sphere (r = 30 \AA) & {$x = y = z$} & 0 & {8415} & {11749} & {0.716}\\
434 Sphere (r = 40 \AA) & {$x = y = z$} & 0 & {47544} & {34464} & {1.380}\\
435 Prolate Ellipsoid & {$x = y$} & {1792} & {3128} & {4991} & {0.627}\\
436 Prolate Ellipsoid & {$z$} & {716} & {1590} & {1993} & {0.798}
437 \\ \hline \hline
438 \label{table:couple}
439 \end{longtable}
440
441 The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold structures solvated by hexane do not exist in the ``slip'' boundary conditions. At this length scale, the nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short axis, which may at first seem counterintuitive. However, the `propellor' motion caused by rotation about short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius) appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of surface features or anomalous solvent behaviors that are not fully understood at this time.
442
443 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
444 % **DISCUSSION**
445 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
446 \section{Discussion}
447
448
449 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450 % **ACKNOWLEDGMENTS**
451 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452 \section*{Acknowledgments}
453
454 We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
455 this project was provided by the National Science Foundation under grant
456 CHE-0848243. Computational time was provided by the Center for Research
457 Computing (CRC) at the University of Notre Dame.
458
459 \newpage
460
461 \bibliography{nonperiodicVSS}
462
463 \end{doublespace}
464 \end{document}

Properties

Name Value
svn:executable