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 3940 by kstocke1, Mon Aug 26 16:45:06 2013 UTC vs.
Revision 4060 by gezelter, Thu Mar 13 13:31:14 2014 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 12 | Line 12
12   \usepackage{amssymb}
13   \usepackage{times}
14   \usepackage{mathptm}
15 \usepackage{setspace}
16 \usepackage{endfloat}
15   \usepackage{caption}
16   \usepackage{tabularx}
17   \usepackage{longtable}
# Line 29 | Line 27
27   9.0in \textwidth 6.5in \brokenpenalty=10000
28  
29   % double space list of tables and figures
30 < % \AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
30 > % \AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
31   \setlength{\abovecaptionskip}{20 pt}
32   \setlength{\belowcaptionskip}{30 pt}
33  
# Line 66 | Line 64 | We have adapted the Velocity Shearing and Scaling Reve
64  
65   \begin{abstract}
66  
67 < 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.
67 > We have adapted the Velocity Shearing and Scaling Reverse Non-Equilibium Molecular Dynamics (VSS-RNEMD) method
68 > for use with non-periodic system geometries. This new method is capable of creating stable temperature and
69 > angular velocity gradients in heterogeneous non-periodic systems while conserving total energy and angular
70 > momentum. To demonstrate the method, we have computed the thermal conductivities of a gold nanoparticle and
71 > water cluster, the shear viscosity of a water cluster, the interfacial thermal conductivity of a solvated gold
72 > nanoparticle and the interfacial friction of solvated gold nanostructures.
73  
74   \end{abstract}
75  
# Line 79 | Line 82 | We have adapted the Velocity Shearing and Scaling Reve
82   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83   \section{Introduction}
84  
85 + Non-equilibrium Molecular Dynamics (NEMD) methods impose a temperature or velocity {\it gradient} on a
86 + system,\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v} and use linear response theory to connect the resulting thermal or
87 + momentum flux to transport coefficients of bulk materials.  However, for heterogeneous systems, such as phase
88 + boundaries or interfaces, it is often unclear what shape of gradient should be imposed at the boundary between
89 + materials.
90  
91 + Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods impose an unphysical {\it flux} between different
92 + regions or ``slabs'' of the simulation box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The system
93 + responds by developing a temperature or velocity {\it gradient} between the two regions. The gradients which
94 + develop in response to the applied flux are then related (via linear response theory) to the transport
95 + coefficient of interest. Since the amount of the applied flux is known exactly, and measurement of a gradient
96 + is generally less complicated, imposed-flux methods typically take shorter simulation times to obtain converged
97 + results. At interfaces, the observed gradients often exhibit near-discontinuities at the boundaries between
98 + dissimilar materials. RNEMD methods do not need many trajectories to provide information about transport
99 + properties, and they have become widely used to compute thermal and mechanical transport in both homogeneous
100 + liquids and solids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as well as heterogeneous
101 + interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
102 +
103 + The strengths of specific algorithms for imposing the flux between two
104 + different slabs of the simulation cell has been the subject of some
105 + renewed interest.  The original RNEMD approach used kinetic energy or
106 + momentum exchange between particles in the two slabs, either through
107 + direct swapping of momentum vectors or via virtual elastic collisions
108 + between atoms in the two regions.  There have been recent
109 + methodological advances which involve scaling all particle velocities
110 + in both slabs.  Constraint equations are simultaneously imposed to
111 + require the simulation to conserve both total energy and total linear
112 + momentum.  The most recent and simplest of the velocity scaling
113 + approaches allows for simultaneous shearing (to provide viscosity
114 + estimates) as well as scaling (to provide information about thermal
115 + conductivity).
116 +
117 + To date, however, the RNEMD methods have only been usable in periodic
118 + simulation cells where the exchange regions are physically separated
119 + along one of the axes of the simulation cell.   This limits the
120 + applicability to infinite planar interfaces.
121 +
122 + In order to model steady-state non-equilibrium distributions for
123 + curved surfaces (e.g. hot nanoparticles in contact with colder
124 + solvent), or for regions that are not planar slabs, the method
125 + requires some generalization for non-parallel exchange regions.  In
126 + the following sections, we present the Velocity Shearing and Scaling
127 + (VSS) RNEMD algorithm which has been explicitly designed for
128 + non-periodic simulations, and use the method to compute some thermal
129 + transport and solid-liquid friction at the surfaces of spherical and
130 + ellipsoidal nanoparticles, and discuss how the method can be extended
131 + to provide other kinds of non-equilibrium fluxes.
132 +
133   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134   % **METHODOLOGY**
135   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
136 < \section{Methodology}
136 > \section{Velocity Shearing and Scaling (VSS) for non-periodic systems}
137  
138 + The periodic VSS-RNEMD approach uses a series of simultaneous velocity
139 + shearing and scaling exchanges between the two slabs.\cite{Kuang2012}
140 + This method imposes energy and momentum conservation constraints while
141 + simultaneously creating a desired flux between the two slabs. These
142 + constraints ensure that all configurations are sampled from the same
143 + microcanonical (NVE) ensemble.
144 +
145 + \begin{figure}
146 + \includegraphics[width=\linewidth]{figures/npVSS}
147 + \caption{Schematics of periodic (left) and non-periodic (right)
148 +  Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
149 +  flux is applied from region B to region A. Thermal gradients are
150 +  depicted by a color gradient. Linear or angular velocity gradients
151 +  are shown as arrows.}
152 + \label{fig:VSS}
153 + \end{figure}
154 +
155 + We have extended the VSS method for use in {\it non-periodic} simulations, in which the ``slabs'' have been
156 + generalized to two separated regions of space. These regions could be defined as concentric spheres (as in
157 + figure \ref{fig:VSS}), or one of the regions can be defined in terms of a dynamically changing ``hull''
158 + comprising the surface atoms of the cluster. This latter definition is identical to the hull used in the
159 + Langevin Hull algorithm.
160 +
161 + We present here a new set of constraints that are more general than the VSS constraints. For the non-periodic
162 + variant, the constraints fix both the total energy and total {\it angular} momentum of the system while
163 + simultaneously imposing a thermal and angular momentum flux between the two regions.
164 +
165 + After each $\Delta t$ time interval, the particle velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two
166 + shells ($A$ and $B$) are modified by a velocity scaling coefficient ($a$ and $b$) and by a rotational shearing
167 + term ($\mathbf{c}_a$ and $\mathbf{c}_b$).
168 +
169 + \begin{displaymath}
170 + \begin{array}{rclcl}
171 + & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
172 + \underline{\mathrm{rotational~shearing}} \\  \\
173 + \mathbf{v}_i $~~~$\leftarrow &
174 +  a \left(\mathbf{v}_i - \langle \omega_a
175 +  \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
176 + \mathbf{v}_j $~~~$\leftarrow &
177 +  b  \left(\mathbf{v}_j - \langle \omega_b
178 +  \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
179 + \end{array}
180 + \end{displaymath}
181 + Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
182 + velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
183 + (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
184 + to their velocities. The amount of shear is governed by the imposed angular momentum flux,
185 + $\mathbf{j}_r(\mathbf{L})$,
186 + \begin{eqnarray}
187 + \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
188 + \overleftrightarrow{I_a}^{-1} \Delta t   + \langle \omega_a \rangle \label{eq:bc}\\
189 + \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
190 + \overleftrightarrow{I_b}^{-1}  \Delta t  + \langle \omega_b \rangle \label{eq:bh}
191 + \end{eqnarray}
192 + where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia tensor for each of the two shells.
193 +
194 + To simultaneously impose a thermal flux ($J_r$) between the shells we use energy conservation constraints,
195 + \begin{eqnarray}
196 + K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
197 + \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
198 + \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
199 + \cdot \mathbf{c}_a \label{eq:Kc}\\
200 + K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
201 + \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
202 + \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
203 + \end{eqnarray}
204 + Simultaneous solution of these quadratic formulae for the scaling coefficients, $a$ and $b$, will ensure that
205 + the simulation samples from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is the instantaneous
206 + translational kinetic energy of each shell. At each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
207 + $\mathbf{c}_b$, subject to the imposed angular momentum flux, $j_r(\mathbf{L})$, and thermal flux, $J_r$,
208 + values. The new particle velocities are computed, and the simulation continues. System configurations after the
209 + transformations have exactly the same energy ({\it and} angular momentum) as before the moves.
210 +
211 + As the simulation progresses, the velocity transformations can be performed on a regular basis, and the system
212 + will develop a temperature and/or angular velocity gradient in response to the applied flux. Using the slope of
213 + the radial temperature or velocity gradients, it is quite simple to obtain both the thermal conductivity
214 + ($\lambda$), interfacial thermal conductance ($G$), or rotational friction coefficients ($\Xi^{rr}$) of any
215 + non-periodic system.
216 +
217 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
218 + % **COMPUTATIONAL DETAILS**
219 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
220 + \section{Computational Details}
221 +
222 + The new VSS-RNEMD methodology for non-periodic system geometries has been implemented in our group molecular
223 + dynamics code, OpenMD.\cite{openmd} We have used the new method to calculate the thermal conductance of a gold
224 + nanoparticle and SPC/E water cluster, and compared the results with previous bulk RNEMD values, as well as
225 + experiment. We have also investigated the interfacial thermal conductance and interfacial rotational friction
226 + for gold nanostructures solvated in hexane as a function of nanoparticle size and shape.
227 +
228   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
229   % FORCE FIELD PARAMETERS
230   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
231   \subsection{Force field parameters}
232  
233 < 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.
233 > Gold -- gold interactions are described by the quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} The QSC
234 > parameters are tuned to experimental properties such as density, cohesive energy, and elastic moduli and
235 > include zero-point quantum corrections.
236  
237 < 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.
237 > We have chosen the SPC/E water model for these simulations, which is particularly useful for method validation
238 > as there are many values for physical properties from previous simulations available for direct
239 > comparison.\cite{Bedrov:2000, Kuang2010}
240  
241 < Hexane molecules are described by the TraPPE united atom model. This model provides good computational efficiency and reasonable accuracy for bulk thermal conductivity values.
241 > Hexane molecules are described by the TraPPE united atom model,\cite{TraPPE-UA.alkanes} which provides good
242 > computational efficiency and reasonable accuracy for bulk thermal conductivity values. In this model, sites are
243 > located at the carbon centers for alkyl groups. Bonding interactions, including bond stretches and bends and
244 > torsions, were used for intra-molecular sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
245 > potentials were used. We have previously utilized both united atom (UA) and all-atom (AA) force fields for
246 > thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united atom force fields cannot populate the
247 > high-frequency modes that are present in AA force fields, they appear to work better for modeling thermal
248 > conductivity.
249  
250 < Metal-nonmetal interactions are governed by parameters derived from Luedtke and Landman.
250 > Gold -- hexane nonbonded interactions are governed by pairwise Lennard-Jones parameters derived from Vlugt
251 > \emph{et al}.\cite{vlugt:cpc2007154} They fitted parameters for the interaction between Au and CH$_{\emph x}$
252 > pseudo-atoms based on the effective potential of Hautman and Klein for the Au(111) surface.\cite{hautman:4994}
253  
101
254   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
255   % NON-PERIODIC DYNAMICS
256   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
257 < \subsection{Dynamics for non-periodic systems}
257 > % \subsection{Dynamics for non-periodic systems}
258 > %
259 > % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
260 > % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
261 > % compressibilities, which are typically problematic for traditional affine transform methods. We have had
262 > % success applying this method to several different systems including bare metal nanoparticles, liquid water
263 > % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
264 > % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
265 > % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
266 > % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
267 > % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
268 > % subject to standard Newtonian dynamics.
269  
270 < 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.
270 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
271 > % SIMULATION PROTOCOL
272 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
273 > \subsection{Simulation protocol}
274  
275 + In all cases, systems were fully equilibrated under non-periodic isobaric-isothermal (NPT) conditions -- using
276 + the Langevin Hull methodology\cite{Vardeman2011} -- before any non-equilibrium methods were introduced. For
277 + heterogeneous systems, the gold nanoparticles and ellipsoid were first created from a bulk lattice and
278 + thermally equilibrated before being solvated in hexane. Packmol\cite{packmol} was used to solvate previously
279 + equilibrated gold nanostructures within a spherical droplet of hexane.
280 +
281 + Once fully equilibrated, a thermal or angular momentum flux was applied for 1 - 2
282 + ns, until a stable temperature or angular velocity gradient had developed. Systems containing liquids were run
283 + under moderate pressure (5 atm) and temperatures (230 K) to avoid the formation of a substantial vapor phase at the boundary of the cluster. Pressure was applied to the system via the non-periodic convex Langevin Hull. Thermal coupling to the external temperature and pressure bath was removed to avoid interference with any
284 + imposed flux.
285 +
286 + To stabilize the gold nanoparticle under the imposed angular momentum flux we altered the gold atom at the
287 + designated coordinate origin to have $10,000$ times its original mass. The nonbonded interactions remain
288 + unchanged and the heavy atom is excluded from the angular momentum exchange. For rotation of the ellipsoid about
289 + its long axis we have added two heavy atoms along the axis of rotation, one at each end of the rod. We collected angular velocity data for the heterogeneous systems after a brief VSS-RNEMD simulation to initialize rotation of the solvated nanostructure. Doing so ensures that we overcome the initial static friction and calculate only the \emph{dynamic} interfacial rotational friction.
290 +
291   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
292 < % NON-PERIODIC RNEMD
292 > % THERMAL CONDUCTIVITIES
293   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
294 < \subsection{VSS-RNEMD for non-periodic systems}
294 > \subsection{Thermal conductivities}
295  
296 < The adaptation of VSS-RNEMD for non-periodic systems is relatively
297 < 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
116 < of rectangular slabs. A temperature profile along the $r$ coordinate is created by recording the average temperature of concentric spherical shells.
296 > Fourier's Law of heat conduction in radial coordinates yields an expression for the heat flow between the
297 > concentric spherical RNEMD shells:
298  
299 < At each time interval, the particle velocities ($\mathbf{v}_i$ and
300 < $\mathbf{v}_j$) in the cold and hot shells ($C$ and $H$) are modified by a
301 < velocity scaling coefficient ($c$ and $h$), an additive linear velocity shearing
302 < 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.
299 > \begin{equation}
300 >        q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
301 > \label{eq:Q}
302 > \end{equation}
303  
304 < \begin{displaymath}
305 < \begin{array}{rclcl}
128 < & \underline{~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~} & &
129 < \underline{\mathrm{rotational \; shearing}} \\  \\
130 < \mathbf{v}_i $~~~$\leftarrow &
131 <  c \, \left(\mathbf{v}_i - \langle \omega_c
132 <  \rangle \times r_i\right) & + & \mathbf{b}_c \times r_i \\
133 < \mathbf{v}_j $~~~$\leftarrow &
134 <  h \, \left(\mathbf{v}_j - \langle \omega_h
135 <  \rangle \times r_j\right) & + & \mathbf{b}_h \times r_j
136 < \end{array}
137 < \end{displaymath}
304 > where $\lambda$ is the thermal conductivity, and $T_{a,b}$ and $r_{a,b}$ are the temperatures and radii of the
305 > two RNEMD regions, respectively.
306  
307 < \begin{eqnarray}
308 < \mathbf{b}_c & = & - \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_c^{-1} + \langle \omega_c \rangle \label{eq:bc}\\
309 < \mathbf{b}_h & = & + \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_h^{-1} + \langle \omega_h \rangle \label{eq:bh}
310 < \end{eqnarray}
307 > A thermal flux is created using VSS-RNEMD moves, and the temperature in each of the radial shells is recorded.
308 > The resulting temperature profiles are analyzed to yield information about the interfacial thermal conductance.
309 > As the simulation progresses, the VSS moves are performed on a regular basis, and the system develops a thermal
310 > or velocity gradient in response to the applied flux. Once a stable thermal gradient has been established
311 > between the two regions, the thermal conductivity, $\lambda$, can be calculated using a linear regression of
312 > the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
313  
314 < The total energy is constrained via two quadratic formulae,
314 > \begin{equation}
315 >        \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
316 > \label{eq:lambda}
317 > \end{equation}
318  
319 < \begin{eqnarray}
320 < 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}\\
148 < 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}
149 < \end{eqnarray}
319 > The rate of heat transfer between the two RNEMD regions is the amount of transferred kinetic energy over the
320 > length of the simulation, t
321  
322 < the simultaneous
323 < solution of which provide the velocity scaling coefficients $c$ and $h$. Given an
324 < imposed angular momentum flux, $\mathbf{j}_{r} \left( \mathbf{L} \right)$, and/or
325 < thermal flux, $J_r$, equations \ref{eq:bc} - \ref{eq:Kh} are sufficient to obtain
326 < the velocity scaling ($c$ and $h$) and shearing ($\mathbf{b}_c,\,$ and $\mathbf{b}_h$) at each time interval.
322 > \begin{equation}
323 >        q_r = \frac{KE}{t}
324 > \label{eq:heat}
325 > \end{equation}
326 >
327 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
328 > % INTERFACIAL THERMAL CONDUCTANCE
329 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330 > \subsection{Interfacial thermal conductance}
331  
332 + \begin{figure}
333 + \includegraphics[width=\linewidth]{figures/NP20}
334 + \caption{A gold nanoparticle with a radius of 20 \AA$\,$ solvated in TraPPE-UA hexane. A thermal flux is applied        between the nanoparticle and an outer shell of solvent.}
335 + \label{fig:NP20}
336 + \end{figure}
337 +
338 + For heterogeneous systems such as a solvated nanoparticle, shown in Figure \ref{fig:NP20}, the interfacial
339 + thermal conductance at the surface of the nanoparticle can be determined using an applied kinetic energy flux.
340 + We can treat the temperature in each radial shell as discrete, making the thermal conductance, $G$, of each
341 + shell the inverse of its Kapitza resistance, $R_K$. To model the thermal conductance across an interface (or
342 + multiple interfaces) it is useful to consider the shells as resistors wired in series. The resistance of the
343 + shells is then additive, and the interfacial thermal conductance is the inverse of the total Kapitza
344 + resistance. The thermal resistance of each shell is
345 +
346 + \begin{equation}
347 +        R_K = \frac{1}{q_r} \Delta T 4 \pi r^2
348 + \label{eq:RK}
349 + \end{equation}
350 +
351 + making the total resistance of two neighboring shells
352 +
353 + \begin{equation}
354 +        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 ] = \frac{1}{G}
355 + \label{eq:Rtotal}
356 + \end{equation}
357 +
358 + This series can be expanded for any number of adjacent shells, allowing for the calculation of the interfacial
359 + thermal conductance for interfaces of considerable thickness, such as self-assembled ligand monolayers on a
360 + metal surface.
361 +
362 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363 + % INTERFACIAL ROTATIONAL FRICTION
364 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
365 + \subsection{Interfacial rotational friction}
366 +
367 + The interfacial rotational friction, $\Xi^{rr}$, can be calculated for heterogeneous nanostructure/solvent
368 + systems by applying an angular momentum flux between the solvated nanostructure and a spherical shell of
369 + solvent at the boundary of the cluster. An angular velocity gradient develops in response to the applied flux,
370 + causing the nanostructure and solvent shell to rotate in opposite directions about a given axis.
371 +
372 + \begin{figure}
373 + \includegraphics[width=\linewidth]{figures/E25-75}
374 + \caption{A gold prolate ellipsoid of length 65 \AA$\,$ and width 25 \AA$\,$ solvated by TraPPE-UA hexane. An angular momentum flux is applied between the ellipsoid and an outer shell of solvent.}
375 + \label{fig:E25-75}
376 + \end{figure}
377 +
378 + 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
379 +
380 + \begin{equation}
381 +        \Xi^{rr}_{stick} = 8 \pi \eta r^3
382 + \label{eq:Xisphere}.
383 + \end{equation}
384 +
385 + where $\eta$ is the dynamic viscosity of the surrounding solvent. An $\eta$ value for TraPPE-UA hexane under
386 + these particular temperature and pressure conditions was determined by applying a traditional VSS-RNEMD linear
387 + momentum flux to a periodic box of solvent.
388 +
389 + For general ellipsoids with semiaxes $a$, $b$, and $c$, Perrin's extension of Stokes' law provides exact
390 + solutions for symmetric prolate $(a \geq b = c)$ and oblate $(a < b = c)$ ellipsoids under ideal ``stick'' conditions. For simplicity, we define
391 + a Perrin Factor, $S$,
392 +
393 + \begin{equation}
394 +        S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right].
395 + \label{eq:S}
396 + \end{equation}
397 +
398 + For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
399 + \begin{equation}
400 +        \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S}
401 + \label{eq:Xia}
402 + \end{equation}\vspace{-0.45in}\\
403 + \begin{equation}
404 +        \Xi^{rr}_{b,c} = \frac{32 \pi}{2} \eta \frac{ \left( a^4 - b^4 \right)}{ \left( 2a^2 - b^2 \right)S - 2a}.
405 + \label{eq:Xibc}
406 + \end{equation}
407 +
408 + corresponding to rotation about the long axis ($a$), and each of the equivalent short axes ($b$ and $c$), respectively.
409 +
410 + Previous VSS-RNEMD simulations of the interfacial friction of the planar Au(111) / hexane interface have shown
411 + that the interface exists within ``slip'' boundary conditions.\cite{Kuang2012} Hu and Zwanzig\cite{Zwanzig}
412 + investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained
413 + numerial results for a scaling factor to be applied to $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau$, the
414 + ratio of the shorter semiaxes and the longer semiaxis of the spheroid. For the sphere and prolate ellipsoid
415 + shown here, the values of $\tau$ are $1$ and $0.3939$, respectively. Under ``slip'' conditions,
416 + $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate ellipsoid about its long axis approaches
417 + $0$, as no solvent is displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for rotation of the
418 + prolate ellipsoid about its short axis is $35.9\%$ of the analytical $\Xi^{rr}_{\mathit{stick}}$ result,
419 + accounting for the reduced interfacial friction under ``slip'' boundary conditions.
420 +
421 + The effective rotational friction coefficient, $\Xi^{rr}_{\mathit{eff}}$ at the interface can be extracted from non-periodic VSS-RNEMD simulations quite easily using the applied torque ($\tau$) and the observed angular velocity of the gold structure ($\omega_{Au}$)
422 +
423 + \begin{equation}
424 +        \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
425 + \label{eq:Xieff}
426 + \end{equation}
427 +
428 + The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
429 +
430 + \begin{equation}
431 +        \tau = \frac{L}{2 t}
432 + \label{eq:tau}  
433 + \end{equation}
434 +
435 + where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
436 +
437   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
438   % **TESTS AND APPLICATIONS**
439   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 164 | Line 444 | Calculated values for the thermal conductivity of a 40
444   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
445   \subsection{Thermal conductivities}
446  
447 < 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.
447 > Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at
448 > different kinetic energy flux values are shown in Table \ref{table:goldTC}. For these calculations, the hot and
449 > cold slabs were excluded from the linear regression of the thermal gradient.
450  
451   \begin{longtable}{ccc}
452   \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.}
453   \\ \hline \hline
454 < {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
455 < {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
456 < 3.25$\times 10^{-6}$ & 0.11435 & 1.9753 \\
457 < 6.50$\times 10^{-6}$ & 0.2324 & 1.9438 \\
458 < 1.30$\times 10^{-5}$ & 0.44922 & 2.0113 \\
459 < 3.25$\times 10^{-5}$ & 1.1802 & 1.9139 \\
460 < 6.50$\times 10^{-5}$ & 2.339 & 1.9314
454 > {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
455 > {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
456 > 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
457 > 6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
458 > 1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
459 > 3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
460 > 6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
461 > \hline
462 > This work & & 1.0040
463   \\ \hline \hline
464 < \label{table:goldconductivity}
464 > \label{table:goldTC}
465   \end{longtable}
182        
183 SPC/E Water Cluster:
466  
467 + The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly dependent on the applied kinetic energy
468 + flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 -- 1.26 W / m $\cdot$ K\cite{Kuang2010}, though still significantly lower than the experimental value
469 + of 320 W / m $\cdot$ K, as the QSC force field neglects significant electronic contributions to
470 + heat conduction.
471 +
472 + Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table
473 + \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were
474 + excluded from the $\langle \frac{dT}{dr} \rangle$ fit.
475 +
476   \begin{longtable}{ccc}
477 < \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and ambient density.}
477 > \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
478   \\ \hline \hline
479 < {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
480 < {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
479 > {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
480 > {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
481 > 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
482 > 3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
483 > 6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
484 > \hline
485 > This work & & 0.8841 \\
486 > Zhang, et al\cite{Zhang2005} & & 0.81 \\
487 > R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
488 > Experiment\cite{WagnerKruse} & & 0.61
489   \\ \hline \hline
490 < \label{table:waterconductivity}
490 > \label{table:waterTC}
491   \end{longtable}
492  
493 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
494 < % SHEAR VISCOSITY
495 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
496 < \subsection{Shear viscosity}
493 > Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average
494 > calculated thermal conductivity from this work, $0.8841$ W / m $\cdot$ K, compares very well to
495 > previous non-equilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental
496 > values.\cite{WagnerKruse}
497  
199 SPC/E Water Cluster:
200
498   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499   % INTERFACIAL THERMAL CONDUCTANCE
500   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501   \subsection{Interfacial thermal conductance}
502  
503 < The interfacial thermal conductance, $G$, is calculated by defining a temperature difference $\Delta T$ across a given interface.
503 > Calculated interfacial thermal conductance ($G$) values for three sizes of gold nanoparticles and Au(111)
504 > surface solvated in TraPPE-UA hexane are shown in Table \ref{table:G}.
505  
506 + \begin{longtable}{ccc}
507 + \caption{Calculated interfacial thermal conductance ($G$) values for gold nanoparticles of varying radii solvated in TraPPE-UA hexane. The nanoparticle $G$ values are compared to previous simulation results for a Au(111) interface in TraPPE-UA hexane.}
508 + \\ \hline \hline
509 + {Nanoparticle Radius} & {$G$}\\
510 + {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
511 + 20 & {47.1} \\
512 + 30 & {45.4} \\
513 + 40 & {46.5} \\
514 + \hline
515 + Au(111) & {30.2}
516 + \\ \hline \hline
517 + \label{table:G}
518 + \end{longtable}
519 +
520 + The introduction of surface curvature increases the interfacial thermal conductance by a factor of
521 + approximately $1.5$ relative to the flat interface. There are no significant differences in the $G$ values for
522 + the varying nanoparticle sizes. It seems likely that for the range of nanoparticle sizes represented here, any
523 + particle size effects are not evident. The simulation of larger nanoparticles may demonstrate an approach to the $G$ value of a flat Au(111) slab but would require prohibitively costly numbers of atoms.
524 +
525   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
526   % INTERFACIAL FRICTION
527   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528   \subsection{Interfacial friction}
529  
530 < The interfacial friction coefficient, $\kappa$, can be calculated from the solvent dynamic viscosity, $\eta$, and the slip length, $\delta$. The slip length is measured from a radial angular momentum profile generated from a nonperiodic VSS-RNEMD simulation, as shown in Figure X.
530 > Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold
531 > nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied
532 > between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius,
533 > respectively.
534  
215 Table \ref{table:interfacialfriction} 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 included in the convex hull, respectively. The resulting angular velocity gradient resulted in the gold structure rotating about the prescribed axis within the solvent.
216
217 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
218
219 \begin{equation}
220        \Xi = 8 \pi \eta r^3 \label{eq:Xi}.
221 \end{equation}
222
223 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.
224
225 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, $S$,
226
227 \begin{equation}
228        S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right]. \label{eq:S}
229 \end{equation}
230
231 For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
232
233 \begin{eqnarray}
234        \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S} \label{eq:Xia}\\
235        \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}
236 \end{eqnarray}
237
238 However, the friction between hexane solvent and gold more likely falls within ``slip'' boundary conditions. Hu and 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.
239
240 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.
241        
535   \begin{longtable}{lccccc}
536 < \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.}
536 > \caption{Comparison of rotational friction coefficients under ideal ``slip'' ($\Xi^{rr}_{\mathit{slip}}$) and ``stick'' ($\Xi^{rr}_{\mathit{stick}}$) conditions and effective ($\Xi^{rr}_{\mathit{eff}}$) rotational friction coefficients of gold nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid is oriented with the long axis along the $z$ direction.}
537   \\ \hline \hline
538 < {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$} & {$f_{VSS}$} & {$f_{calc}$}\\
539 < {} & {} & {\small($10^4$ Pa s m$^{-1}$)} & {\small($10^4$ Pa s m$^{-1}$)} & {} & {}\\  \hline
540 < {Sphere} & {$x = y = z$} & {} & {} & {1} & {1}\\
541 < {Prolate Ellipsoid} & {$x = y$} & {} & {} & {} & {}\\
542 < {Prolate Ellipsoid} & {$z$} & {} & {} & {} & {}\\  \hline \hline
543 < \label{table:interfacialfriction}
538 > {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
539 > {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\  \hline
540 > Sphere (r = 20 \AA) & {$x = y = z$} & 0 & {2386} & {3314} & {0.720}\\
541 > Sphere (r = 30 \AA) & {$x = y = z$} & 0 & {8415} & {11749} & {0.716}\\
542 > Sphere (r = 40 \AA) & {$x = y = z$} & 0 & {47544} & {34464} & {1.380}\\
543 > Prolate Ellipsoid & {$x = y$} & {1792} & {3128} & {4991} & {0.627}\\
544 > Prolate Ellipsoid & {$z$} & 0 & {1590} & {1993} & {0.798}
545 > \\ \hline \hline
546 > \label{table:couple}
547   \end{longtable}
548  
549 + The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold
550 + structures solvated by hexane do not exist in the ``slip'' boundary condition. At this length scale, the
551 + nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience
552 + increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius
553 + nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of
554 + the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short
555 + axis, which at first seems counterintuitive. However, the `propellor' motion caused by rotation about the
556 + short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the
557 + gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius)
558 + appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of
559 + surface features or anomalous solvent behaviors that are not fully understood at this time.
560 +
561   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
562   % **DISCUSSION**
563   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
564   \section{Discussion}
565  
566 + We have demonstrated a novel adaptation of the VSS-RNEMD methodology for the application of thermal and angular momentum fluxes in explicitly non-periodic system geometries. Non-periodic VSS-RNEMD preserves the virtues of traditional VSS-RNEMD, namely Boltzmann thermal velocity distributions and minimal thermal anisotropy, while extending the constraints to conserve total energy and total \emph{angular} momentum. We also still have the ability to impose the thermal and angular momentum fluxes simultaneously or individually.
567  
568 + Most strikingly, this method enables calculation of thermal conductivity in homogeneous non-periodic geometries, as well as interfacial thermal conductance and interfacial rotational friction in heterogeneous clusters. The ability to interrogate explicitly non-periodic effects -- such as surface curvature -- on interfacial transport properties is an exciting prospect that will be explored in the future.
569 +
570   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
571   % **ACKNOWLEDGMENTS**
572   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
573   \section*{Acknowledgments}
574  
575 < We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
576 < this project was provided by the National Science Foundation under grant
266 < CHE-0848243. Computational time was provided by the Center for Research
575 > We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for this project was provided by the
576 > National Science Foundation under grant CHE-0848243. Computational time was provided by the Center for Research
577   Computing (CRC) at the University of Notre Dame.
578  
579   \newpage

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines