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 3944 by kstocke1, Tue Sep 3 19:30:21 2013 UTC vs.
Revision 4058 by kstocke1, Thu Mar 6 15:28:21 2014 UTC

# Line 2 | Line 2
2   \setkeys{acs}{usetitle = true}
3  
4   \usepackage{caption}
5 < \usepackage{endfloat}
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:2
87 + 002dp,PhysRevA.34.1449,JiangHao_jp802942v} and use linear response theory to connect the resulting thermal or
88 + momentum flux to transport coefficients of bulk materials. However, for heterogeneous systems, such as phase
89 + boundaries or interfaces, it is often unclear what shape of gradient should be imposed at the boundary between
90 + materials.
91  
92 + Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods impose an unphysical {\it flux} between different
93 + regions or ``slabs'' of the simulation box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The system
94 + responds by developing a temperature or velocity {\it gradient} between the two regions. The gradients which
95 + develop in response to the applied flux are then related (via linear response theory) to the transport
96 + coefficient of interest. Since the amount of the applied flux is known exactly, and measurement of a gradient
97 + is generally less complicated, imposed-flux methods typically take shorter simulation times to obtain converged
98 + results. At interfaces, the observed gradients often exhibit near-discontinuities at the boundaries between
99 + dissimilar materials. RNEMD methods do not need many trajectories to provide information about transport
100 + properties, and they have become widely used to compute thermal and mechanical transport in both homogeneous
101 + liquids and solids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as well as heterogeneous
102 + interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
103 +
104 +
105   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106   % **METHODOLOGY**
107   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
108 < \section{Methodology}
108 > \section{Velocity Shearing and Scaling (VSS) for non-periodic systems}
109  
110 + The periodic VSS-RNEMD approach uses a series of simultaneous velocity shearing and scaling exchanges between the two
111 + slabs.\cite{Kuang2012} This method imposes energy and momentum conservation constraints while simultaneously
112 + creating a desired flux between the two slabs. These constraints ensure that all configurations are sampled
113 + from the same microcanonical (NVE) ensemble.
114 +
115 + \begin{figure}
116 + \includegraphics[width=\linewidth]{figures/npVSS}
117 + \caption{Schematics of periodic (left) and non-periodic (right)
118 +  Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum
119 +  flux is applied from region B to region A. Thermal gradients are
120 +  depicted by a color gradient. Linear or angular velocity gradients
121 +  are shown as arrows.}
122 + \label{fig:VSS}
123 + \end{figure}
124 +
125 + We have extended the VSS method for use in {\it non-periodic} simulations, in which the ``slabs'' have been
126 + generalized to two separated regions of space. These regions could be defined as concentric spheres (as in
127 + figure \ref{fig:VSS}), or one of the regions can be defined in terms of a dynamically changing ``hull''
128 + comprising the surface atoms of the cluster. This latter definition is identical to the hull used in the
129 + Langevin Hull algorithm.
130 +
131 + We present here a new set of constraints that are more general than the VSS constraints. For the non-periodic
132 + variant, the constraints fix both the total energy and total {\it angular} momentum of the system while
133 + simultaneously imposing a thermal and angular momentum flux between the two regions.
134 +
135 + After each $\Delta t$ time interval, the particle velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two
136 + shells ($A$ and $B$) are modified by a velocity scaling coefficient ($a$ and $b$) and by a rotational shearing
137 + term ($\mathbf{c}_a$ and $\mathbf{c}_b$).
138 +
139 + \begin{displaymath}
140 + \begin{array}{rclcl}
141 + & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
142 + \underline{\mathrm{rotational~shearing}} \\  \\
143 + \mathbf{v}_i $~~~$\leftarrow &
144 +  a \left(\mathbf{v}_i - \langle \omega_a
145 +  \rangle \times \mathbf{r}_i\right) & + & \mathbf{c}_a \times \mathbf{r}_i \\
146 + \mathbf{v}_j $~~~$\leftarrow &
147 +  b  \left(\mathbf{v}_j - \langle \omega_b
148 +  \rangle \times \mathbf{r}_j\right) & + & \mathbf{c}_b \times \mathbf{r}_j
149 + \end{array}
150 + \end{displaymath}
151 + Here $\langle\mathbf{\omega}_a\rangle$ and $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
152 + velocities of each shell, and $\mathbf{r}_i$ is the position of particle $i$ relative to a fixed point in space
153 + (usually the center of mass of the cluster). Particles in the shells also receive an additive ``angular shear''
154 + to their velocities. The amount of shear is governed by the imposed angular momentum flux,
155 + $\mathbf{j}_r(\mathbf{L})$,
156 + \begin{eqnarray}
157 + \mathbf{c}_a & = & - \mathbf{j}_r(\mathbf{L}) \cdot
158 + \overleftrightarrow{I_a}^{-1} \Delta t   + \langle \omega_a \rangle \label{eq:bc}\\
159 + \mathbf{c}_b & = & + \mathbf{j}_r(\mathbf{L}) \cdot
160 + \overleftrightarrow{I_b}^{-1}  \Delta t  + \langle \omega_b \rangle \label{eq:bh}
161 + \end{eqnarray}
162 + where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment of inertia tensor for each of the two shells.
163 +
164 + To simultaneously impose a thermal flux ($J_r$) between the shells we use energy conservation constraints,
165 + \begin{eqnarray}
166 + K_a - J_r \Delta t & = & a^2 \left(K_a - \frac{1}{2}\langle
167 + \omega_a \rangle \cdot \overleftrightarrow{I_a} \cdot \langle \omega_a
168 + \rangle \right) + \frac{1}{2} \mathbf{c}_a \cdot \overleftrightarrow{I_a}
169 + \cdot \mathbf{c}_a \label{eq:Kc}\\
170 + K_b + J_r \Delta t & = & b^2 \left(K_b - \frac{1}{2}\langle
171 + \omega_b \rangle \cdot \overleftrightarrow{I_b} \cdot \langle \omega_b
172 + \rangle \right) + \frac{1}{2} \mathbf{c}_b \cdot \overleftrightarrow{I_b} \cdot \mathbf{c}_b \label{eq:Kh}
173 + \end{eqnarray}
174 + Simultaneous solution of these quadratic formulae for the scaling coefficients, $a$ and $b$, will ensure that
175 + the simulation samples from the original microcanonical (NVE) ensemble. Here $K_{\{a,b\}}$ is the instantaneous
176 + translational kinetic energy of each shell. At each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
177 + $\mathbf{c}_b$, subject to the imposed angular momentum flux, $j_r(\mathbf{L})$, and thermal flux, $J_r$,
178 + values. The new particle velocities are computed, and the simulation continues. System configurations after the
179 + transformations have exactly the same energy ({\it and} angular momentum) as before the moves.
180 +
181 + As the simulation progresses, the velocity transformations can be performed on a regular basis, and the system
182 + will develop a temperature and/or angular velocity gradient in response to the applied flux. Using the slope of
183 + the radial temperature or velocity gradients, it is quite simple to obtain both the thermal conductivity
184 + ($\lambda$), interfacial thermal conductance ($G$), or rotational friction coefficients ($\Xi^{rr}$) of any
185 + non-periodic system.
186 +
187 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
188 + % **COMPUTATIONAL DETAILS**
189 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
190 + \section{Computational Details}
191 +
192 + The new VSS-RNEMD methodology for non-periodic system geometries has been implemented in our group molecular
193 + dynamics code, OpenMD.\cite{openmd} We have used the new method to calculate the thermal conductance of a gold
194 + nanoparticle and SPC/E water cluster, and compared the results with previous bulk RNEMD values, as well as
195 + experiment. We have also investigated the interfacial thermal conductance and interfacial rotational friction
196 + for gold nanostructures solvated in hexane as a function of nanoparticle size and shape.
197 +
198   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
199   % FORCE FIELD PARAMETERS
200   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
201   \subsection{Force field parameters}
202  
203 < 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.
203 > Gold -- gold interactions are described by the quantum Sutton-Chen (QSC) model.\cite{PhysRevB.59.3527} The QSC
204 > parameters are tuned to experimental properties such as density, cohesive energy, and elastic moduli and
205 > include zero-point quantum corrections.
206  
207 < 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.
207 > We have chosen the SPC/E water model for these simulations, which is particularly useful for method validation
208 > as there are many values for physical properties from previous simulations available for direct
209 > comparison.\cite{Bedrov:2000, Kuang2010}
210  
211 < Hexane molecules are described by the TraPPE united atom model. This model provides good computational efficiency and reasonable accuracy for bulk thermal conductivity values.
211 > Hexane molecules are described by the TraPPE united atom model,\cite{TraPPE-UA.alkanes} which provides good
212 > computational efficiency and reasonable accuracy for bulk thermal conductivity values. In this model, sites are
213 > located at the carbon centers for alkyl groups. Bonding interactions, including bond stretches and bends and
214 > torsions, were used for intra-molecular sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
215 > potentials were used. We have previously utilized both united atom (UA) and all-atom (AA) force fields for
216 > thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united atom force fields cannot populate the
217 > high-frequency modes that are present in AA force fields, they appear to work better for modeling thermal
218 > conductivity.
219  
220 < Metal-nonmetal interactions are governed by parameters derived from Luedtke and Landman.
220 > Gold -- hexane nonbonded interactions are governed by pairwise Lennard-Jones parameters derived from Vlugt
221 > \emph{et al}.\cite{vlugt:cpc2007154} They fitted parameters for the interaction between Au and CH$_{\emph x}$
222 > pseudo-atoms based on the effective potential of Hautman and Klein for the Au(111) surface.\cite{hautman:4994}
223  
101
224   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225   % NON-PERIODIC DYNAMICS
226   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
227 < \subsection{Dynamics for non-periodic systems}
227 > % \subsection{Dynamics for non-periodic systems}
228 > %
229 > % We have run all tests using the Langevin Hull non-periodic simulation methodology.\cite{Vardeman2011} The
230 > % Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different
231 > % compressibilities, which are typically problematic for traditional affine transform methods. We have had
232 > % success applying this method to several different systems including bare metal nanoparticles, liquid water
233 > % clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal
234 > % compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous
235 > % theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex
236 > % hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled
237 > % to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are
238 > % subject to standard Newtonian dynamics.
239  
240 < 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.
240 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
241 > % SIMULATION PROTOCOL
242 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
243 > \subsection{Simulation protocol}
244  
245 + In all cases, systems were fully equilibrated under non-periodic isobaric-isothermal (NPT) conditions -- using
246 + the Langevin Hull methodology\cite{Vardeman2011} -- before any non-equilibrium methods were introduced. For
247 + heterogeneous systems, the gold nanoparticles and ellipsoid were first created from a bulk lattice and
248 + thermally equilibrated before being solvated in hexane. Packmol\cite{packmol} was used to solvate previously
249 + equilibrated gold nanostructures within a spherical droplet of hexane.
250 +
251 + Once fully equilibrated, a thermal or angular momentum flux was applied for 1 - 2
252 + ns, until a stable temperature or angular velocity gradient had developed. Systems containing liquids were run
253 + 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
254 + imposed flux.
255 +
256 + To stabilize the gold nanoparticle under the imposed angular momentum flux we altered the gold atom at the
257 + designated coordinate origin to have $10,000$ times its original mass. The nonbonded interactions remain
258 + unchanged and the heavy atom is excluded from the angular momentum exchange. For rotation of the ellipsoid about
259 + 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.
260 +
261   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262 < % NON-PERIODIC RNEMD
262 > % THERMAL CONDUCTIVITIES
263   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264 < \subsection{VSS-RNEMD for non-periodic systems}
264 > \subsection{Thermal conductivities}
265  
266 < The adaptation of VSS-RNEMD for non-periodic systems is relatively
267 < 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.
266 > Fourier's Law of heat conduction in radial coordinates yields an expression for the heat flow between the
267 > concentric spherical RNEMD shells:
268  
269 + \begin{equation}
270 +        q_r = - \frac{4 \pi \lambda (T_b - T_a)}{\frac{1}{r_a} - \frac{1}{r_b}}
271 + \label{eq:Q}
272 + \end{equation}
273 +
274 + where $\lambda$ is the thermal conductivity, and $T_{a,b}$ and $r_{a,b}$ are the temperatures and radii of the
275 + two RNEMD regions, respectively.
276 +
277 + A thermal flux is created using VSS-RNEMD moves, and the temperature in each of the radial shells is recorded.
278 + The resulting temperature profiles are analyzed to yield information about the interfacial thermal conductance.
279 + As the simulation progresses, the VSS moves are performed on a regular basis, and the system develops a thermal
280 + or velocity gradient in response to the applied flux. Once a stable thermal gradient has been established
281 + between the two regions, the thermal conductivity, $\lambda$, can be calculated using a linear regression of
282 + the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
283 +
284 + \begin{equation}
285 +        \lambda = \frac{r_a}{r_b} \frac{q_r}{4 \pi \langle \frac{dT}{dr} \rangle}
286 + \label{eq:lambda}
287 + \end{equation}
288 +
289 + The rate of heat transfer between the two RNEMD regions is the amount of transferred kinetic energy over the
290 + length of the simulation, t
291 +
292 + \begin{equation}
293 +        q_r = \frac{KE}{t}
294 + \label{eq:heat}
295 + \end{equation}
296 +
297 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
298 + % INTERFACIAL THERMAL CONDUCTANCE
299 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300 + \subsection{Interfacial thermal conductance}
301 +
302   \begin{figure}
303 <        \center{\includegraphics[width=7in]{figures/VSS}}
304 <        \caption{Schematics of periodic (left) and nonperiodic (right) Velocity Shearing and Scaling RNEMD. A kinetic energy or momentum flux is applied from region B to region A. Thermal gradients are depicted by a color gradient. Linear or angular velocity gradients are shown as arrows.}
305 <        \label{fig:VSS}
303 > \includegraphics[width=\linewidth]{figures/NP20}
304 > \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.}
305 > \label{fig:NP20}
306   \end{figure}
307  
308 < At each time interval, the particle velocities ($\mathbf{v}_i$ and
309 < $\mathbf{v}_j$) in the cold and hot shells ($C$ and $H$) are modified by a
310 < velocity scaling coefficient ($c$ and $h$), an additive linear velocity shearing
311 < term ($\mathbf{a}_c$ and $\mathbf{a}_h$), and an additive angular velocity
312 < shearing term ($\mathbf{b}_c$ and $\mathbf{b}_h$). Here, $\langle
313 < \mathbf{v}_{i,j} \rangle$ and $\langle \mathbf{\omega}_{i,j} \rangle$ are the
314 < average linear and angular velocities for each shell.
308 > For heterogeneous systems such as a solvated nanoparticle, shown in Figure \ref{fig:NP20}, the interfacial
309 > thermal conductance at the surface of the nanoparticle can be determined using an applied kinetic energy flux.
310 > We can treat the temperature in each radial shell as discrete, making the thermal conductance, $G$, of each
311 > shell the inverse of its Kapitza resistance, $R_K$. To model the thermal conductance across an interface (or
312 > multiple interfaces) it is useful to consider the shells as resistors wired in series. The resistance of the
313 > shells is then additive, and the interfacial thermal conductance is the inverse of the total Kapitza
314 > resistance. The thermal resistance of each shell is
315  
316 < \begin{displaymath}
317 < \begin{array}{rclcl}
318 < & \underline{~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~} & &
319 < \underline{\mathrm{rotational \; shearing}} \\  \\
136 < \mathbf{v}_i $~~~$\leftarrow &
137 <  c \, \left(\mathbf{v}_i - \langle \omega_c
138 <  \rangle \times r_i\right) & + & \mathbf{b}_c \times r_i \\
139 < \mathbf{v}_j $~~~$\leftarrow &
140 <  h \, \left(\mathbf{v}_j - \langle \omega_h
141 <  \rangle \times r_j\right) & + & \mathbf{b}_h \times r_j
142 < \end{array}
143 < \end{displaymath}
316 > \begin{equation}
317 >        R_K = \frac{1}{q_r} \Delta T 4 \pi r^2
318 > \label{eq:RK}
319 > \end{equation}
320  
321 < \begin{eqnarray}
146 < \mathbf{b}_c & = & - \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_c^{-1} + \langle \omega_c \rangle \label{eq:bc}\\
147 < \mathbf{b}_h & = & + \mathbf{j}_r(\mathbf{L}) \; \Delta \, t \; I_h^{-1} + \langle \omega_h \rangle \label{eq:bh}
148 < \end{eqnarray}
321 > making the total resistance of two neighboring shells
322  
323 < The total energy is constrained via two quadratic formulae,
323 > \begin{equation}
324 >        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}
325 > \label{eq:Rtotal}
326 > \end{equation}
327  
328 < \begin{eqnarray}
329 < 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}\\
330 < 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}
155 < \end{eqnarray}
328 > This series can be expanded for any number of adjacent shells, allowing for the calculation of the interfacial
329 > thermal conductance for interfaces of considerable thickness, such as self-assembled ligand monolayers on a
330 > metal surface.
331  
332 < the simultaneous
333 < solution of which provide the velocity scaling coefficients $c$ and $h$. Given an
334 < imposed angular momentum flux, $\mathbf{j}_{r} \left( \mathbf{L} \right)$, and/or
335 < thermal flux, $J_r$, equations \ref{eq:bc} - \ref{eq:Kh} are sufficient to obtain
161 < the velocity scaling ($c$ and $h$) and shearing ($\mathbf{b}_c,\,$ and $\mathbf{b}_h$) at each time interval.
332 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
333 > % INTERFACIAL ROTATIONAL FRICTION
334 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335 > \subsection{Interfacial rotational friction}
336  
337 + The interfacial rotational friction, $\Xi^{rr}$, can be calculated for heterogeneous nanostructure/solvent
338 + systems by applying an angular momentum flux between the solvated nanostructure and a spherical shell of
339 + solvent at the boundary of the cluster. An angular velocity gradient develops in response to the applied flux,
340 + causing the nanostructure and solvent shell to rotate in opposite directions about a given axis.
341 +
342 + \begin{figure}
343 + \includegraphics[width=\linewidth]{figures/E25-75}
344 + \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.}
345 + \label{fig:E25-75}
346 + \end{figure}
347 +
348 + 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
349 +
350 + \begin{equation}
351 +        \Xi^{rr}_{stick} = 8 \pi \eta r^3
352 + \label{eq:Xisphere}.
353 + \end{equation}
354 +
355 + where $\eta$ is the dynamic viscosity of the surrounding solvent. An $\eta$ value for TraPPE-UA hexane under
356 + these particular temperature and pressure conditions was determined by applying a traditional VSS-RNEMD linear
357 + momentum flux to a periodic box of solvent.
358 +
359 + For general ellipsoids with semiaxes $a$, $b$, and $c$, Perrin's extension of Stokes' law provides exact
360 + solutions for symmetric prolate $(a \geq b = c)$ and oblate $(a < b = c)$ ellipsoids under ideal ``stick'' conditions. For simplicity, we define
361 + a Perrin Factor, $S$,
362 +
363 + \begin{equation}
364 +        S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right].
365 + \label{eq:S}
366 + \end{equation}
367 +
368 + For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
369 + \begin{equation}
370 +        \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S}
371 + \label{eq:Xia}
372 + \end{equation}\vspace{-0.45in}\\
373 + \begin{equation}
374 +        \Xi^{rr}_{b,c} = \frac{32 \pi}{2} \eta \frac{ \left( a^4 - b^4 \right)}{ \left( 2a^2 - b^2 \right)S - 2a}.
375 + \label{eq:Xibc}
376 + \end{equation}
377 +
378 + corresponding to rotation about the long axis ($a$), and each of the equivalent short axes ($b$ and $c$), respectively.
379 +
380 + Previous VSS-RNEMD simulations of the interfacial friction of the planar Au(111) / hexane interface have shown
381 + that the interface exists within ``slip'' boundary conditions.\cite{Kuang2012} Hu and Zwanzig\cite{Zwanzig}
382 + investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained
383 + numerial results for a scaling factor to be applied to $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau$, the
384 + ratio of the shorter semiaxes and the longer semiaxis of the spheroid. For the sphere and prolate ellipsoid
385 + shown here, the values of $\tau$ are $1$ and $0.3939$, respectively. Under ``slip'' conditions,
386 + $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate ellipsoid about its long axis approaches
387 + $0$, as no solvent is displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for rotation of the
388 + prolate ellipsoid about its short axis is $35.9\%$ of the analytical $\Xi^{rr}_{\mathit{stick}}$ result,
389 + accounting for the reduced interfacial friction under ``slip'' boundary conditions.
390 +
391 + 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 ($\omega_{Au}$)
392 +
393 + \begin{equation}
394 +        \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
395 + \label{eq:Xieff}
396 + \end{equation}
397 +
398 + The applied torque required to overcome the interfacial friction and maintain constant rotation of the gold is
399 +
400 + \begin{equation}
401 +        \tau = \frac{L}{2 t}
402 + \label{eq:tau}  
403 + \end{equation}
404 +
405 + where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
406 +
407   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
408   % **TESTS AND APPLICATIONS**
409   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 170 | Line 414 | Calculated values for the thermal conductivity of a 40
414   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
415   \subsection{Thermal conductivities}
416  
417 < 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.
417 > Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at
418 > different kinetic energy flux values are shown in Table \ref{table:goldTC}. For these calculations, the hot and
419 > cold slabs were excluded from the linear regression of the thermal gradient.
420  
421   \begin{longtable}{ccc}
422   \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.}
423   \\ \hline \hline
424 < {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
425 < {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
426 < 3.25$\times 10^{-6}$ & 0.11435 & 1.9753 \\
427 < 6.50$\times 10^{-6}$ & 0.2324 & 1.9438 \\
428 < 1.30$\times 10^{-5}$ & 0.44922 & 2.0113 \\
429 < 3.25$\times 10^{-5}$ & 1.1802 & 1.9139 \\
430 < 6.50$\times 10^{-5}$ & 2.339 & 1.9314
424 > {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
425 > {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
426 > 3.25$\times 10^{-6}$ & 0.11435 & 1.0143 \\
427 > 6.50$\times 10^{-6}$ & 0.2324 & 0.9982 \\
428 > 1.30$\times 10^{-5}$ & 0.44922 & 1.0328 \\
429 > 3.25$\times 10^{-5}$ & 1.1802 & 0.9828 \\
430 > 6.50$\times 10^{-5}$ & 2.339 & 0.9918 \\
431 > \hline
432 > This work & & 1.0040
433   \\ \hline \hline
434 < \label{table:goldconductivity}
434 > \label{table:goldTC}
435   \end{longtable}
188        
189 SPC/E Water Cluster:
436  
437 + The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly dependent on the applied kinetic energy
438 + 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
439 + of 320 W / m $\cdot$ K, as the QSC force field neglects significant electronic contributions to
440 + heat conduction.
441 +
442 + Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table
443 + \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were
444 + excluded from the $\langle \frac{dT}{dr} \rangle$ fit.
445 +
446   \begin{longtable}{ccc}
447 < \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and ambient density.}
447 > \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
448   \\ \hline \hline
449 < {$J_r$} & {$\langle dT / dr \rangle$} & {$\boldsymbol \lambda$}\\
450 < {\small(kcal fs$^{-1}$ \AA$^{-2}$)} & {\small(K \AA$^{-1}$)} & {\small(W m$^{-1}$ K$^{-1}$)}\\ \hline
449 > {$J_r$} & {$\langle \frac{dT}{dr} \rangle$} & {$\boldsymbol \lambda$}\\
450 > {\footnotesize(kcal / fs $\cdot$ \AA$^{2}$)} & {\footnotesize(K / \AA)} & {\footnotesize(W / m $\cdot$ K)}\\ \hline
451 > 1$\times 10^{-5}$ & 0.38683 & 0.8698 \\
452 > 3$\times 10^{-5}$ & 1.1643 & 0.9098 \\
453 > 6$\times 10^{-5}$ & 2.2262 & 0.8727 \\
454 > \hline
455 > This work & & 0.8841 \\
456 > Zhang, et al\cite{Zhang2005} & & 0.81 \\
457 > R$\ddot{\mathrm{o}}$mer, et al\cite{Romer2012} & & 0.87 \\
458 > Experiment\cite{WagnerKruse} & & 0.61
459   \\ \hline \hline
460 < \label{table:waterconductivity}
460 > \label{table:waterTC}
461   \end{longtable}
462  
463 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
464 < % SHEAR VISCOSITY
465 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
466 < \subsection{Shear viscosity}
463 > Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average
464 > calculated thermal conductivity from this work, $0.8841$ W / m $\cdot$ K, compares very well to
465 > previous non-equilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental
466 > values.\cite{WagnerKruse}
467  
205 SPC/E Water Cluster:
206
468   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
469   % INTERFACIAL THERMAL CONDUCTANCE
470   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
471   \subsection{Interfacial thermal conductance}
472  
473 < The interfacial thermal conductance, $G$, is calculated by defining a temperature difference $\Delta T$ across a given interface.
473 > Calculated interfacial thermal conductance ($G$) values for three sizes of gold nanoparticles and Au(111)
474 > surface solvated in TraPPE-UA hexane are shown in Table \ref{table:G}.
475  
476 + \begin{longtable}{ccc}
477 + \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.}
478 + \\ \hline \hline
479 + {Nanoparticle Radius} & {$G$}\\
480 + {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
481 + 20 & {47.1} \\
482 + 30 & {45.4} \\
483 + 40 & {46.5} \\
484 + \hline
485 + Au(111) & {30.2}
486 + \\ \hline \hline
487 + \label{table:G}
488 + \end{longtable}
489 +
490 + The introduction of surface curvature increases the interfacial thermal conductance by a factor of
491 + approximately $1.5$ relative to the flat interface. There are no significant differences in the $G$ values for
492 + the varying nanoparticle sizes. It seems likely that for the range of nanoparticle sizes represented here, any
493 + 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.
494 +
495   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
496   % INTERFACIAL FRICTION
497   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
498   \subsection{Interfacial friction}
499  
500 < 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.
500 > Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold
501 > nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied
502 > between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius,
503 > respectively.
504  
221 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.
222
223 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
224
225 \begin{equation}
226        \Xi = 8 \pi \eta r^3 \label{eq:Xi}.
227 \end{equation}
228
229 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.
230
231 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, $S$,
232
233 \begin{equation}
234        S = \frac{2}{\sqrt{a^2 - b^2}} ln \left[ \frac{a + \sqrt{a^2 - b^2}}{b} \right]. \label{eq:S}
235 \end{equation}
236
237 For a prolate ellipsoidal rod, demonstrated here, the rotational resistance tensor $\Xi$ is a $3 \times 3$ diagonal matrix with elements
238
239 \begin{eqnarray}
240        \Xi^{rr}_a = \frac{32 \pi}{2} \eta \frac{ \left( a^2 - b^2 \right) b^2}{2a - b^2 S} \label{eq:Xia}\\
241        \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}
242 \end{eqnarray}
243
244 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.
245
246 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.
247        
505   \begin{longtable}{lccccc}
506 < \caption{Calculated ``stick'' interfacial friction coefficients ($\kappa$) and friction factors ($f$) of gold nanostructures solvated in TraPPE-UA hexane. The ellipsoid is oriented with the long axis along the $z$ direction.}
506 > \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.}
507   \\ \hline \hline
508 < {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$} & {$f_{VSS}$} & {$f_{calc}$}\\
509 < {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)} & {} & {}\\  \hline
510 < {Sphere} & {$x = y = z$} & {} & {5.37237} & {1} & {1}\\
511 < {Prolate Ellipsoid} & {$x = y$} & {} & {3.59881} & {} & {0.768726}\\
512 < {Prolate Ellipsoid} & {$z$} & {} & {9.01084} & {} & {1.92477}\\  \hline \hline
513 < \label{table:interfacialfriction}
514 < \end{longtable}
258 <
259 < \begin{longtable}{lccc}
260 < \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.}
508 > {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
509 > {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\  \hline
510 > Sphere (r = 20 \AA) & {$x = y = z$} & 0 & {2386} & {3314} & {0.720}\\
511 > Sphere (r = 30 \AA) & {$x = y = z$} & 0 & {8415} & {11749} & {0.716}\\
512 > Sphere (r = 40 \AA) & {$x = y = z$} & 0 & {47544} & {34464} & {1.380}\\
513 > Prolate Ellipsoid & {$x = y$} & {1792} & {3128} & {4991} & {0.627}\\
514 > Prolate Ellipsoid & {$z$} & 0 & {1590} & {1993} & {0.798}
515   \\ \hline \hline
516 < {Structure} & {Axis of rotation} & {$\kappa_{VSS}$} & {$\kappa_{calc}$}\\
263 < {} & {} & {\small($10^{-29}$ Pa s m$^{3}$)} & {\small($10^{-29}$ Pa s m$^{3}$)}\\  \hline
264 < {Sphere} & {$x = y = z$} & {} & {0}\\
265 < {Prolate Ellipsoid} & {$x = y$} & {} & {0}\\
266 < {Prolate Ellipsoid} & {$z$} & {} & {7.9295392}\\  \hline \hline
267 < \label{table:interfacialfriction}
516 > \label{table:couple}
517   \end{longtable}
518  
519 + The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold
520 + structures solvated by hexane do not exist in the ``slip'' boundary condition. At this length scale, the
521 + nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience
522 + increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius
523 + nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of
524 + the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short
525 + axis, which at first seems counterintuitive. However, the `propellor' motion caused by rotation about the
526 + short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the
527 + gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius)
528 + appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of
529 + surface features or anomalous solvent behaviors that are not fully understood at this time.
530 +
531   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
532   % **DISCUSSION**
533   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
534   \section{Discussion}
535  
536 + 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.
537  
538 + 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.
539 +
540   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541   % **ACKNOWLEDGMENTS**
542   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
543   \section*{Acknowledgments}
544  
545 < We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
546 < this project was provided by the National Science Foundation under grant
283 < CHE-0848243. Computational time was provided by the Center for Research
545 > We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for this project was provided by the
546 > National Science Foundation under grant CHE-0848243. Computational time was provided by the Center for Research
547   Computing (CRC) at the University of Notre Dame.
548  
549   \newpage

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines