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 4004 by kstocke1, Fri Jan 17 22:44:07 2014 UTC vs.
Revision 4009 by kstocke1, Thu Jan 23 23:26:39 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 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 | Non-equilibrium Molecular Dynamics (NEMD) methods impo
82   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83   \section{Introduction}
84  
85 < Non-equilibrium Molecular Dynamics (NEMD) methods impose a temperature
86 < or velocity {\it gradient} on a
87 < system,\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
88 < and use linear response theory to connect the resulting thermal or
89 < momentum flux to transport coefficients of bulk materials.  However,
90 < for heterogeneous systems, such as phase boundaries or interfaces, it
88 < is often unclear what shape of gradient should be imposed at the
89 < boundary between materials.
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{Velocity Shearing and Scaling (VSS) for non-periodic systems}
109 +
110 + The 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)
# Line 98 | Line 122 | Reverse Non-Equilibrium Molecular Dynamics (RNEMD) met
122   \label{fig:VSS}
123   \end{figure}
124  
125 < Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods impose an
126 < unphysical {\it flux} between different regions or ``slabs'' of the
127 < simulation
128 < box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
129 < system responds by developing a temperature or velocity {\it gradient}
106 < between the two regions.  The gradients which develop in response to
107 < the applied flux are then related (via linear response theory) to the
108 < transport coefficient of interest. Since the amount of the applied
109 < flux is known exactly, and measurement of a gradient is generally less
110 < complicated, imposed-flux methods typically take shorter simulation
111 < times to obtain converged results. At interfaces, the observed
112 < gradients often exhibit near-discontinuities at the boundaries between
113 < dissimilar materials.  RNEMD methods do not need many trajectories to
114 < provide information about transport properties, and they have become
115 < widely used to compute thermal and mechanical transport in both
116 < homogeneous liquids and
117 < solids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
118 < well as heterogeneous
119 < interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
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 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
132 < % **METHODOLOGY**
133 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124 < \section{Velocity Shearing and Scaling (VSS) for non-periodic systems}
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 < The VSS-RNEMD approach uses a series of simultaneous velocity shearing
136 < and scaling exchanges between the two slabs.\cite{Kuang2012}
137 < This method imposes energy and momentum conservation constraints while
129 < simultaneously creating a desired flux between the two slabs.  These
130 < constraints ensure that all configurations are sampled from the same
131 < microcanonical (NVE) ensemble.
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  
133 We have extended the VSS method for use in {\it non-periodic}
134 simulations, in which the ``slabs'' have been generalized to two
135 separated regions of space.  These regions could be defined as
136 concentric spheres (as in figure \ref{fig:VSS}), or one of the regions
137 can be defined in terms of a dynamically changing ``hull'' comprising
138 the surface atoms of the cluster.  This latter definition is identical
139 to the hull used in the Langevin Hull algorithm.
140
141 We present here a new set of constraints that are more general than
142 the VSS constraints.  For the non-periodic variant, the constraints
143 fix both the total energy and total {\it angular} momentum of the
144 system while simultaneously imposing a thermal and angular momentum
145 flux between the two regions.
146
147 After each $\Delta t$ time interval, the particle velocities
148 ($\mathbf{v}_i$ and $\mathbf{v}_j$) in the two shells ($A$ and $B$)
149 are modified by a velocity scaling coefficient ($a$ and $b$) and by a
150 rotational shearing term ($\mathbf{c}_a$ and $\mathbf{c}_b$).
139   \begin{displaymath}
140   \begin{array}{rclcl}
141   & \underline{~~~~~~~~\mathrm{scaling}~~~~~~~~} & &
# Line 160 | Line 148 | Here $\langle\mathbf{\omega}_a\rangle$ and
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
152 < $\langle\mathbf{\omega}_b\rangle$ are the instantaneous angular
153 < velocities of each shell, and $\mathbf{r}_i$ is the position of
154 < particle $i$ relative to a fixed point in space (usually the center of
167 < mass of the cluster).  Particles in the shells also receive an
168 < additive ``angular shear'' to their velocities.  The amount of shear
169 < is governed by the imposed angular momentum flux,
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
# Line 174 | Line 159 | where $\overleftrightarrow{I}_{\{a,b\}}$ is the moment
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
178 < each of the two shells.
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
181 < use energy conservation constraints,
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
# Line 188 | Line 171 | Simultaneous solution of these quadratic formulae for
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
175 < coefficients, $a$ and $b$, will ensure that the simulation samples
176 < from the original microcanonical (NVE) ensemble.  Here $K_{\{a,b\}}$
177 < is the instantaneous translational kinetic energy of each shell.  At
178 < each time interval, we solve for $a$, $b$, $\mathbf{c}_a$, and
179 < $\mathbf{c}_b$, subject to the imposed angular momentum flux,
197 < $j_r(\mathbf{L})$, and thermal flux, $J_r$, values.  The new particle
198 < velocities are computed, and the simulation continues. System
199 < configurations after the transformations have exactly the same energy
200 < ({\it and} angular momentum) as before the moves.
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
182 < performed on a regular basis, and the system will develop a
183 < temperature and/or angular velocity gradient in response to the
184 < applied flux.  Using the slope of the radial temperature or velocity
185 < gradients, it is quite simple to obtain both the thermal conductivity
207 < ($\lambda$), interfacial thermal conductance ($G$), or rotational friction coefficients ($\Xi^{rr}$) of any nonperiodic system.
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 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
193 < % NON-PERIODIC DYNAMICS
194 < %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
195 < \subsection{Dynamics for non-periodic systems}
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  
219 We have run all tests using the Langevin Hull nonperiodic simulation methodology.\cite{Vardeman2011} The Langevin Hull is especially useful for simulating heterogeneous mixtures of materials with different compressibilities, which are typically problematic for traditional affine transform methods. We have had success applying this method to
220 several different systems including bare metal nanoparticles, liquid water clusters, and explicitly solvated nanoparticles. Calculated physical properties such as the isothermal compressibility of water and the bulk modulus of gold nanoparticles are in good agreement with previous theoretical and experimental results. The Langevin Hull uses a Delaunay tesselation to create a dynamic convex hull composed of triangular facets with vertices at atomic sites. Atomic sites included in the hull are coupled to an external bath defined by a temperature, pressure and viscosity. Atoms not included in the hull are subject to standard Newtonian dynamics.
221
198   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
223 % SIMULATION PROTOCOL
224 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225 \subsection{Simulation protocol}
226
227 Systems containing liquids were run under moderate pressure ($\sim$ 5 atm) to avoid the formation of a substantial vapor phase. Thermal coupling to the Langevin Hull external bath was turned off to avoid interference with any imposed flux.
228
229 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
199   % FORCE FIELD PARAMETERS
200   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
201   \subsection{Force field parameters}
202  
203 < We have chosen the SPC/E water model for these simulations, which is particularly useful for method validation as there are many values for physical properties from previous simulations available for direct comparison.\cite{Bedrov:2000, Kuang2010}
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.\cite{PhysRevB.59.3527} The QSC parameters are tuned to experimental properties such as density, cohesive energy, and elastic moduli and include zero-point quantum corrections.
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,\cite{TraPPE-UA.alkanes} which provides good computational efficiency and reasonable accuracy for bulk thermal conductivity values. In this model,
212 < sites are located at the carbon centers for alkyl groups. Bonding
213 < interactions, including bond stretches and bends and torsions, were
214 < used for intra-molecular sites closer than 3 bonds. For non-bonded
215 < interactions, Lennard-Jones potentials were used.  We have previously
216 < utilized both united atom (UA) and all-atom (AA) force fields for
217 < thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
218 < atom force fields cannot populate the high-frequency modes that are
246 < present in AA force fields, they appear to work better for modeling
247 < thermal conductivity.
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 < Gold -- hexane nonbonded interactions are governed by pairwise Lennard-Jones parameters derived from Vlugt \emph{et al}.\cite{vlugt:cpc2007154} They fitted parameters for the interaction between Au and CH$_{\emph x}$ pseudo-atoms based on the effective potential of Hautman and Klein for the Au(111) surface.\cite{hautman:4994}
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  
224   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225 + % NON-PERIODIC DYNAMICS
226 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
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 + %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
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 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. 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. 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   % THERMAL CONDUCTIVITIES
263   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
264   \subsection{Thermal conductivities}
265  
266 < Fourier's Law of heat conduction in radial coordinates yields an expression for the heat flow between the concentric spherical RNEMD 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 two RNEMD regions, respectively.
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 < Once a stable thermal gradient has been established between the two regions, the thermal conductivity, $\lambda$, can be calculated using a linear regression of the thermal gradient, $\langle \frac{dT}{dr} \rangle$:
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 length of the simulation, t
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}
# Line 281 | Line 299 | A thermal flux is created using VSS-RNEMD moves, and t
299   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300   \subsection{Interfacial thermal conductance}
301  
302 < A thermal flux is created using VSS-RNEMD moves, and the temperature in each of the radial shells is recorded. The resulting temperature
303 < profiles are analyzed to yield information about the interfacial thermal
304 < conductance. As the simulation progresses, the VSS moves are performed on a regular basis, and
305 < the system develops a thermal or velocity gradient in response to the applied
306 < flux. We can treat the temperature in each radial shell as discrete, making the thermal conductance, $G$, of each shell the inverse of its Kapitza resistance, $R_K$. To model the thermal conductance across an interface (or multiple interfaces) it is useful to consider the shells as resistors wired in series. The resistance of the shells is then additive, and the interfacial thermal conductance is the inverse of the total Kapitza resistance. The thermal resistance of each shell is
302 > \begin{figure}
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 + 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{equation}
317          R_K = \frac{1}{q_r} \Delta T 4 \pi r^2
318   \label{eq:RK}
# Line 299 | Line 325 | This series can be expanded for any number of adjacent
325   \label{eq:Rtotal}
326   \end{equation}
327  
328 < This series can be expanded for any number of adjacent shells, allowing for the calculation of the interfacial thermal conductance for interfaces of significant thickness, such as self-assembled ligand monolayers on a metal surface.
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 significant thickness, such as self-assembled ligand monolayers on a
330 > metal surface.
331  
332   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
333 < % INTERFACIAL FRICTION
333 > % INTERFACIAL ROTATIONAL FRICTION
334   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335 < \subsection{Interfacial friction}
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} = 8 \pi \eta r^3
352 < \label{eq:Xistick}.
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 UA hexane under these particular temperature and pressure conditions was determined by applying a traditional VSS-RNEMD linear momentum flux to a periodic box of solvent.
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 solutions for symmetric prolate $(a \geq b = c)$ and oblate $(a < b = c)$ ellipsoids. For simplicity, we define a Perrin Factor, $S$,
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. 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].
# Line 332 | Line 375 | The effective rotational friction coefficient at the i
375   \label{eq:Xibc}
376   \end{equation}
377  
378 < The effective rotational friction coefficient at the interface can be extracted from nonperiodic VSS-RNEMD simulations quite easily using the applied torque ($\tau$) and the observed angular velocity of the gold structure ($\omega_{Au}$)
378 > The effective rotational friction coefficient 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 ($\omega_{Au}$)
379  
380   \begin{equation}
381          \Xi^{rr}_{\mathit{eff}} = \frac{\tau}{\omega_{Au}}
# Line 348 | Line 391 | Previous VSS-RNEMD simulations of the interfacial fric
391  
392   where $L$ is the total angular momentum exchanged between the two RNEMD regions and $t$ is the length of the simulation.
393  
394 < Previous VSS-RNEMD simulations of the interfacial friction of the planar Au(111) / hexane interface have shown that the interface exists within ``slip'' boundary conditions.\cite{Kuang2012} Hu and Zwanzig\cite{Zwanzig} investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained numerial results for a scaling factor to be applied to $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau$, the ratio of the shorter semiaxes and the longer semiaxis of the spheroid. For the sphere and prolate ellipsoid shown here, the values of $\tau$ are $1$ and $0.3939$, respectively. Under ``slip'' conditions, $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate ellipsoid about its long axis approaches $0$, as no solvent is displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for rotation of the prolate ellipsoid about its short axis is $35.9\%$ of the analytical $\Xi^{rr}_{\mathit{stick}}$ result, accounting for the reduced interfacial friction under ``slip'' boundary conditions.
394 > Previous VSS-RNEMD simulations of the interfacial friction of the planar Au(111) / hexane interface have shown
395 > that the interface exists within ``slip'' boundary conditions.\cite{Kuang2012} Hu and Zwanzig\cite{Zwanzig}
396 > investigated the rotational friction coefficients for spheroids under slip boundary conditions and obtained
397 > numerial results for a scaling factor to be applied to $\Xi^{rr}_{\mathit{stick}}$ as a function of $\tau$, the
398 > ratio of the shorter semiaxes and the longer semiaxis of the spheroid. For the sphere and prolate ellipsoid
399 > shown here, the values of $\tau$ are $1$ and $0.3939$, respectively. Under ``slip'' conditions,
400 > $\Xi^{rr}_{\mathit{slip}}$ for any sphere and rotation of the prolate ellipsoid about its long axis approaches
401 > $0$, as no solvent is displaced by either of these rotations. $\Xi^{rr}_{\mathit{slip}}$ for rotation of the
402 > prolate ellipsoid about its short axis is $35.9\%$ of the analytical $\Xi^{rr}_{\mathit{stick}}$ result,
403 > accounting for the reduced interfacial friction under ``slip'' boundary conditions.
404  
405   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
406   % **TESTS AND APPLICATIONS**
# Line 360 | Line 412 | Calculated values for the thermal conductivity of a 40
412   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
413   \subsection{Thermal conductivities}
414  
415 < Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at different kinetic energy flux values are shown in Table \ref{table:goldTC}. For these calculations, the hot and cold slabs were excluded from the linear regression of the thermal gradient. The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly dependent on the applied kinetic energy flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 -- 1.26 {\footnotesize W / m $\cdot$ K}\cite{Kuang2010}, though still significantly lower than the experimental value of 320 {\footnotesize W / m $\cdot$ K}, as the QSC force field neglects significant electronic contributions to heat conduction.
415 > Calculated values for the thermal conductivity of a 40 \AA$~$ radius gold nanoparticle (15707 atoms) at
416 > different kinetic energy flux values are shown in Table \ref{table:goldTC}. For these calculations, the hot and
417 > cold slabs were excluded from the linear regression of the thermal gradient.
418  
365 % 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.
366
419   \begin{longtable}{ccc}
420   \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.}
421   \\ \hline \hline
# Line 380 | Line 432 | Calculated values for the thermal conductivity of a cl
432   \label{table:goldTC}
433   \end{longtable}
434  
435 < Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were excluded from the $\langle \frac{dT}{dr} \rangle$ fit. Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average calculated thermal conductivity from this work, $0.8841$ {\footnotesize W / m $\cdot$ K}, compares very well to previous nonequilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental values.\cite{WagnerKruse}
435 > The measured linear slope $\langle \frac{dT}{dr} \rangle$ is linearly dependent on the applied kinetic energy
436 > flux $J_r$. Calculated thermal conductivity values compare well with previous bulk QSC values of 1.08 -- 1.26
437 > {\footnotesize W / m $\cdot$ K}\cite{Kuang2010}, though still significantly lower than the experimental value
438 > of 320 {\footnotesize W / m $\cdot$ K}, as the QSC force field neglects significant electronic contributions to
439 > heat conduction.
440  
441 + Calculated values for the thermal conductivity of a cluster of 6912 SPC/E water molecules are shown in Table
442 + \ref{table:waterTC}. As with the gold nanoparticle thermal conductivity calculations, the RNEMD regions were
443 + excluded from the $\langle \frac{dT}{dr} \rangle$ fit.
444 +
445   \begin{longtable}{ccc}
446   \caption{Calculated thermal conductivity of a cluster of 6912 SPC/E water molecules. Calculations were performed at 300 K and 5 atm.}
447   \\ \hline \hline
# Line 399 | Line 459 | Experiment\cite{WagnerKruse} & & 0.61
459   \label{table:waterTC}
460   \end{longtable}
461  
462 + Again, the measured slope is linearly dependent on the applied kinetic energy flux $J_r$. The average
463 + calculated thermal conductivity from this work, $0.8841$ {\footnotesize W / m $\cdot$ K}, compares very well to
464 + previous non-equilibrium molecular dynamics results\cite{Romer2012, Zhang2005} and experimental
465 + values.\cite{WagnerKruse}
466 +
467   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468   % INTERFACIAL THERMAL CONDUCTANCE
469   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
470   \subsection{Interfacial thermal conductance}
471  
472 < Calculated interfacial thermal conductance ($G$) values for three sizes of gold nanoparticles and Au(111) surface solvated in TraPPE-UA hexane. The introduction of surface curvature increases the interfacial thermal conductance by a factor of approximately $1.5$ relative to the flat interface. There are no significant differences in the $G$ values for the varying nanoparticle sizes. It seems likely that for the range of nanoparticle sizes represented here, size effects are not evident.
472 > Calculated interfacial thermal conductance ($G$) values for three sizes of gold nanoparticles and Au(111)
473 > surface solvated in TraPPE-UA hexane are shown in Table \ref{table:G}.
474  
475   \begin{longtable}{ccc}
476 < \caption{Calculated interfacial thermal conductance ($G$) values for gold nanoparticles of varying radii solvated in explicit TraPPE-UA hexane. The nanoparticle $G$ values are compared to previous results for a Au(111) interface in TraPPE-UA hexane, revealing increased interfacial thermal conductance for non-planar interfaces.}
476 > \caption{Calculated interfacial thermal conductance ($G$) values for gold nanoparticles of varying radii solvated in explicit TraPPE-UA hexane. The nanoparticle $G$ values are compared to previous results for a Au(111) interface in TraPPE-UA hexane.}
477   \\ \hline \hline
478   {Nanoparticle Radius} & {$G$}\\
479   {\footnotesize(\AA)} & {\footnotesize(MW / m$^{2}$ $\cdot$ K)}\\ \hline
# Line 417 | Line 483 | Au(111) & {30.2}
483   \hline
484   Au(111) & {30.2}
485   \\ \hline \hline
486 < \label{table:interfacialconductance}
486 > \label{table:G}
487   \end{longtable}
488  
489 + The introduction of surface curvature increases the interfacial thermal conductance by a factor of
490 + approximately $1.5$ relative to the flat interface. There are no significant differences in the $G$ values for
491 + the varying nanoparticle sizes. It seems likely that for the range of nanoparticle sizes represented here, any
492 + particle size effects are not evident.
493 +
494   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
495   % INTERFACIAL FRICTION
496   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
497   \subsection{Interfacial friction}
498  
499 < Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius, respectively. The resulting angular velocity gradient causes the gold structure to rotate about the prescribed axis.
499 > Table \ref{table:couple} shows the calculated rotational friction coefficients $\Xi^{rr}$ for spherical gold
500 > nanoparticles and a prolate ellipsoidal gold nanorod in TraPPE-UA hexane. An angular momentum flux was applied
501 > between the A and B regions defined as the gold structure and hexane molecules beyond a certain radius,
502 > respectively.
503  
504   \begin{longtable}{lccccc}
505 < \caption{Comparison of rotational friction coefficients under ideal ``slip'' ($\Xi^{rr}_{\mathit{slip}}$) and ``stick'' conditions ($\Xi^{rr}_{\mathit{stick}}$) and effective rotational friction coefficients ($\Xi^{rr}_{\mathit{eff}}$) of gold nanostructures solvated in TraPPE-UA hexane at 230 K. The ellipsoid is oriented with the long axis along the $z$ direction.}
505 > \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.}
506   \\ \hline \hline
507   {Structure} & {Axis of Rotation} & {$\Xi^{rr}_{\mathit{slip}}$} & {$\Xi^{rr}_{\mathit{eff}}$} & {$\Xi^{rr}_{\mathit{stick}}$} & {$\Xi^{rr}_{\mathit{eff}}$ / $\Xi^{rr}_{\mathit{stick}}$}\\
508   {} & {} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & {\footnotesize(amu A$^2$ / fs)} & \\  \hline
# Line 441 | Line 515 | The results for $\Xi^{rr}_{\mathit{eff}}$ show that, c
515   \label{table:couple}
516   \end{longtable}
517  
518 < The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold structures solvated by hexane do not exist in the ``slip'' boundary conditions. At this length scale, the nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short axis, which may at first seem counterintuitive. However, the `propellor' motion caused by rotation about short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius) appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of surface features or anomalous solvent behaviors that are not fully understood at this time.
518 > The results for $\Xi^{rr}_{\mathit{eff}}$ show that, contrary to the flat Au(111) / hexane interface, gold
519 > structures solvated by hexane do not exist in the ``slip'' boundary conditions. At this length scale, the
520 > nanostructures are not perfect spheroids due to atomic `roughening' of the surface and therefore experience
521 > increased interfacial friction which deviates from the ideal ``slip'' case. The 20 and 30 \AA$\,$ radius
522 > nanoparticles experience approximately 70\% of the ideal ``stick'' boundary interfacial friction. Rotation of
523 > the ellipsoid about its long axis more closely approaches the ``stick'' limit than rotation about the short
524 > axis, which may at first seem counterintuitive. However, the `propellor' motion caused by rotation about the
525 > short axis may exclude solvent from the rotation cavity or move a sufficient amount of solvent along with the
526 > gold that a smaller interfacial friction is actually experienced. The largest nanoparticle (40 \AA$\,$ radius)
527 > appears to experience more than the ``stick'' limit of interfacial friction, which may be a consequence of
528 > surface features or anomalous solvent behaviors that are not fully understood at this time.
529  
530   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
531   % **DISCUSSION**
532   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
533   \section{Discussion}
534  
535 + 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.
536  
537 + 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 investigated in the future.
538 +
539   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540   % **ACKNOWLEDGMENTS**
541   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
542   \section*{Acknowledgments}
543  
544 < We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
545 < this project was provided by the National Science Foundation under grant
459 < CHE-0848243. Computational time was provided by the Center for Research
544 > We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for this project was provided by the
545 > National Science Foundation under grant CHE-0848243. Computational time was provided by the Center for Research
546   Computing (CRC) at the University of Notre Dame.
547  
548   \newpage

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines