ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
(Generate patch)

Comparing trunk/chainLength/GoldThiolsPaper.tex (file contents):
Revision 3802 by kstocke1, Wed Dec 5 16:34:23 2012 UTC vs.
Revision 3803 by gezelter, Wed Dec 5 21:49:28 2012 UTC

# Line 22 | Line 22
22  
23   \begin{document}
24  
25 < \title{Interfacial Thermal Conductance of Thiolate-Capped Gold}
25 > \title{The role chain length and solvent penetration in the
26 >  interfacial thermal conductance of thiolate-capped gold surfaces}
27  
28 < \author{Kelsey M. Stocker and J. Daniel Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
28 > \author{Kelsey M. Stocker and J. Daniel
29 >  Gezelter\footnote{Corresponding author. \ Electronic mail:
30 >    gezelter@nd.edu} \\
31 >  251 Nieuwland Science Hall, \\
32          Department of Chemistry and Biochemistry,\\
33          University of Notre Dame\\
34          Notre Dame, Indiana 46556}
# Line 48 | Line 52
52   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53   \section{Introduction}
54  
55 + The structural and dynamical details of interfaces between metal
56 + nanoparticles and solvents determines how energy flows between these
57 + particles and their surroundings. Understanding this energy flow is
58 + essential in designing and functionalizing metallic nanoparticles for
59 + plasmonic photothermal
60 + therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff
61 +  :2007ye,Larson:2007hw} which rely on the ability of metallic
62 + nanoparticles to absorb light in the near-IR, a portion of the
63 + spectrum in which living tissue is very nearly transparent.  The
64 + principle of this therapy is to pump the particles at high power at
65 + the plasmon resonance and to allow heat dissipation to kill targeted
66 + (e.g. cancerous) cells.  The relevant physical property controlling
67 + this transfer of energy is the interfacial thermal conductance, $G$,
68 + which can be somewhat difficult to determine
69 + experimentally.\cite{Wilson:2002uq,Plech:2005kx}
70 +
71 + Metallic particles have also been proposed for use in highly efficient
72 + thermal-transfer fluids, although careful experiments by Eapen {\it et al.}
73 + have shown that metal-particle-based ``nanofluids'' have thermal
74 + conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
75 + likely cause of previously reported non-Maxwell
76 + behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
77 + is percolation networks of nanoparticles exchanging energy via the
78 + solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
79 + picture of particle-solvent interactions in order to understand the
80 + thermal behavior of complex fluids. To date, there have been few
81 + reported values (either from theory or experiment) for $G$ for
82 + ligand-protected nanoparticles embedded in liquids, and there is a
83 + significant gap in knowledge about how chemically distinct ligands or
84 + protecting groups will affect heat transport from the particles.
85 +
86 + The thermal properties of various nanostructured interfaces have been
87 + investigated experimentally by a number of groups: Cahill and
88 + coworkers studied nanoscale thermal transport from metal
89 + nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxides
90 + interfaces, and hydrophilic and hydrophobic interfaces between water
91 + and solids with different self-assembled
92 + monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevL
93 + ett.96.186101}
94 + Wang {\it et al.} studied heat transport through long-chain
95 + hydrocarbon monolayers on gold substrate at the individual molecular
96 + level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
97 + cetyltrimethylammonium bromide (CTAB) on the thermal transport between
98 + gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
99 +  et al.} studied the cooling dynamics, which is controlled by thermal
100 + interface resistance of glass-embedded metal
101 + nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
102 + normally considered barriers for heat transport, Alper {\it et al.}
103 + suggested that specific ligands (capping agents) could completely
104 + eliminate this barrier
105 + ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
106 +
107 + In previous simulations, we applied a variant of reverse
108 + non-equilibrium molecular dynamics (RNEMD) to calculate the
109 + interfacial thermal conductance at a metal / organic solvent interface
110 + that had been chemically protected by butanethiolate groups.  Our
111 + calculations suggest an explanation for the very large thermal
112 + conductivity at alkanethiol-capped metal surfaces when compared with
113 + bare metal/solvent interfaces.  Specifically, the chemical bond
114 + between the metal and the ligand introduces a vibrational overlap that
115 + is not present without the protecting group, and the overlap between
116 + the vibrational spectra (metal to ligand, ligand to solvent) provides
117 + a mechanism for rapid thermal transport across the interface.
118 +
119 + One interesting result of our previous work was the observation of
120 + {\it non-monotonic dependence} of the thermal conductance on the
121 + coverage of a metal surface by a chemical protecting group.  Our
122 + explanation for this behavior was that gaps in surface coverage
123 + allowed solvent to penetrate close to the capping molecules that had
124 + been heated by the metal surface, to absorb thermal energy from these
125 + molecules, and then diffuse away.  The effect of surface coverage is
126 + relatively difficult to study as the individual protecting groups have
127 + lateral mobility on the surface and can aggregate to form domains on
128 + the timescale of the simulation.
129 +
130 + The work reported here involves the use of velocity shearing and
131 + scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD) to
132 + study surfaces composed of mixed-length chains which collectively form
133 + a complete monolayer on the surfaces.  These complete (but
134 + mixed-chain) surfaces are significantly less prone to surface domain
135 + formation, and would allow us to further investigate the mechanism of
136 + heat transport to the solvent.
137 +
138   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
139   %                          **METHODOLOGY**
140   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
141   \section{Methodology}
142  
143 + There are many ways to compute bulk transport properties from
144 + classical molecular dynamics simulations.  Equilibrium Molecular
145 + Dynamics (EMD) simulations can be used by computing relevant time
146 + correlation functions and assuming that linear response theory
147 + holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
148 + For some transport properties (notably thermal conductivity), EMD
149 + approaches are subject to noise and poor convergence of the relevant
150 + correlation functions. Traditional Non-equilibrium Molecular Dynamics
151 + (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
152 + simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
153 + However, the resulting flux is often difficult to
154 + measure. Furthermore, problems arise for NEMD simulations of
155 + heterogeneous systems, such as phase-phase boundaries or interfaces,
156 + where the type of gradient to enforce at the boundary between
157 + materials is unclear.
158 +
159 + {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
160 + a different approach in that an unphysical {\it flux} is imposed
161 + between different regions or ``slabs'' of the simulation
162 + box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
163 + system responds by developing a temperature or momentum {\it gradient}
164 + between the two regions. Since the amount of the applied flux is known
165 + exactly, and the measurement of a gradient is generally less
166 + complicated, imposed-flux methods typically take shorter simulation
167 + times to obtain converged results for transport properties.  The
168 + corresponding temperature or velocity gradients which develop in
169 + response to the applied flux are then related (via linear response
170 + theory) to the transport coefficient of interest.  These methods are
171 + quite efficient, in that they do not need many trajectories to provide
172 + information about transport properties. To date, they have been
173 + utilized in computing thermal and mechanical transfer of both
174 + homogeneous
175 + liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
176 + well as heterogeneous
177 + systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
178 +
179          %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
180          %                           VSS-RNEMD
181          %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
182          \subsection{VSS-RNEMD}
183 <        
183 > The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
184 >  al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
185 > as a sequence of imaginary elastic collisions between particles in
186 > regions separated by half of the simulation cell.  In each collision,
187 > the entire momentum vectors of both particles may be exchanged to
188 > generate a thermal flux. Alternatively, a single component of the
189 > momentum vectors may be exchanged to generate a shear flux.  This
190 > algorithm turns out to be quite useful in many simulations of bulk
191 > liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
192 > system away from ideal Maxwell-Boltzmann distributions, and this has
193 > undesirable side-effects when the applied flux becomes
194 > large.\cite{Maginn:2010}        
195 >
196 > Instead of having momentum exchange between {\it individual particles}
197 > in each slab, the NIVS algorithm applies velocity scaling to all of
198 > the selected particles in both slabs.\cite{Kuang:2010uq} A combination
199 > of linear momentum, kinetic energy, and flux constraint equations
200 > governs the amount of velocity scaling performed at each step.  NIVS
201 > has been shown to be very effective at producing thermal gradients and
202 > for computing thermal conductivities, particularly for heterogeneous
203 > interfaces.  Although the NIVS algorithm can also be applied to impose
204 > a directional momentum flux, thermal anisotropy was observed in
205 > relatively high flux simulations, and the method is not suitable for
206 > imposing a shear flux or for computing shear viscosities.
207 >
208 > The most useful RNEMD
209 > approach developed so far utilizes a series of simultaneous velocity
210 > shearing and scaling exchanges between the two
211 > slabs.\cite{2012MolPh.110..691K} This method provides a set of
212 > conservation constraints while simultaneously creating a desired flux
213 > between the two slabs.  Satisfying the constraint equations ensures
214 > that the new configurations are sampled from the same NVE ensemble.
215 >
216 > The VSS moves are applied periodically to scale and shift the particle
217 > velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
218 > $C$) which are separated by half of the simulation box,
219 > \begin{displaymath}
220 > \begin{array}{rclcl}
221 >
222 > & \underline{\mathrm{shearing}} & &
223 > \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\  \\
224 > \mathbf{v}_i \leftarrow &
225 >  \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
226 >  \rangle\right)  +  \langle\mathbf{v}_c\rangle \\
227 > \mathbf{v}_j \leftarrow &
228 >  \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
229 >    \rangle\right) + \langle\mathbf{v}_h\rangle .
230 >
231 > \end{array}
232 > \end{displaymath}
233 > Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
234 > the center of mass velocities in the $C$ and $H$ slabs, respectively.
235 > Within the two slabs, particles receive incremental changes or a
236 > ``shear'' to their velocities.  The amount of shear is governed by the
237 > imposed momentum flux, $j_z(\mathbf{p})$
238 > \begin{eqnarray}
239 > \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
240 > \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
241 > \end{eqnarray}
242 > where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
243 > is the interval between two separate operations.
244 >
245 > To simultaneously impose a thermal flux ($J_z$) between the slabs we
246 > use energy conservation constraints,
247 > \begin{eqnarray}
248 > K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
249 > \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
250 > K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
251 > \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
252 > \mathbf{a}_h)^2 \label{vss4}.
253 > \label{constraint}
254 > \end{eqnarray}
255 > Simultaneous solution of these quadratic formulae for the scaling
256 > coefficients, $c$ and $h$, will ensure that the simulation samples from
257 > the original microcanonical (NVE) ensemble.  Here $K_{\{c,h\}}$ is the
258 > instantaneous translational kinetic energy of each slab.  At each time
259 > interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
260 > and $\mathbf{a}_h$, subject to the imposed momentum flux,
261 > $j_z(\mathbf{p})$, and thermal flux, $J_z$ values.  Since the VSS
262 > operations do not change the kinetic energy due to orientational
263 > degrees of freedom or the potential energy of a system, configurations
264 > after the VSS move have exactly the same energy ({\it and} linear
265 > momentum) as before the move.
266 >
267 > As the simulation progresses, the VSS moves are performed on a regular
268 > basis, and the system develops a thermal or velocity gradient in
269 > response to the applied flux.  Using the slope of the temperature or
270 > velocity gradient, it is quite simple to obtain of thermal
271 > conductivity ($\lambda$),
272 > \begin{equation}
273 > J_z = -\lambda \frac{\partial T}{\partial z},
274 > \end{equation}
275 > and shear viscosity ($\eta$),
276 > \begin{equation}
277 > j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
278 > \end{equation}
279 > Here, the quantities on the left hand side are the imposed flux
280 > values, while the slopes are obtained from linear fits to the
281 > gradients that develop in the simulated system.
282 >
283 > The VSS-RNEMD approach is versatile in that it may be used to
284 > implement both thermal and shear transport either separately or
285 > simultaneously.  Perturbations of velocities away from the ideal
286 > Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
287 > kept to a minimum.  This ability to generate simultaneous thermal and
288 > shear fluxes has been previously utilized to map out the shear
289 > viscosity of SPC/E water over a wide range of temperatures (90~K) with
290 > a {\it single 1 ns simulation}.\cite{2012MolPh.110..691K}
291 >
292          \begin{figure}
293                  \includegraphics[width=\linewidth]{figures/rnemd}
294                  \caption{VSS-RNEMD}
295                  \label{fig:rnemd}
296          \end{figure}
297          
67        \begin{equation}
68                \boldsymbol{\nu}_{i} \leftarrow c \cdot \left( \boldsymbol{\nu}_{i} - \langle \boldsymbol{\nu}_{c} \rangle \right) + \left( \langle \boldsymbol{\nu}_{c} \rangle + \bold{a}_{c} \right),
69        \end{equation}
298          
71        \begin{equation}
72                \boldsymbol{\nu}_{j} \leftarrow h \cdot \left( \boldsymbol{\nu}_{j} - \langle \boldsymbol{\nu}_{h} \rangle \right) + \left( \langle \boldsymbol{\nu}_{h} \rangle + \bold{a}_{h} \right),
73        \end{equation}
299          
75        \begin{equation}
76                K_{c} - J_{z} \Delta t = c^{2} \left( K_{c} - \frac{1}{2} M_{c} \langle \boldsymbol{\nu}_{c} \rangle ^{2}  \right) + \frac{1}{2} M_{c} \left( \langle \boldsymbol{\nu}_{c} \rangle + \bold{a}_{c} \right) ^{2},
77        \end{equation}
78        
79        \begin{equation}
80                K_{h} + J_{z} \Delta t = h^{2} \left( K_{h} - \frac{1}{2} M_{h} \langle \boldsymbol{\nu}_{h} \rangle ^{2}  \right) + \frac{1}{2} M_{h} \left( \langle \boldsymbol{\nu}_{h} \rangle + \bold{a}_{h} \right) ^{2},
81        \end{equation}
82                
83        
84        
300          %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
301          %                          INTERFACIAL CONDUCTANCE
302          %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303 <        \subsection{Defining Interfacial Conductance (G)}
304 <        
90 <        \begin{figure}
91 <                \includegraphics[width=\linewidth]{figures/resistor_series}
92 <                \caption{RESISTOR SERIES}
93 <                \label{fig:resistor_series}
94 <        \end{figure}
95 <        
96 <        
97 < There are two interfaces involved in the transfer of heat from the gold slab to the solvent: the gold/thiolate interface and the thiolate/solvent interface. We can treat the temperature on each side of an interface as discrete, making the interfacial conductance the inverse of the Kaptiza resistance, or $G = \frac{J}{\Delta T}$. To model the total conductance across multiple interfaces, we treat the interfaces as resistors in series. Resistors in series are additive, ($R_{total} = R_{1} + R_{2} + R_{3} + ...$) and the total conductance is the inverse of the total resistance, or $G = \frac{1}{(R_{1} + R_{2} + R_{3} + ...}$). We treat each bin in the VSS-RNEMD temperature profile as a resistor with resistance $\frac{T_{2}-T_{1}}{J}$, $\frac{T_{3}-T_{2}}{J}$, etc. The sum of this resistor series which spans the gold/thiolate interface, thiolate chains, and thiolate/solvent interface simplifies to
303 >        \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
304 >  to interfacial transport}
305  
306 < \begin{equation}
307 <        \frac{T_{n}-T_{1}}{J},
308 <        \label{eq:finalG}
309 <        \end{equation}
306 > Interfaces between dissimilar materials have transport properties
307 > which can be defined as derivatives of the standard transport
308 > coefficients in a direction normal to the interface. For example, the
309 > {\it interfacial} thermal conductance ($G$) can be thought of as the
310 > change in the thermal conductivity ($\lambda$) across the boundary
311 > between materials:
312 > \begin{align}
313 >  G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
314 >  &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
315 >  \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
316 >  \label{derivativeG}
317 > \end{align}
318 > where $z_0$ is the location of the interface between two materials and
319 > $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
320 > to be the $z$ direction from here on).  RNEMD simulations, and
321 > particularly the VSS-RNEMD approach, function by applying a momentum
322 > or thermal flux and watching the gradient response of the
323 > material. This means that the {\it interfacial} conductance is a
324 > second derivative property which is subject to significant noise and
325 > therefore requires extensive sampling.  We have been able to
326 > demonstrate the use of the second derivative approach to compute
327 > interfacial conductance at chemically-modified metal / solvent
328 > interfaces.  However, a definition of the interfacial conductance in
329 > terms of a temperature difference ($\Delta T$) measured at $z_0$,
330 > \begin{displaymath}
331 > G = \frac{J_z}{\Delta T_{z_0}},
332 > \end{displaymath}
333 > is useful once the RNEMD approach has generated a stable temperature
334 > gap across the interface.
335  
336 < or the temperature difference between the gold side of the gold/thiolate interface and the solvent side of the thiolate/solvent interface over the applied flux.
336 > \begin{figure}
337 >  \includegraphics[width=\linewidth]{figures/resistor_series}
338 >  \caption{RESISTOR SERIES}
339 >  \label{fig:resistor_series}
340 > \end{figure}
341 >
342 > In the particular case we are studying here, there are two interfaces
343 > involved in the transfer of heat from the gold slab to the solvent:
344 > the gold/thiolate interface and the thiolate/solvent interface. We
345 > could treat the temperature on each side of an interface as discrete,
346 > making the interfacial conductance the inverse of the Kaptiza
347 > resistance, or $G = \frac{1}{R_k}$. To model the total conductance
348 > across multiple interfaces, it is useful to think of the interfaces as
349 > a set of resistors wired in series. The total resistance is then
350 > additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
351 > is the inverse of the total resistance, or $G = \frac{1}{\sum_i
352 >  R_i}$).  In the interfacial region, we treat each bin in the
353 > VSS-RNEMD temperature profile as a resistor with resistance
354 > $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc.  The sum of
355 > the set of resistors which spans the gold/thiolate interface, thiolate
356 > chains, and thiolate/solvent interface simplifies to
357 > \begin{equation}
358 >  \frac{T_{n}-T_{1}}{J_z},
359 >  \label{eq:finalG}
360 > \end{equation}
361 > or the temperature difference between the gold side of the
362 > gold/thiolate interface and the solvent side of the thiolate/solvent
363 > interface over the applied flux.  
364          
365          
366   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
# Line 109 | Line 368 | or the temperature difference between the gold side of
368   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
369   \section{Computational Details}
370  
371 <        %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372 <        %                          SIMULATION PROTOCOL
373 <        %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374 <        \subsection{Simulation Protocol}
116 <        
117 < We have implemented the VSS-RNEMD algorithm in OpenMD, our group molecular dynamics code. A gold slab 11 layers thick was equilibrated at 1 atm and 200 K. The periodic box was expanded in the z direction to expose two Au(111) faces.
371 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372 > % SIMULATION PROTOCOL
373 > %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374 > \subsection{Simulation Protocol}
375  
376 + We have implemented the VSS-RNEMD algorithm in OpenMD, our
377 + group molecular dynamics code. A gold slab 11 atoms thick was
378 + equilibrated at 1 atm and 200 K. The periodic box was expanded
379 + in the z direction to expose two Au(111) faces.
380 +
381   A full monolayer of thiolates (1/3 the number of surface gold atoms) was placed on three-fold hollow sites on the gold interfaces. To efficiently test the effect of thiolate binding sites on the thermal conductance, all systems had one gold interface with thiolates placed only on fcc hollow sites and the other interface with thiolates only on hcp hollow sites. To test the effect of thiolate chain length on interfacial thermal conductance, full coverages of five chain lengths were tested: butanethiolate, hexanethiolate, octanethiolate, decanethiolate, and dodecanethiolate. To test the effect of mixed chain lengths, full coverages of butanethiolate/decanethiolate and butanethiolate/dodecanethiolate mixtures were created in short/long chain ratios of 25/75, 50/50, and 75/25. The short and long chains were placed on the surface in a random configuration.
382  
383   The simulation box z dimension was set to roughly double the length of the gold/thiolate block. Solvent molecules were placed in the vacant portion of the box using the packmol algorithm. Two solvent molecules were examined: hexane and toluene. Hexane, a straight chain flexible alkane, is very structurally similar to the thiolate alkane tails while toluene is a rigid planar molecule.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines