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

Comparing interfacial/interfacial.tex (file contents):
Revision 3750 by skuang, Tue Jul 26 18:10:49 2011 UTC vs.
Revision 3751 by gezelter, Tue Jul 26 19:43:10 2011 UTC

# Line 73 | Line 73 | Interfacial thermal conductance is extensively studied
73   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74  
75   \section{Introduction}
76 < Interfacial thermal conductance is extensively studied both
77 < experimentally and computationally\cite{cahill:793}, due to its
78 < importance in nanoscale science and technology. Reliability of
79 < nanoscale devices depends on their thermal transport
80 < properties. Unlike bulk homogeneous materials, nanoscale materials
81 < features significant presence of interfaces, and these interfaces
82 < could dominate the heat transfer behavior of these
83 < materials. Furthermore, these materials are generally heterogeneous,
84 < which challenges traditional research methods for homogeneous
85 < systems.
76 > Due to the importance of heat flow in nanotechnology, interfacial
77 > thermal conductance has been studied extensively both experimentally
78 > and computationally.\cite{cahill:793} Unlike bulk materials, nanoscale
79 > materials have a significant fraction of their atoms at interfaces,
80 > and the chemical details of these interfaces govern the heat transfer
81 > behavior. Furthermore, the interfaces are
82 > heterogeneous (e.g. solid - liquid), which provides a challenge to
83 > traditional methods developed for homogeneous systems.
84  
85 < Heat conductance of molecular and nano-scale interfaces will be
86 < affected by the chemical details of the surface. Experimentally,
87 < various interfaces have been investigated for their thermal
88 < conductance properties. Wang {\it et al.} studied heat transport
89 < through long-chain hydrocarbon monolayers on gold substrate at
90 < individual molecular level\cite{Wang10082007}; Schmidt {\it et al.}
93 < studied the role of CTAB on thermal transport between gold nanorods
94 < and solvent\cite{doi:10.1021/jp8051888}; Juv\'e {\it et al.} studied
85 > Experimentally, various interfaces have been investigated for their
86 > thermal conductance. Wang {\it et al.} studied heat transport through
87 > long-chain hydrocarbon monolayers on gold substrate at individual
88 > molecular level,\cite{Wang10082007} Schmidt {\it et al.} studied the
89 > role of CTAB on thermal transport between gold nanorods and
90 > solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it et al.} studied
91   the cooling dynamics, which is controlled by thermal interface
92   resistence of glass-embedded metal
93 < nanoparticles\cite{PhysRevB.80.195406}. Although interfaces are
94 < commonly barriers for heat transport, Alper {\it et al.} suggested
95 < that specific ligands (capping agents) could completely eliminate this
96 < barrier ($G\rightarrow\infty$)\cite{doi:10.1021/la904855s}.
93 > nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
94 > normally considered barriers for heat transport, Alper {\it et al.}
95 > suggested that specific ligands (capping agents) could completely
96 > eliminate this barrier
97 > ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
98  
99   Theoretical and computational models have also been used to study the
100   interfacial thermal transport in order to gain an understanding of
# Line 105 | Line 102 | atoms)\cite{hase:2010,hase:2011}. However, ensemble av
102   employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
103   study thermal transport from hot Au(111) substrate to a self-assembled
104   monolayer of alkylthiol with relatively long chain (8-20 carbon
105 < atoms)\cite{hase:2010,hase:2011}. However, ensemble averaged
105 > atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
106   measurements for heat conductance of interfaces between the capping
107 < monolayer on Au and a solvent phase has yet to be studied.
108 < The comparatively low thermal flux through interfaces is
107 > monolayer on Au and a solvent phase have yet to be studied with their
108 > approach. The comparatively low thermal flux through interfaces is
109   difficult to measure with Equilibrium MD or forward NEMD simulation
110   methods. Therefore, the Reverse NEMD (RNEMD)
111 < methods\cite{MullerPlathe:1997xw} would have the advantage of having
112 < this difficult to measure flux known when studying the thermal
113 < transport across interfaces, given that the simulation methods being
114 < able to effectively apply an unphysical flux in non-homogeneous
115 < systems. Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009}
116 < applied this approach to various liquid interfaces and studied how
117 < thermal conductance (or resistance) is dependent on chemistry details
118 < of interfaces, e.g. hydrophobic and hydrophilic aqueous interfaces.
111 > methods\cite{MullerPlathe:1997xw,kuang:164101} would have the
112 > advantage of applying this difficult to measure flux (while measuring
113 > the resulting gradient), given that the simulation methods being able
114 > to effectively apply an unphysical flux in non-homogeneous systems.
115 > Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
116 > this approach to various liquid interfaces and studied how thermal
117 > conductance (or resistance) is dependent on chemistry details of
118 > interfaces, e.g. hydrophobic and hydrophilic aqueous interfaces.
119  
120 < Recently, we have developed the Non-Isotropic Velocity Scaling (NIVS)
120 > Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
121   algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
122   retains the desirable features of RNEMD (conservation of linear
123   momentum and total energy, compatibility with periodic boundary
# Line 137 | Line 134 | underlying mechanism for the phenomena was investigate
134   thermal transport across these interfaces was studied and the
135   underlying mechanism for the phenomena was investigated.
136  
140 [MAY ADD WHY STUDY AU-THIOL SURFACE; CITE SHAOYI JIANG]
141
137   \section{Methodology}
138   \subsection{Imposd-Flux Methods in MD Simulations}
139 < Steady state MD simulations has the advantage that not many
139 > Steady state MD simulations have an advantage in that not many
140   trajectories are needed to study the relationship between thermal flux
141 < and thermal gradients. For systems including low conductance
142 < interfaces one must have a method capable of generating or measuring
143 < relatively small fluxes, compared to those required for bulk
144 < conductivity. This requirement makes the calculation even more
145 < difficult for those slowly-converging equilibrium
146 < methods\cite{Viscardy:2007lq}. Forward methods may impose gradient,
147 < but in interfacial conditions it is not clear what behavior to impose
148 < at the interfacial boundaries. Imposed-flux reverse non-equilibrium
149 < methods\cite{MullerPlathe:1997xw} have the flux set {\it a priori} and
150 < the thermal response becomes easier to measure than the flux. Although
141 > and thermal gradients. For systems with low interfacial conductance,
142 > one must have a method capable of generating or measuring relatively
143 > small fluxes, compared to those required for bulk conductivity. This
144 > requirement makes the calculation even more difficult for
145 > slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
146 > NEMD methods impose a gradient (and measure a flux), but at interfaces
147 > it is not clear what behavior should be imposed at the boundaries
148 > between materials.  Imposed-flux reverse non-equilibrium
149 > methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
150 > the thermal response becomes an easy-to-measure quantity.  Although
151   M\"{u}ller-Plathe's original momentum swapping approach can be used
152   for exchanging energy between particles of different identity, the
153   kinetic energy transfer efficiency is affected by the mass difference
154   between the particles, which limits its application on heterogeneous
155   interfacial systems.
156  
157 < The non-isotropic velocity scaling (NIVS)\cite{kuang:164101} approach to
158 < non-equilibrium MD simulations is able to impose a wide range of
157 > The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
158 > to non-equilibrium MD simulations is able to impose a wide range of
159   kinetic energy fluxes without obvious perturbation to the velocity
160   distributions of the simulated systems. Furthermore, this approach has
161   the advantage in heterogeneous interfaces in that kinetic energy flux
# Line 177 | Line 172 | momenta and energy and does not depend on an external
172   for computing thermal conductivities. The NIVS algorithm conserves
173   momenta and energy and does not depend on an external thermostat.
174  
175 < \subsection{Defining Interfacial Thermal Conductivity $G$}
176 < Given a system with thermal gradients and the corresponding thermal
177 < flux, for interfaces with a relatively low interfacial conductance,
178 < the bulk regions on either side of an interface rapidly come to a
179 < state in which the two phases have relatively homogeneous (but
180 < distinct) temperatures. The interfacial thermal conductivity $G$ can
181 < therefore be approximated as:
175 > \subsection{Defining Interfacial Thermal Conductivity ($G$)}
176 >
177 > For an interface with relatively low interfacial conductance, and a
178 > thermal flux between two distinct bulk regions, the regions on either
179 > side of the interface rapidly come to a state in which the two phases
180 > have relatively homogeneous (but distinct) temperatures. The
181 > interfacial thermal conductivity $G$ can therefore be approximated as:
182   \begin{equation}
183 < G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
183 >  G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
184      \langle T_\mathrm{cold}\rangle \right)}
185   \label{lowG}
186   \end{equation}
187 < where ${E_{total}}$ is the imposed non-physical kinetic energy
188 < transfer and ${\langle T_\mathrm{hot}\rangle}$ and ${\langle
189 <  T_\mathrm{cold}\rangle}$ are the average observed temperature of the
190 < two separated phases.
187 > where ${E_{total}}$ is the total imposed non-physical kinetic energy
188 > transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
189 > and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
190 > temperature of the two separated phases.
191  
192   When the interfacial conductance is {\it not} small, there are two
193 < ways to define $G$.
193 > ways to define $G$. One way is to assume the temperature is discrete
194 > on the two sides of the interface. $G$ can be calculated using the
195 > applied thermal flux $J$ and the maximum temperature difference
196 > measured along the thermal gradient max($\Delta T$), which occurs at
197 > the Gibbs deviding surface (Figure \ref{demoPic}): \begin{equation}
198 >  G=\frac{J}{\Delta T} \label{discreteG} \end{equation}
199  
200 One way is to assume the temperature is discrete on the two sides of
201 the interface. $G$ can be calculated using the applied thermal flux
202 $J$ and the maximum temperature difference measured along the thermal
203 gradient max($\Delta T$), which occurs at the Gibbs deviding surface
204 (Figure \ref{demoPic}):
205 \begin{equation}
206 G=\frac{J}{\Delta T}
207 \label{discreteG}
208 \end{equation}
209
200   \begin{figure}
201   \includegraphics[width=\linewidth]{method}
202   \caption{Interfacial conductance can be calculated by applying an
# Line 223 | Line 213 | the magnitude of thermal conductivity $\lambda$ change
213  
214   The other approach is to assume a continuous temperature profile along
215   the thermal gradient axis (e.g. $z$) and define $G$ at the point where
216 < the magnitude of thermal conductivity $\lambda$ change reach its
216 > the magnitude of thermal conductivity ($\lambda$) change reaches its
217   maximum, given that $\lambda$ is well-defined throughout the space:
218   \begin{equation}
219   G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
# Line 234 | Line 224 | With the temperature profile obtained from simulations
224   \label{derivativeG}
225   \end{equation}
226  
227 < With the temperature profile obtained from simulations, one is able to
227 > With temperature profiles obtained from simulation, one is able to
228   approximate the first and second derivatives of $T$ with finite
229 < difference methods and thus calculate $G^\prime$.
229 > difference methods and calculate $G^\prime$. In what follows, both
230 > definitions have been used, and are compared in the results.
231  
232 < In what follows, both definitions have been used for calculation and
233 < are compared in the results.
234 <
235 < To compare the above definitions ($G$ and $G^\prime$), we have modeled
236 < a metal slab with its (111) surfaces perpendicular to the $z$-axis of
246 < our simulation cells. Both with and without capping agents on the
247 < surfaces, the metal slab is solvated with simple organic solvents, as
232 > To investigate the interfacial conductivity at metal / solvent
233 > interfaces, we have modeled a metal slab with its (111) surfaces
234 > perpendicular to the $z$-axis of our simulation cells. The metal slab
235 > has been prepared both with and without capping agents on the exposed
236 > surface, and has been solvated with simple organic solvents, as
237   illustrated in Figure \ref{gradT}.
238  
239   With the simulation cell described above, we are able to equilibrate
240   the system and impose an unphysical thermal flux between the liquid
241   and the metal phase using the NIVS algorithm. By periodically applying
242 < the unphysical flux, we are able to obtain a temperature profile and
243 < its spatial derivatives. These quantities enable the evaluation of the
244 < interfacial thermal conductance of a surface. Figure \ref{gradT} is an
245 < example of how an applied thermal flux can be used to obtain the 1st
257 < and 2nd derivatives of the temperature profile.
242 > the unphysical flux, we obtained a temperature profile and its spatial
243 > derivatives. Figure \ref{gradT} shows how an applied thermal flux can
244 > be used to obtain the 1st and 2nd derivatives of the temperature
245 > profile.
246  
247   \begin{figure}
248   \includegraphics[width=\linewidth]{gradT}
# Line 268 | Line 256 | OpenMD\cite{Meineke:2005gd,openmd}, and was used for o
256   \section{Computational Details}
257   \subsection{Simulation Protocol}
258   The NIVS algorithm has been implemented in our MD simulation code,
259 < OpenMD\cite{Meineke:2005gd,openmd}, and was used for our
260 < simulations. Different metal slab thickness (layer numbers of Au) was
261 < simulated. Metal slabs were first equilibrated under atmospheric
262 < pressure (1 atm) and a desired temperature (e.g. 200K). After
263 < equilibration, butanethiol capping agents were placed at three-fold
264 < hollow sites on the Au(111) surfaces. These sites could be either a
265 < {\it fcc} or {\it hcp} site. However, Hase {\it et al.} found that
266 < they are equivalent in a heat transfer process\cite{hase:2010}, so
279 < they are not distinguished in our study. The maximum butanethiol
259 > OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
260 > Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
261 > under atmospheric pressure (1 atm) and 200K. After equilibration,
262 > butanethiol capping agents were placed at three-fold hollow sites on
263 > the Au(111) surfaces. These sites are either {\it fcc} or {\it
264 >  hcp} sites, although Hase {\it et al.} found that they are
265 > equivalent in a heat transfer process,\cite{hase:2010} so we did not
266 > distinguish between these sites in our study. The maximum butanethiol
267   capacity on Au surface is $1/3$ of the total number of surface Au
268   atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
269   structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
270 < series of different coverages was derived by evenly eliminating
271 < butanethiols on the surfaces, and was investigated in order to study
272 < the relation between coverage and interfacial conductance.
270 > series of lower coverages was also prepared by eliminating
271 > butanethiols from the higher coverage surface in a regular manner. The
272 > lower coverages were prepared in order to study the relation between
273 > coverage and interfacial conductance.
274  
275   The capping agent molecules were allowed to migrate during the
276   simulations. They distributed themselves uniformly and sampled a
277   number of three-fold sites throughout out study. Therefore, the
278 < initial configuration would not noticeably affect the sampling of a
278 > initial configuration does not noticeably affect the sampling of a
279   variety of configurations of the same coverage, and the final
280   conductance measurement would be an average effect of these
281 < configurations explored in the simulations. [MAY NEED SNAPSHOTS]
281 > configurations explored in the simulations.
282  
283 < After the modified Au-butanethiol surface systems were equilibrated
284 < under canonical ensemble, organic solvent molecules were packed in the
285 < previously empty part of the simulation cells\cite{packmol}. Two
283 > After the modified Au-butanethiol surface systems were equilibrated in
284 > the canonical (NVT) ensemble, organic solvent molecules were packed in
285 > the previously empty part of the simulation cells.\cite{packmol} Two
286   solvents were investigated, one which has little vibrational overlap
287 < with the alkanethiol and a planar shape (toluene), and one which has
288 < similar vibrational frequencies and chain-like shape ({\it n}-hexane).
287 > with the alkanethiol and which has a planar shape (toluene), and one
288 > which has similar vibrational frequencies to the capping agent and
289 > chain-like shape ({\it n}-hexane).
290  
291 < The space filled by solvent molecules, i.e. the gap between
292 < periodically repeated Au-butanethiol surfaces should be carefully
293 < chosen. A very long length scale for the thermal gradient axis ($z$)
305 < may cause excessively hot or cold temperatures in the middle of the
291 > The simulation cells were not particularly extensive along the
292 > $z$-axis, as a very long length scale for the thermal gradient may
293 > cause excessively hot or cold temperatures in the middle of the
294   solvent region and lead to undesired phenomena such as solvent boiling
295   or freezing when a thermal flux is applied. Conversely, too few
296   solvent molecules would change the normal behavior of the liquid
297   phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
298 < these extreme cases did not happen to our simulations. And the
299 < corresponding spacing is usually $35 \sim 75$\AA.
298 > these extreme cases did not happen to our simulations. The spacing
299 > between periodic images of the gold interfaces is $35 \sim 75$\AA.
300  
301   The initial configurations generated are further equilibrated with the
302 < $x$ and $y$ dimensions fixed, only allowing length scale change in $z$
303 < dimension. This is to ensure that the equilibration of liquid phase
304 < does not affect the metal crystal structure in $x$ and $y$ dimensions.
305 < To investigate this effect, comparisons were made with simulations
306 < that allow changes of $L_x$ and $L_y$ during NPT equilibration, and
307 < the results are shown in later sections. After ensuring the liquid
308 < phase reaches equilibrium at atmospheric pressure (1 atm), further
309 < equilibration are followed under NVT and then NVE ensembles.
302 > $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
303 > change. This is to ensure that the equilibration of liquid phase does
304 > not affect the metal's crystalline structure. Comparisons were made
305 > with simulations that allowed changes of $L_x$ and $L_y$ during NPT
306 > equilibration. No substantial changes in the box geometry were noticed
307 > in these simulations. After ensuring the liquid phase reaches
308 > equilibrium at atmospheric pressure (1 atm), further equilibration was
309 > carried out under canonical (NVT) and microcanonical (NVE) ensembles.
310  
311 < After the systems reach equilibrium, NIVS is implemented to impose a
312 < periodic unphysical thermal flux between the metal and the liquid
313 < phase. Most of our simulations are under an average temperature of
314 < $\sim$200K. Therefore, this flux usually comes from the metal to the
311 > After the systems reach equilibrium, NIVS was used to impose an
312 > unphysical thermal flux between the metal and the liquid phases. Most
313 > of our simulations were done under an average temperature of
314 > $\sim$200K. Therefore, thermal flux usually came from the metal to the
315   liquid so that the liquid has a higher temperature and would not
316 < freeze due to excessively low temperature. After this induced
317 < temperature gradient is stablized, the temperature profile of the
318 < simulation cell is recorded. To do this, the simulation cell is
319 < devided evenly into $N$ slabs along the $z$-axis and $N$ is maximized
332 < for highest possible spatial resolution but not too many to have some
333 < slabs empty most of the time. The average temperatures of each slab
316 > freeze due to lowered temperatures. After this induced temperature
317 > gradient had stablized, the temperature profile of the simulation cell
318 > was recorded. To do this, the simulation cell is devided evenly into
319 > $N$ slabs along the $z$-axis. The average temperatures of each slab
320   are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
321   the same, the derivatives of $T$ with respect to slab number $n$ can
322 < be directly used for $G^\prime$ calculations:
323 < \begin{equation}
338 < G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
322 > be directly used for $G^\prime$ calculations: \begin{equation}
323 >  G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
324           \Big/\left(\frac{\partial T}{\partial z}\right)^2
325           = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
326           \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
# Line 344 | Line 329 | All of the above simulation procedures use a time step
329   \label{derivativeG2}
330   \end{equation}
331  
332 < All of the above simulation procedures use a time step of 1 fs. And
333 < each equilibration / stabilization step usually takes 100 ps, or
334 < longer, if necessary.
332 > All of the above simulation procedures use a time step of 1 fs. Each
333 > equilibration stage took a minimum of 100 ps, although in some cases,
334 > longer equilibration stages were utilized.
335  
336   \subsection{Force Field Parameters}
337 < Our simulations include various components. Figure \ref{demoMol}
338 < demonstrates the sites defined for both United-Atom and All-Atom
339 < models of the organic solvent and capping agent molecules in our
340 < simulations. Force field parameter descriptions are needed for
337 > Our simulations include a number of chemically distinct components.
338 > Figure \ref{demoMol} demonstrates the sites defined for both
339 > United-Atom and All-Atom models of the organic solvent and capping
340 > agents in our simulations. Force field parameters are needed for
341   interactions both between the same type of particles and between
342   particles of different species.
343  
# Line 369 | Line 354 | quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59
354   \end{figure}
355  
356   The Au-Au interactions in metal lattice slab is described by the
357 < quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59.3527}. The QSC
357 > quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
358   potentials include zero-point quantum corrections and are
359   reparametrized for accurate surface energies compared to the
360 < Sutton-Chen potentials\cite{Chen90}.
360 > Sutton-Chen potentials.\cite{Chen90}
361  
362 < For both solvent molecules, straight chain {\it n}-hexane and aromatic
363 < toluene, United-Atom (UA) and All-Atom (AA) models are used
364 < respectively. The TraPPE-UA
362 > For the two solvent molecules, {\it n}-hexane and toluene, two
363 > different atomistic models were utilized. Both solvents were modeled
364 > using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
365   parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
366   for our UA solvent molecules. In these models, sites are located at
367   the carbon centers for alkyl groups. Bonding interactions, including
368   bond stretches and bends and torsions, were used for intra-molecular
369 < sites not separated by more than 3 bonds. Otherwise, for non-bonded
370 < interactions, Lennard-Jones potentials are used. [CHECK CITATION]
369 > sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
370 > potentials are used.
371  
372 < By eliminating explicit hydrogen atoms, these models are simple and
373 < computationally efficient, while maintains good accuracy. However, the
374 < TraPPE-UA for alkanes is known to predict a lower boiling point than
375 < experimental values. Considering that after an unphysical thermal flux
376 < is applied to a system, the temperature of ``hot'' area in the liquid
377 < phase would be significantly higher than the average of the system, to
378 < prevent over heating and boiling of the liquid phase, the average
379 < temperature in our simulations should be much lower than the liquid
380 < boiling point.
372 > By eliminating explicit hydrogen atoms, the TraPPE-UA models are
373 > simple and computationally efficient, while maintaining good accuracy.
374 > However, the TraPPE-UA model for alkanes is known to predict a slighly
375 > lower boiling point than experimental values. This is one of the
376 > reasons we used a lower average temperature (200K) for our
377 > simulations. If heat is transferred to the liquid phase during the
378 > NIVS simulation, the liquid in the hot slab can actually be
379 > substantially warmer than the mean temperature in the simulation. The
380 > lower mean temperatures therefore prevent solvent boiling.
381  
382 < For UA-toluene model, the non-bonded potentials between
383 < inter-molecular sites have a similar Lennard-Jones formulation. For
384 < intra-molecular interactions, considering the stiffness of the benzene
385 < ring, rigid body constraints are applied for further computational
386 < efficiency. All bonds in the benzene ring and between the ring and the
402 < methyl group remain rigid during the progress of simulations.
382 > For UA-toluene, the non-bonded potentials between intermolecular sites
383 > have a similar Lennard-Jones formulation. The toluene molecules were
384 > treated as a single rigid body, so there was no need for
385 > intramolecular interactions (including bonds, bends, or torsions) in
386 > this solvent model.
387  
388   Besides the TraPPE-UA models, AA models for both organic solvents are
389   included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
390 < force field is used. Additional explicit hydrogen sites were
390 > force field is used, and additional explicit hydrogen sites were
391   included. Besides bonding and non-bonded site-site interactions,
392   partial charges and the electrostatic interactions were added to each
393   CT and HC site. For toluene, the United Force Field developed by
394 < Rapp\'{e} {\it et al.}\cite{doi:10.1021/ja00051a040} is
395 < adopted. Without the rigid body constraints, bonding interactions were
396 < included. For the aromatic ring, improper torsions (inversions) were
397 < added as an extra potential for maintaining the planar shape.
414 < [CHECK CITATION]
394 > Rapp\'{e} {\it et al.}\cite{doi:10.1021/ja00051a040} was adopted, and
395 > a flexible model for the toluene molecule was utilized which included
396 > bond, bend, torsion, and inversion potentials to enforce ring
397 > planarity.
398  
399 < The capping agent in our simulations, the butanethiol molecules can
400 < either use UA or AA model. The TraPPE-UA force fields includes
399 > The butanethiol capping agent in our simulations, were also modeled
400 > with both UA and AA model. The TraPPE-UA force field includes
401   parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
402   UA butanethiol model in our simulations. The OPLS-AA also provides
403   parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
404 < surfaces do not have the hydrogen atom bonded to sulfur. To adapt this
405 < change and derive suitable parameters for butanethiol adsorbed on
406 < Au(111) surfaces, we adopt the S parameters from Luedtke and
407 < Landman\cite{landman:1998}[CHECK CITATION]
408 < and modify parameters for its neighbor C
409 < atom for charge balance in the molecule. Note that the model choice
410 < (UA or AA) of capping agent can be different from the
411 < solvent. Regardless of model choice, the force field parameters for
429 < interactions between capping agent and solvent can be derived using
430 < Lorentz-Berthelot Mixing Rule:
404 > surfaces do not have the hydrogen atom bonded to sulfur. To derive
405 > suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
406 > adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
407 > modify the parameters for the CTS atom to maintain charge neutrality
408 > in the molecule.  Note that the model choice (UA or AA) for the capping
409 > agent can be different from the solvent. Regardless of model choice,
410 > the force field parameters for interactions between capping agent and
411 > solvent can be derived using Lorentz-Berthelot Mixing Rule:
412   \begin{eqnarray}
413 < \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
414 < \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
413 >  \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
414 >  \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
415   \end{eqnarray}
416  
417 < To describe the interactions between metal Au and non-metal capping
418 < agent and solvent particles, we refer to an adsorption study of alkyl
419 < thiols on gold surfaces by Vlugt {\it et
420 <  al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
421 < form of potential parameters for the interaction between Au and
422 < pseudo-atoms CH$_x$ and S based on a well-established and widely-used
423 < effective potential of Hautman and Klein\cite{hautman:4994} for the
424 < Au(111) surface. As our simulations require the gold lattice slab to
425 < be non-rigid so that it could accommodate kinetic energy for thermal
445 < transport study purpose, the pair-wise form of potentials is
446 < preferred.
417 > To describe the interactions between metal (Au) and non-metal atoms,
418 > we refer to an adsorption study of alkyl thiols on gold surfaces by
419 > Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
420 > Lennard-Jones form of potential parameters for the interaction between
421 > Au and pseudo-atoms CH$_x$ and S based on a well-established and
422 > widely-used effective potential of Hautman and Klein for the Au(111)
423 > surface.\cite{hautman:4994} As our simulations require the gold slab
424 > to be flexible to accommodate thermal excitation, the pair-wise form
425 > of potentials they developed was used for our study.
426  
427 < Besides, the potentials developed from {\it ab initio} calculations by
428 < Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
429 < interactions between Au and aromatic C/H atoms in toluene. A set of
430 < pseudo Lennard-Jones parameters were provided for Au in their force
431 < fields. By using the Mixing Rule, this can be used to derive pair-wise
432 < potentials for non-bonded interactions between Au and non-metal sites.
433 <
434 < However, the Lennard-Jones parameters between Au and other types of
435 < particles, such as All-Atom normal alkanes in our simulations are not
436 < yet well-established. For these interactions, we attempt to derive
437 < their parameters using the Mixing Rule. To do this, Au pseudo
459 < Lennard-Jones parameters for ``Metal-non-Metal'' (MnM) interactions
460 < were first extracted from the Au-CH$_x$ parameters by applying the
461 < Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
462 < parameters in our simulations.
427 > The potentials developed from {\it ab initio} calculations by Leng
428 > {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
429 > interactions between Au and aromatic C/H atoms in toluene. However,
430 > the Lennard-Jones parameters between Au and other types of particles,
431 > (e.g. AA alkanes) have not yet been established. For these
432 > interactions, the Lorentz-Berthelot mixing rule can be used to derive
433 > effective single-atom LJ parameters for the metal using the fit values
434 > for toluene. These are then used to construct reasonable mixing
435 > parameters for the interactions between the gold and other atoms.
436 > Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
437 > our simulations.
438  
439   \begin{table*}
440    \begin{minipage}{\linewidth}
# Line 496 | Line 471 | parameters in our simulations.
471    \end{minipage}
472   \end{table*}
473  
474 < \subsection{Vibrational Spectrum}
474 > \subsection{Vibrational Power Spectrum}
475 >
476   To investigate the mechanism of interfacial thermal conductance, the
477 < vibrational spectrum is utilized as a complementary tool. Vibrational
478 < spectra were taken for individual components in different
479 < simulations. To obtain these spectra, simulations were run after
480 < equilibration, in the NVE ensemble. Snapshots of configurations were
481 < collected at a frequency that is higher than that of the fastest
482 < vibrations occuring in the simulations. With these configurations, the
483 < velocity auto-correlation functions can be computed:
477 > vibrational power spectrum was computed. Power spectra were taken for
478 > individual components in different simulations. To obtain these
479 > spectra, simulations were run after equilibration, in the NVE
480 > ensemble, and without a thermal gradient. Snapshots of configurations
481 > were collected at a frequency that is higher than that of the fastest
482 > vibrations occuring in the simulations. With these configurations, the
483 > velocity auto-correlation functions can be computed:
484   \begin{equation}
485   C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
486 < \label{vCorr}
487 < \end{equation}
488 < Followed by Fourier transforms, the power spectrum can be constructed:
486 > \label{vCorr}
487 > \end{equation}
488 > The power spectrum is constructed via a Fourier transform of the
489 > symmetrized velocity autocorrelation function,
490   \begin{equation}
491 < \hat{f}(\omega) = \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
492 < \label{fourier}
491 >  \hat{f}(\omega) =
492 >  \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
493 > \label{fourier}
494   \end{equation}
495  
496   \section{Results and Discussions}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines