ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3755
Committed: Fri Jul 29 15:45:14 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 50087 byte(s)
Log Message:
new edits

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{setspace}
5 \usepackage{endfloat}
6 \usepackage{caption}
7 %\usepackage{tabularx}
8 \usepackage{graphicx}
9 \usepackage{multirow}
10 %\usepackage{booktabs}
11 %\usepackage{bibentry}
12 %\usepackage{mathrsfs}
13 %\usepackage[ref]{overcite}
14 \usepackage[square, comma, sort&compress]{natbib}
15 \usepackage{url}
16 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18 9.0in \textwidth 6.5in \brokenpenalty=10000
19
20 % double space list of tables and figures
21 \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22 \setlength{\abovecaptionskip}{20 pt}
23 \setlength{\belowcaptionskip}{30 pt}
24
25 %\renewcommand\citemid{\ } % no comma in optional reference note
26 \bibpunct{[}{]}{,}{n}{}{;}
27 \bibliographystyle{achemso}
28
29 \begin{document}
30
31 \title{Simulating interfacial thermal conductance at metal-solvent
32 interfaces: the role of chemical capping agents}
33
34 \author{Shenyu Kuang and J. Daniel
35 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
36 Department of Chemistry and Biochemistry,\\
37 University of Notre Dame\\
38 Notre Dame, Indiana 46556}
39
40 \date{\today}
41
42 \maketitle
43
44 \begin{doublespace}
45
46 \begin{abstract}
47
48 With the Non-Isotropic Velocity Scaling algorithm (NIVS) we have
49 developed, an unphysical thermal flux can be effectively set up even
50 for non-homogeneous systems like interfaces in non-equilibrium
51 molecular dynamics simulations. In this work, this algorithm is
52 applied for simulating thermal conductance at metal / organic solvent
53 interfaces with various coverages of butanethiol capping
54 agents. Different solvents and force field models were tested. Our
55 results suggest that the United-Atom models are able to provide an
56 estimate of the interfacial thermal conductivity comparable to
57 experiments in our simulations with satisfactory computational
58 efficiency. From our results, the acoustic impedance mismatch between
59 metal and liquid phase is effectively reduced by the capping
60 agents, and thus leads to interfacial thermal conductance
61 enhancement. Furthermore, this effect is closely related to the
62 capping agent coverage on the metal surfaces and the type of solvent
63 molecules, and is affected by the models used in the simulations.
64
65 \end{abstract}
66
67 \newpage
68
69 %\narrowtext
70
71 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72 % BODY OF TEXT
73 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74
75 \section{Introduction}
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 Experimentally, various interfaces have been investigated for their
86 thermal conductance. Cahill and coworkers studied nanoscale thermal
87 transport from metal nanoparticle/fluid interfaces, to epitaxial
88 TiN/single crystal oxides interfaces, to hydrophilic and hydrophobic
89 interfaces between water and solids with different self-assembled
90 monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
91 Wang {\it et al.} studied heat transport through
92 long-chain hydrocarbon monolayers on gold substrate at individual
93 molecular level,\cite{Wang10082007} Schmidt {\it et al.} studied the
94 role of CTAB on thermal transport between gold nanorods and
95 solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it et al.} studied
96 the cooling dynamics, which is controlled by thermal interface
97 resistence of glass-embedded metal
98 nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
99 normally considered barriers for heat transport, Alper {\it et al.}
100 suggested that specific ligands (capping agents) could completely
101 eliminate this barrier
102 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
103
104 Theoretical and computational models have also been used to study the
105 interfacial thermal transport in order to gain an understanding of
106 this phenomena at the molecular level. Recently, Hase and coworkers
107 employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
108 study thermal transport from hot Au(111) substrate to a self-assembled
109 monolayer of alkylthiol with relatively long chain (8-20 carbon
110 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
111 measurements for heat conductance of interfaces between the capping
112 monolayer on Au and a solvent phase have yet to be studied with their
113 approach. The comparatively low thermal flux through interfaces is
114 difficult to measure with Equilibrium
115 MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
116 methods. Therefore, the Reverse NEMD (RNEMD)
117 methods\cite{MullerPlathe:1997xw,kuang:164101} would have the
118 advantage of applying this difficult to measure flux (while measuring
119 the resulting gradient), given that the simulation methods being able
120 to effectively apply an unphysical flux in non-homogeneous systems.
121 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
122 this approach to various liquid interfaces and studied how thermal
123 conductance (or resistance) is dependent on chemistry details of
124 interfaces, e.g. hydrophobic and hydrophilic aqueous interfaces.
125
126 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
127 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
128 retains the desirable features of RNEMD (conservation of linear
129 momentum and total energy, compatibility with periodic boundary
130 conditions) while establishing true thermal distributions in each of
131 the two slabs. Furthermore, it allows effective thermal exchange
132 between particles of different identities, and thus makes the study of
133 interfacial conductance much simpler.
134
135 The work presented here deals with the Au(111) surface covered to
136 varying degrees by butanethiol, a capping agent with short carbon
137 chain, and solvated with organic solvents of different molecular
138 properties. Different models were used for both the capping agent and
139 the solvent force field parameters. Using the NIVS algorithm, the
140 thermal transport across these interfaces was studied and the
141 underlying mechanism for the phenomena was investigated.
142
143 \section{Methodology}
144 \subsection{Imposd-Flux Methods in MD Simulations}
145 Steady state MD simulations have an advantage in that not many
146 trajectories are needed to study the relationship between thermal flux
147 and thermal gradients. For systems with low interfacial conductance,
148 one must have a method capable of generating or measuring relatively
149 small fluxes, compared to those required for bulk conductivity. This
150 requirement makes the calculation even more difficult for
151 slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
152 NEMD methods impose a gradient (and measure a flux), but at interfaces
153 it is not clear what behavior should be imposed at the boundaries
154 between materials. Imposed-flux reverse non-equilibrium
155 methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
156 the thermal response becomes an easy-to-measure quantity. Although
157 M\"{u}ller-Plathe's original momentum swapping approach can be used
158 for exchanging energy between particles of different identity, the
159 kinetic energy transfer efficiency is affected by the mass difference
160 between the particles, which limits its application on heterogeneous
161 interfacial systems.
162
163 The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
164 to non-equilibrium MD simulations is able to impose a wide range of
165 kinetic energy fluxes without obvious perturbation to the velocity
166 distributions of the simulated systems. Furthermore, this approach has
167 the advantage in heterogeneous interfaces in that kinetic energy flux
168 can be applied between regions of particles of arbitary identity, and
169 the flux will not be restricted by difference in particle mass.
170
171 The NIVS algorithm scales the velocity vectors in two separate regions
172 of a simulation system with respective diagonal scaling matricies. To
173 determine these scaling factors in the matricies, a set of equations
174 including linear momentum conservation and kinetic energy conservation
175 constraints and target energy flux satisfaction is solved. With the
176 scaling operation applied to the system in a set frequency, bulk
177 temperature gradients can be easily established, and these can be used
178 for computing thermal conductivities. The NIVS algorithm conserves
179 momenta and energy and does not depend on an external thermostat.
180
181 \subsection{Defining Interfacial Thermal Conductivity ($G$)}
182
183 For an interface with relatively low interfacial conductance, and a
184 thermal flux between two distinct bulk regions, the regions on either
185 side of the interface rapidly come to a state in which the two phases
186 have relatively homogeneous (but distinct) temperatures. The
187 interfacial thermal conductivity $G$ can therefore be approximated as:
188 \begin{equation}
189 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
190 \langle T_\mathrm{cold}\rangle \right)}
191 \label{lowG}
192 \end{equation}
193 where ${E_{total}}$ is the total imposed non-physical kinetic energy
194 transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
195 and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
196 temperature of the two separated phases.
197
198 When the interfacial conductance is {\it not} small, there are two
199 ways to define $G$. One common way is to assume the temperature is
200 discrete on the two sides of the interface. $G$ can be calculated
201 using the applied thermal flux $J$ and the maximum temperature
202 difference measured along the thermal gradient max($\Delta T$), which
203 occurs at the Gibbs deviding surface (Figure \ref{demoPic}). This is
204 known as the Kapitza conductance, which is the inverse of the Kapitza
205 resistance.
206 \begin{equation}
207 G=\frac{J}{\Delta T}
208 \label{discreteG}
209 \end{equation}
210
211 \begin{figure}
212 \includegraphics[width=\linewidth]{method}
213 \caption{Interfacial conductance can be calculated by applying an
214 (unphysical) kinetic energy flux between two slabs, one located
215 within the metal and another on the edge of the periodic box. The
216 system responds by forming a thermal response or a gradient. In
217 bulk liquids, this gradient typically has a single slope, but in
218 interfacial systems, there are distinct thermal conductivity
219 domains. The interfacial conductance, $G$ is found by measuring the
220 temperature gap at the Gibbs dividing surface, or by using second
221 derivatives of the thermal profile.}
222 \label{demoPic}
223 \end{figure}
224
225 The other approach is to assume a continuous temperature profile along
226 the thermal gradient axis (e.g. $z$) and define $G$ at the point where
227 the magnitude of thermal conductivity ($\lambda$) change reaches its
228 maximum, given that $\lambda$ is well-defined throughout the space:
229 \begin{equation}
230 G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
231 = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
232 \left(\frac{\partial T}{\partial z}\right)\right)\Big|
233 = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
234 \Big/\left(\frac{\partial T}{\partial z}\right)^2
235 \label{derivativeG}
236 \end{equation}
237
238 With temperature profiles obtained from simulation, one is able to
239 approximate the first and second derivatives of $T$ with finite
240 difference methods and calculate $G^\prime$. In what follows, both
241 definitions have been used, and are compared in the results.
242
243 To investigate the interfacial conductivity at metal / solvent
244 interfaces, we have modeled a metal slab with its (111) surfaces
245 perpendicular to the $z$-axis of our simulation cells. The metal slab
246 has been prepared both with and without capping agents on the exposed
247 surface, and has been solvated with simple organic solvents, as
248 illustrated in Figure \ref{gradT}.
249
250 With the simulation cell described above, we are able to equilibrate
251 the system and impose an unphysical thermal flux between the liquid
252 and the metal phase using the NIVS algorithm. By periodically applying
253 the unphysical flux, we obtained a temperature profile and its spatial
254 derivatives. Figure \ref{gradT} shows how an applied thermal flux can
255 be used to obtain the 1st and 2nd derivatives of the temperature
256 profile.
257
258 \begin{figure}
259 \includegraphics[width=\linewidth]{gradT}
260 \caption{A sample of Au-butanethiol/hexane interfacial system and the
261 temperature profile after a kinetic energy flux is imposed to
262 it. The 1st and 2nd derivatives of the temperature profile can be
263 obtained with finite difference approximation (lower panel).}
264 \label{gradT}
265 \end{figure}
266
267 \section{Computational Details}
268 \subsection{Simulation Protocol}
269 The NIVS algorithm has been implemented in our MD simulation code,
270 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
271 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
272 under atmospheric pressure (1 atm) and 200K. After equilibration,
273 butanethiol capping agents were placed at three-fold hollow sites on
274 the Au(111) surfaces. These sites are either {\it fcc} or {\it
275 hcp} sites, although Hase {\it et al.} found that they are
276 equivalent in a heat transfer process,\cite{hase:2010} so we did not
277 distinguish between these sites in our study. The maximum butanethiol
278 capacity on Au surface is $1/3$ of the total number of surface Au
279 atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
280 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
281 series of lower coverages was also prepared by eliminating
282 butanethiols from the higher coverage surface in a regular manner. The
283 lower coverages were prepared in order to study the relation between
284 coverage and interfacial conductance.
285
286 The capping agent molecules were allowed to migrate during the
287 simulations. They distributed themselves uniformly and sampled a
288 number of three-fold sites throughout out study. Therefore, the
289 initial configuration does not noticeably affect the sampling of a
290 variety of configurations of the same coverage, and the final
291 conductance measurement would be an average effect of these
292 configurations explored in the simulations.
293
294 After the modified Au-butanethiol surface systems were equilibrated in
295 the canonical (NVT) ensemble, organic solvent molecules were packed in
296 the previously empty part of the simulation cells.\cite{packmol} Two
297 solvents were investigated, one which has little vibrational overlap
298 with the alkanethiol and which has a planar shape (toluene), and one
299 which has similar vibrational frequencies to the capping agent and
300 chain-like shape ({\it n}-hexane).
301
302 The simulation cells were not particularly extensive along the
303 $z$-axis, as a very long length scale for the thermal gradient may
304 cause excessively hot or cold temperatures in the middle of the
305 solvent region and lead to undesired phenomena such as solvent boiling
306 or freezing when a thermal flux is applied. Conversely, too few
307 solvent molecules would change the normal behavior of the liquid
308 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
309 these extreme cases did not happen to our simulations. The spacing
310 between periodic images of the gold interfaces is $45 \sim 75$\AA.
311
312 The initial configurations generated are further equilibrated with the
313 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
314 change. This is to ensure that the equilibration of liquid phase does
315 not affect the metal's crystalline structure. Comparisons were made
316 with simulations that allowed changes of $L_x$ and $L_y$ during NPT
317 equilibration. No substantial changes in the box geometry were noticed
318 in these simulations. After ensuring the liquid phase reaches
319 equilibrium at atmospheric pressure (1 atm), further equilibration was
320 carried out under canonical (NVT) and microcanonical (NVE) ensembles.
321
322 After the systems reach equilibrium, NIVS was used to impose an
323 unphysical thermal flux between the metal and the liquid phases. Most
324 of our simulations were done under an average temperature of
325 $\sim$200K. Therefore, thermal flux usually came from the metal to the
326 liquid so that the liquid has a higher temperature and would not
327 freeze due to lowered temperatures. After this induced temperature
328 gradient had stablized, the temperature profile of the simulation cell
329 was recorded. To do this, the simulation cell is devided evenly into
330 $N$ slabs along the $z$-axis. The average temperatures of each slab
331 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
332 the same, the derivatives of $T$ with respect to slab number $n$ can
333 be directly used for $G^\prime$ calculations: \begin{equation}
334 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
335 \Big/\left(\frac{\partial T}{\partial z}\right)^2
336 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
337 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
338 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
339 \Big/\left(\frac{\partial T}{\partial n}\right)^2
340 \label{derivativeG2}
341 \end{equation}
342
343 All of the above simulation procedures use a time step of 1 fs. Each
344 equilibration stage took a minimum of 100 ps, although in some cases,
345 longer equilibration stages were utilized.
346
347 \subsection{Force Field Parameters}
348 Our simulations include a number of chemically distinct components.
349 Figure \ref{demoMol} demonstrates the sites defined for both
350 United-Atom and All-Atom models of the organic solvent and capping
351 agents in our simulations. Force field parameters are needed for
352 interactions both between the same type of particles and between
353 particles of different species.
354
355 \begin{figure}
356 \includegraphics[width=\linewidth]{structures}
357 \caption{Structures of the capping agent and solvents utilized in
358 these simulations. The chemically-distinct sites (a-e) are expanded
359 in terms of constituent atoms for both United Atom (UA) and All Atom
360 (AA) force fields. Most parameters are from
361 Refs. \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
362 (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
363 atoms are given in Table \ref{MnM}.}
364 \label{demoMol}
365 \end{figure}
366
367 The Au-Au interactions in metal lattice slab is described by the
368 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
369 potentials include zero-point quantum corrections and are
370 reparametrized for accurate surface energies compared to the
371 Sutton-Chen potentials.\cite{Chen90}
372
373 For the two solvent molecules, {\it n}-hexane and toluene, two
374 different atomistic models were utilized. Both solvents were modeled
375 using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
376 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
377 for our UA solvent molecules. In these models, sites are located at
378 the carbon centers for alkyl groups. Bonding interactions, including
379 bond stretches and bends and torsions, were used for intra-molecular
380 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
381 potentials are used.
382
383 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
384 simple and computationally efficient, while maintaining good accuracy.
385 However, the TraPPE-UA model for alkanes is known to predict a slighly
386 lower boiling point than experimental values. This is one of the
387 reasons we used a lower average temperature (200K) for our
388 simulations. If heat is transferred to the liquid phase during the
389 NIVS simulation, the liquid in the hot slab can actually be
390 substantially warmer than the mean temperature in the simulation. The
391 lower mean temperatures therefore prevent solvent boiling.
392
393 For UA-toluene, the non-bonded potentials between intermolecular sites
394 have a similar Lennard-Jones formulation. The toluene molecules were
395 treated as a single rigid body, so there was no need for
396 intramolecular interactions (including bonds, bends, or torsions) in
397 this solvent model.
398
399 Besides the TraPPE-UA models, AA models for both organic solvents are
400 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
401 were used. For hexane, additional explicit hydrogen sites were
402 included. Besides bonding and non-bonded site-site interactions,
403 partial charges and the electrostatic interactions were added to each
404 CT and HC site. For toluene, a flexible model for the toluene molecule
405 was utilized which included bond, bend, torsion, and inversion
406 potentials to enforce ring planarity.
407
408 The butanethiol capping agent in our simulations, were also modeled
409 with both UA and AA model. The TraPPE-UA force field includes
410 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
411 UA butanethiol model in our simulations. The OPLS-AA also provides
412 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
413 surfaces do not have the hydrogen atom bonded to sulfur. To derive
414 suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
415 adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
416 modify the parameters for the CTS atom to maintain charge neutrality
417 in the molecule. Note that the model choice (UA or AA) for the capping
418 agent can be different from the solvent. Regardless of model choice,
419 the force field parameters for interactions between capping agent and
420 solvent can be derived using Lorentz-Berthelot Mixing Rule:
421 \begin{eqnarray}
422 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
423 \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
424 \end{eqnarray}
425
426 To describe the interactions between metal (Au) and non-metal atoms,
427 we refer to an adsorption study of alkyl thiols on gold surfaces by
428 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
429 Lennard-Jones form of potential parameters for the interaction between
430 Au and pseudo-atoms CH$_x$ and S based on a well-established and
431 widely-used effective potential of Hautman and Klein for the Au(111)
432 surface.\cite{hautman:4994} As our simulations require the gold slab
433 to be flexible to accommodate thermal excitation, the pair-wise form
434 of potentials they developed was used for our study.
435
436 The potentials developed from {\it ab initio} calculations by Leng
437 {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
438 interactions between Au and aromatic C/H atoms in toluene. However,
439 the Lennard-Jones parameters between Au and other types of particles,
440 (e.g. AA alkanes) have not yet been established. For these
441 interactions, the Lorentz-Berthelot mixing rule can be used to derive
442 effective single-atom LJ parameters for the metal using the fit values
443 for toluene. These are then used to construct reasonable mixing
444 parameters for the interactions between the gold and other atoms.
445 Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
446 our simulations.
447
448 \begin{table*}
449 \begin{minipage}{\linewidth}
450 \begin{center}
451 \caption{Non-bonded interaction parameters (including cross
452 interactions with Au atoms) for both force fields used in this
453 work.}
454 \begin{tabular}{lllllll}
455 \hline\hline
456 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
457 $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
458 & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
459 \hline
460 United Atom (UA)
461 &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
462 &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
463 &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
464 &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
465 \hline
466 All Atom (AA)
467 &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
468 &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
469 &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
470 &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
471 &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
472 &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
473 \hline
474 Both UA and AA
475 & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
476 \hline\hline
477 \end{tabular}
478 \label{MnM}
479 \end{center}
480 \end{minipage}
481 \end{table*}
482
483
484 \section{Results}
485 There are many factors contributing to the measured interfacial
486 conductance; some of these factors are physically motivated
487 (e.g. coverage of the surface by the capping agent coverage and
488 solvent identity), while some are governed by parameters of the
489 methodology (e.g. applied flux and the formulas used to obtain the
490 conductance). In this section we discuss the major physical and
491 calculational effects on the computed conductivity.
492
493 \subsection{Effects due to capping agent coverage}
494
495 A series of different initial conditions with a range of surface
496 coverages was prepared and solvated with various with both of the
497 solvent molecules. These systems were then equilibrated and their
498 interfacial thermal conductivity was measured with the NIVS
499 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
500 with respect to surface coverage.
501
502 \begin{figure}
503 \includegraphics[width=\linewidth]{coverage}
504 \caption{Comparison of interfacial thermal conductivity ($G$) values
505 for the Au-butanethiol/solvent interface with various UA models and
506 different capping agent coverages at $\langle T\rangle\sim$200K.}
507 \label{coverage}
508 \end{figure}
509
510 In partially covered surfaces, the derivative definition for $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the location of maximum change of $\lambda$ becomes washed out. The discrete definition (Eq. \ref{discreteG}) is easier to apply, as the Gibbs dividing surface is still well-defined. Therefore, $G$ (not $G^\prime$) was used in this section.
511
512 From Figure \ref{coverage}, one can see the significance of the presence of capping agents. When even a small fraction of the Au(111) surface sites are covered with butanethiols, the conductivity exhibits an enhancement by at least a factor of 3. Cappping agents are clearly playing a major role in thermal transport at metal / organic solvent surfaces.
513
514 We note a non-monotonic behavior in the interfacial conductance as a function of surface coverage. The maximum conductance (largest $G$) happens when the surfaces are about 75\% covered with butanethiol caps. The reason for this behavior is not entirely clear. One explanation is that incomplete butanethiol coverage allows small gaps between butanethiols to form. These gaps can be filled by transient solvent molecules. These solvent molecules couple very strongly with the hot capping agent molecules near the surface, and can then carry away (diffusively) the excess thermal energy from the surface.
515
516 There appears to be a competition between the conduction of the thermal energy away from the surface by the capping agents (enhanced by greater coverage) and the coupling of the capping agents with the solvent (enhanced by interdigitation at lower coverages). This competition would lead to the non-monotonic coverage behavior observed here.
517
518 Results for rigid body toluene solvent, as well as the UA hexane, are within the ranges expected from prior experimental work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests that explicit hydrogen atoms might not be required for modeling thermal transport in these systems. C-H vibrational modes do not see significant excited state population at low temperatures, and are not likely to carry lower frequency excitations from the solid layer into the bulk liquid.
519
520 The toluene solvent does not exhibit the same behavior as hexane in that $G$ remains at approximately the same magnitude when the capping coverage increases from 25\% to 75\%. Toluene, as a rigid planar molecule, cannot occupy the relatively small gaps between the capping agents as easily as the chain-like {\it n}-hexane. The effect of solvent coupling to the capping agent is therefore weaker in toluene except at the very lowest coverage levels. This effect counters the coverage-dependent conduction of heat away from the metal surface, leading to a much flatter $G$ vs. coverage trend than is observed in {\it n}-hexane.
521
522 \subsection{Effects due to Solvent \& Solvent Models}
523 In addition to UA solvent and capping agent models, AA models have also been included in our simulations. In most of this work, the same (UA or AA) model for solvent and capping agent was used, but it is also possible to utilize different models for different components. We have also included isotopic substitutions (Hydrogen to Deuterium) to decrease the explicit vibrational overlap between solvent and capping agent. Table \ref{modelTest} summarizes the results of these studies.
524
525 \begin{table*}
526 \begin{minipage}{\linewidth}
527 \begin{center}
528
529 \caption{Computed interfacial thermal conductance ($G$ and
530 $G^\prime$) values for interfaces using various models for
531 solvent and capping agent (or without capping agent) at
532 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
533 or capping agent molecules; ``Avg.'' denotes results that are
534 averages of simulations under different applied thermal flux values $(J_z)$. Error
535 estimates are indicated in parentheses.)}
536
537 \begin{tabular}{llccc}
538 \hline\hline
539 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
540 (or bare surface) & model & (GW/m$^2$) &
541 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
542 \hline
543 UA & UA hexane & Avg. & 131(9) & 87(10) \\
544 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
545 & AA hexane & Avg. & 131(6) & 122(10) \\
546 & UA toluene & 1.96 & 187(16) & 151(11) \\
547 & AA toluene & 1.89 & 200(36) & 149(53) \\
548 \hline
549 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
550 & AA hexane & Avg. & 442(14) & 356(31) \\
551 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
552 & UA toluene & 1.98 & 125(25) & 97(60) \\
553 & AA toluene & 3.79 & 487(56) & 290(42) \\
554 \hline
555 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
556 & AA hexane & 1.92 & 243(29) & 191(11) \\
557 & AA toluene & 1.93 & 364(36) & 322(67) \\
558 \hline
559 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
560 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
561 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
562 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
563 \hline\hline
564 \end{tabular}
565 \label{modelTest}
566 \end{center}
567 \end{minipage}
568 \end{table*}
569
570 To facilitate direct comparison between force fields, systems with the same capping agent and solvent were prepared with the same length scales for the simulation cells.
571
572 On bare metal / solvent surfaces, different force field models for hexane yield similar results for both $G$ and $G^\prime$, and these two definitions agree with each other very well. This is primarily an indicator of weak interactions between the metal and the solvent, and is a typical case for acoustic impedance mismatch between these two phases.
573
574 For the fully-covered surfaces, the choice of force field for the capping agent and solvent has a large impact on the calulated values of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are much larger than their UA to UA counterparts, and these values exceed the experimental estimates by a large measure. The AA force field allows significant energy to go into C-H (or C-D) stretching modes, and since these modes are high frequency, this non-quantum behavior is likely responsible for the overestimate of the conductivity.
575
576 The similarity in the vibrational modes available to solvent and capping agent can be reduced by deuterating one of the two components. Once either the hexanes or the butanethiols are deuterated, one can see a significantly lower $G$ and $G^\prime$ (Figure \ref{aahxntln}). Compared to the AA model, the UA model yields more reasonable conductivity values with much higher computational efficiency.
577
578 \begin{figure}
579 \includegraphics[width=\linewidth]{aahxntln}
580 \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
581 systems. When butanethiol is deuterated (lower left), its
582 vibrational overlap with hexane decreases significantly. Since aromatic molecules and the butanethiol are vibrationally dissimilar, the change is not as dramatic when toluene is the solvent (right).}
583 \label{aahxntln}
584 \end{figure}
585
586 For the Au / butanethiol / toluene interfaces, having the AA butanethiol deuterated did not yield a significant change in the measured conductance. Compared to the C-H vibrational overlap between hexane and butanethiol, both of which have alkyl chains, the overlap between toluene and butanethiol is not as significant and thus does not contribute as much to the heat exchange process. The presence of extra degrees of freedom in the AA force field for toluene yields higher heat exchange rates between the two phases and results in a much higher conductivity than in the UA force field.
587
588 \subsubsection{Are electronic excitations in the metal important?}
589 Because they lack electronic excitations, the QSC and related embedded atom method (EAM) models for gold are known to predict unreasonably low values for bulk conductivity ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the conductance between the phases ($G$) is governed primarily by phonon excitation (and not electronic degrees of freedom), one would expect a classical model to capture most of the interfacial thermal conductance. Our results for $G$ and $G^\prime$ indicate that this is indeed the case, and suggest that the modeling of interfacial thermal transport depends primarily on the description of the interactions between the various components at the interface. When the metal is chemically capped, the primary barrier to thermal conductivity appears to be the interface between the capping agent and the surrounding solvent, so the excitations in the metal have little impact on the value of $G$.
590
591 \subsection{Effects due to methodology and simulation parameters}
592
593 START HERE
594
595 We have varied our protocol or other parameters of the simulations in order to investigate how these factors would affect the computation of $G$.
596
597 We allowed $L_x$ and $L_y$ to change during equilibrating the liquid phase. Due to the stiffness of the crystalline Au structure, $L_x$ and $L_y$ would not change noticeably after equilibration. Although $L_z$ could fluctuate $\sim$1\% after a system is fully equilibrated in the NPT ensemble, this fluctuation, as well as those of $L_x$ and $L_y$ (which is significantly smaller), would not be magnified on the calculated $G$'s, as shown in Table \ref{AuThiolHexaneUA}. This insensivity to $L_i$ fluctuations allows reliable measurement of $G$'s without the necessity of extremely cautious equilibration process.
598
599 As stated in our computational details, the spacing filled with
600 solvent molecules can be chosen within a range. This allows some
601 change of solvent molecule numbers for the same Au-butanethiol
602 surfaces. We did this study on our Au-butanethiol/hexane
603 simulations. Nevertheless, the results obtained from systems of
604 different $N_{hexane}$ did not indicate that the measurement of $G$ is
605 susceptible to this parameter. For computational efficiency concern,
606 smaller system size would be preferable, given that the liquid phase
607 structure is not affected.
608
609 \subsubsection{Effects of applied flux}
610 Our NIVS algorithm allows change of unphysical thermal flux both in
611 direction and in quantity. This feature extends our investigation of
612 interfacial thermal conductance. However, the magnitude of this
613 thermal flux is not arbitary if one aims to obtain a stable and
614 reliable thermal gradient. A temperature profile would be
615 substantially affected by noise when $|J_z|$ has a much too low
616 magnitude; while an excessively large $|J_z|$ that overwhelms the
617 conductance capacity of the interface would prevent a thermal gradient
618 to reach a stablized steady state. NIVS has the advantage of allowing
619 $J$ to vary in a wide range such that the optimal flux range for $G$
620 measurement can generally be simulated by the algorithm. Within the
621 optimal range, we were able to study how $G$ would change according to
622 the thermal flux across the interface. For our simulations, we denote
623 $J_z$ to be positive when the physical thermal flux is from the liquid
624 to metal, and negative vice versa. The $G$'s measured under different
625 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and
626 \ref{AuThiolToluene}. These results do not suggest that $G$ is
627 dependent on $J_z$ within this flux range. The linear response of flux
628 to thermal gradient simplifies our investigations in that we can rely
629 on $G$ measurement with only a couple $J_z$'s and do not need to test
630 a large series of fluxes.
631
632 \begin{table*}
633 \begin{minipage}{\linewidth}
634 \begin{center}
635 \caption{Computed interfacial thermal conductivity ($G$ and
636 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
637 interfaces with UA model and different hexane molecule numbers
638 at different temperatures using a range of energy
639 fluxes. Error estimates indicated in parenthesis.}
640
641 \begin{tabular}{ccccccc}
642 \hline\hline
643 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
644 $J_z$ & $G$ & $G^\prime$ \\
645 (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
646 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
647 \hline
648 200 & 266 & No & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
649 & 200 & Yes & 0.694 & 1.92 & 129(11) & 87.3(0.3) \\
650 & & Yes & 0.672 & 1.93 & 131(16) & 78(13) \\
651 & & No & 0.688 & 0.96 & 125(16) & 90.2(15) \\
652 & & & & 1.91 & 139(10) & 101(10) \\
653 & & & & 2.83 & 141(6) & 89.9(9.8) \\
654 & 166 & Yes & 0.679 & 0.97 & 115(19) & 69(18) \\
655 & & & & 1.94 & 125(9) & 87.1(0.2) \\
656 & & No & 0.681 & 0.97 & 141(30) & 78(22) \\
657 & & & & 1.92 & 138(4) & 98.9(9.5) \\
658 \hline
659 250 & 200 & No & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
660 & & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
661 & 166 & Yes & 0.570 & 0.98 & 79.0(3.5) & 62.9(3.0) \\
662 & & No & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
663 & & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
664 & & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
665 & & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
666 \hline\hline
667 \end{tabular}
668 \label{AuThiolHexaneUA}
669 \end{center}
670 \end{minipage}
671 \end{table*}
672
673 \subsubsection{Effects due to average temperature}
674
675 Furthermore, we also attempted to increase system average temperatures
676 to above 200K. These simulations are first equilibrated in the NPT
677 ensemble under normal pressure. As stated above, the TraPPE-UA model
678 for hexane tends to predict a lower boiling point. In our simulations,
679 hexane had diffculty to remain in liquid phase when NPT equilibration
680 temperature is higher than 250K. Additionally, the equilibrated liquid
681 hexane density under 250K becomes lower than experimental value. This
682 expanded liquid phase leads to lower contact between hexane and
683 butanethiol as well.[MAY NEED SLAB DENSITY FIGURE]
684 And this reduced contact would
685 probably be accountable for a lower interfacial thermal conductance,
686 as shown in Table \ref{AuThiolHexaneUA}.
687
688 A similar study for TraPPE-UA toluene agrees with the above result as
689 well. Having a higher boiling point, toluene tends to remain liquid in
690 our simulations even equilibrated under 300K in NPT
691 ensembles. Furthermore, the expansion of the toluene liquid phase is
692 not as significant as that of the hexane. This prevents severe
693 decrease of liquid-capping agent contact and the results (Table
694 \ref{AuThiolToluene}) show only a slightly decreased interface
695 conductance. Therefore, solvent-capping agent contact should play an
696 important role in the thermal transport process across the interface
697 in that higher degree of contact could yield increased conductance.
698
699 \begin{table*}
700 \begin{minipage}{\linewidth}
701 \begin{center}
702 \caption{Computed interfacial thermal conductivity ($G$ and
703 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
704 interface at different temperatures using a range of energy
705 fluxes. Error estimates indicated in parenthesis.}
706
707 \begin{tabular}{ccccc}
708 \hline\hline
709 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
710 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
711 \hline
712 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
713 & & -1.86 & 180(3) & 135(21) \\
714 & & -3.93 & 176(5) & 113(12) \\
715 \hline
716 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
717 & & -4.19 & 135(9) & 113(12) \\
718 \hline\hline
719 \end{tabular}
720 \label{AuThiolToluene}
721 \end{center}
722 \end{minipage}
723 \end{table*}
724
725 Besides lower interfacial thermal conductance, surfaces in relatively
726 high temperatures are susceptible to reconstructions, when
727 butanethiols have a full coverage on the Au(111) surface. These
728 reconstructions include surface Au atoms migrated outward to the S
729 atom layer, and butanethiol molecules embedded into the original
730 surface Au layer. The driving force for this behavior is the strong
731 Au-S interactions in our simulations. And these reconstructions lead
732 to higher ratio of Au-S attraction and thus is energetically
733 favorable. Furthermore, this phenomenon agrees with experimental
734 results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
735 {\it et al.} had kept their Au(111) slab rigid so that their
736 simulations can reach 300K without surface reconstructions. Without
737 this practice, simulating 100\% thiol covered interfaces under higher
738 temperatures could hardly avoid surface reconstructions. However, our
739 measurement is based on assuming homogeneity on $x$ and $y$ dimensions
740 so that measurement of $T$ at particular $z$ would be an effective
741 average of the particles of the same type. Since surface
742 reconstructions could eliminate the original $x$ and $y$ dimensional
743 homogeneity, measurement of $G$ is more difficult to conduct under
744 higher temperatures. Therefore, most of our measurements are
745 undertaken at $\langle T\rangle\sim$200K.
746
747 However, when the surface is not completely covered by butanethiols,
748 the simulated system is more resistent to the reconstruction
749 above. Our Au-butanethiol/toluene system had the Au(111) surfaces 90\%
750 covered by butanethiols, but did not see this above phenomena even at
751 $\langle T\rangle\sim$300K. The empty three-fold sites not occupied by
752 capping agents could help prevent surface reconstruction in that they
753 provide other means of capping agent relaxation. It is observed that
754 butanethiols can migrate to their neighbor empty sites during a
755 simulation. Therefore, we were able to obtain $G$'s for these
756 interfaces even at a relatively high temperature without being
757 affected by surface reconstructions.
758
759
760 \section{Discussion}
761
762 \subsection{Capping agent acts as a vibrational coupler between solid
763 and solvent phases}
764 To investigate the mechanism of interfacial thermal conductance, the
765 vibrational power spectrum was computed. Power spectra were taken for
766 individual components in different simulations. To obtain these
767 spectra, simulations were run after equilibration, in the NVE
768 ensemble, and without a thermal gradient. Snapshots of configurations
769 were collected at a frequency that is higher than that of the fastest
770 vibrations occuring in the simulations. With these configurations, the
771 velocity auto-correlation functions can be computed:
772 \begin{equation}
773 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
774 \label{vCorr}
775 \end{equation}
776 The power spectrum is constructed via a Fourier transform of the
777 symmetrized velocity autocorrelation function,
778 \begin{equation}
779 \hat{f}(\omega) =
780 \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
781 \label{fourier}
782 \end{equation}
783
784
785 \subsubsection{The role of specific vibrations}
786 The vibrational spectra for gold slabs in different environments are
787 shown as in Figure \ref{specAu}. Regardless of the presence of
788 solvent, the gold surfaces covered by butanethiol molecules, compared
789 to bare gold surfaces, exhibit an additional peak observed at the
790 frequency of $\sim$170cm$^{-1}$, which is attributed to the S-Au
791 bonding vibration. This vibration enables efficient thermal transport
792 from surface Au layer to the capping agents. Therefore, in our
793 simulations, the Au/S interfaces do not appear major heat barriers
794 compared to the butanethiol / solvent interfaces.
795
796 \subsubsection{Overlap of power spectrum}
797 Simultaneously, the vibrational overlap between butanethiol and
798 organic solvents suggests higher thermal exchange efficiency between
799 these two components. Even exessively high heat transport was observed
800 when All-Atom models were used and C-H vibrations were treated
801 classically. Compared to metal and organic liquid phase, the heat
802 transfer efficiency between butanethiol and organic solvents is closer
803 to that within bulk liquid phase.
804
805 Furthermore, our observation validated previous
806 results\cite{hase:2010} that the intramolecular heat transport of
807 alkylthiols is highly effecient. As a combinational effects of these
808 phenomena, butanethiol acts as a channel to expedite thermal transport
809 process. The acoustic impedance mismatch between the metal and the
810 liquid phase can be effectively reduced with the presence of suitable
811 capping agents.
812
813 \begin{figure}
814 \includegraphics[width=\linewidth]{vibration}
815 \caption{Vibrational spectra obtained for gold in different
816 environments.}
817 \label{specAu}
818 \end{figure}
819
820 \subsubsection{Isotopic substitution and vibrational overlap}
821 A comparison of the results obtained from the two different organic
822 solvents can also provide useful information of the interfacial
823 thermal transport process. The deuterated hexane (UA) results do not
824 appear to be substantially different from those of normal hexane (UA),
825 given that butanethiol (UA) is non-deuterated for both solvents. The
826 UA models, even though they have eliminated C-H vibrational overlap,
827 still have significant overlap in the infrared spectra. Because
828 differences in the infrared range do not seem to produce an observable
829 difference for the results of $G$ (Figure \ref{uahxnua}).
830
831 \begin{figure}
832 \includegraphics[width=\linewidth]{uahxnua}
833 \caption{Vibrational spectra obtained for normal (upper) and
834 deuterated (lower) hexane in Au-butanethiol/hexane
835 systems. Butanethiol spectra are shown as reference. Both hexane and
836 butanethiol were using United-Atom models.}
837 \label{uahxnua}
838 \end{figure}
839
840 \section{Conclusions}
841 The NIVS algorithm we developed has been applied to simulations of
842 Au-butanethiol surfaces with organic solvents. This algorithm allows
843 effective unphysical thermal flux transferred between the metal and
844 the liquid phase. With the flux applied, we were able to measure the
845 corresponding thermal gradient and to obtain interfacial thermal
846 conductivities. Under steady states, single trajectory simulation
847 would be enough for accurate measurement. This would be advantageous
848 compared to transient state simulations, which need multiple
849 trajectories to produce reliable average results.
850
851 Our simulations have seen significant conductance enhancement with the
852 presence of capping agent, compared to the bare gold / liquid
853 interfaces. The acoustic impedance mismatch between the metal and the
854 liquid phase is effectively eliminated by proper capping
855 agent. Furthermore, the coverage precentage of the capping agent plays
856 an important role in the interfacial thermal transport
857 process. Moderately lower coverages allow higher contact between
858 capping agent and solvent, and thus could further enhance the heat
859 transfer process.
860
861 Our measurement results, particularly of the UA models, agree with
862 available experimental data. This indicates that our force field
863 parameters have a nice description of the interactions between the
864 particles at the interfaces. AA models tend to overestimate the
865 interfacial thermal conductance in that the classically treated C-H
866 vibration would be overly sampled. Compared to the AA models, the UA
867 models have higher computational efficiency with satisfactory
868 accuracy, and thus are preferable in interfacial thermal transport
869 modelings. Of the two definitions for $G$, the discrete form
870 (Eq. \ref{discreteG}) was easier to use and gives out relatively
871 consistent results, while the derivative form (Eq. \ref{derivativeG})
872 is not as versatile. Although $G^\prime$ gives out comparable results
873 and follows similar trend with $G$ when measuring close to fully
874 covered or bare surfaces, the spatial resolution of $T$ profile is
875 limited for accurate computation of derivatives data.
876
877 Vlugt {\it et al.} has investigated the surface thiol structures for
878 nanocrystal gold and pointed out that they differs from those of the
879 Au(111) surface\cite{landman:1998,vlugt:cpc2007154}. This difference
880 might lead to change of interfacial thermal transport behavior as
881 well. To investigate this problem, an effective means to introduce
882 thermal flux and measure the corresponding thermal gradient is
883 desirable for simulating structures with spherical symmetry.
884
885 \section{Acknowledgments}
886 Support for this project was provided by the National Science
887 Foundation under grant CHE-0848243. Computational time was provided by
888 the Center for Research Computing (CRC) at the University of Notre
889 Dame.
890 \newpage
891
892 \bibliography{interfacial}
893
894 \end{doublespace}
895 \end{document}
896