ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3751
Committed: Tue Jul 26 19:43:10 2011 UTC (12 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 49772 byte(s)
Log Message:
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. 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 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
101 this phenomena at the molecular level. Recently, Hase and coworkers
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
106 measurements for heat conductance of interfaces between the capping
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,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 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
124 conditions) while establishing true thermal distributions in each of
125 the two slabs. Furthermore, it allows effective thermal exchange
126 between particles of different identities, and thus makes the study of
127 interfacial conductance much simpler.
128
129 The work presented here deals with the Au(111) surface covered to
130 varying degrees by butanethiol, a capping agent with short carbon
131 chain, and solvated with organic solvents of different molecular
132 properties. Different models were used for both the capping agent and
133 the solvent force field parameters. Using the NIVS algorithm, the
134 thermal transport across these interfaces was studied and the
135 underlying mechanism for the phenomena was investigated.
136
137 \section{Methodology}
138 \subsection{Imposd-Flux Methods in MD Simulations}
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 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
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
162 can be applied between regions of particles of arbitary identity, and
163 the flux will not be restricted by difference in particle mass.
164
165 The NIVS algorithm scales the velocity vectors in two separate regions
166 of a simulation system with respective diagonal scaling matricies. To
167 determine these scaling factors in the matricies, a set of equations
168 including linear momentum conservation and kinetic energy conservation
169 constraints and target energy flux satisfaction is solved. With the
170 scaling operation applied to the system in a set frequency, bulk
171 temperature gradients can be easily established, and these can be used
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
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 -
184 \langle T_\mathrm{cold}\rangle \right)}
185 \label{lowG}
186 \end{equation}
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 $G$. 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 \begin{figure}
201 \includegraphics[width=\linewidth]{method}
202 \caption{Interfacial conductance can be calculated by applying an
203 (unphysical) kinetic energy flux between two slabs, one located
204 within the metal and another on the edge of the periodic box. The
205 system responds by forming a thermal response or a gradient. In
206 bulk liquids, this gradient typically has a single slope, but in
207 interfacial systems, there are distinct thermal conductivity
208 domains. The interfacial conductance, $G$ is found by measuring the
209 temperature gap at the Gibbs dividing surface, or by using second
210 derivatives of the thermal profile.}
211 \label{demoPic}
212 \end{figure}
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 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|
220 = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
221 \left(\frac{\partial T}{\partial z}\right)\right)\Big|
222 = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
223 \Big/\left(\frac{\partial T}{\partial z}\right)^2
224 \label{derivativeG}
225 \end{equation}
226
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 calculate $G^\prime$. In what follows, both
230 definitions have been used, and are compared in the results.
231
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 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}
249 \caption{A sample of Au-butanethiol/hexane interfacial system and the
250 temperature profile after a kinetic energy flux is imposed to
251 it. The 1st and 2nd derivatives of the temperature profile can be
252 obtained with finite difference approximation (lower panel).}
253 \label{gradT}
254 \end{figure}
255
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 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 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 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.
282
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 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 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. 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 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 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 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: \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
327 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
328 \Big/\left(\frac{\partial T}{\partial n}\right)^2
329 \label{derivativeG2}
330 \end{equation}
331
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 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
344 \begin{figure}
345 \includegraphics[width=\linewidth]{structures}
346 \caption{Structures of the capping agent and solvents utilized in
347 these simulations. The chemically-distinct sites (a-e) are expanded
348 in terms of constituent atoms for both United Atom (UA) and All Atom
349 (AA) force fields. Most parameters are from
350 Refs. \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} (UA) and
351 \protect\cite{OPLSAA} (AA). Cross-interactions with the Au atoms are given
352 in Table \ref{MnM}.}
353 \label{demoMol}
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
358 potentials include zero-point quantum corrections and are
359 reparametrized for accurate surface energies compared to the
360 Sutton-Chen potentials.\cite{Chen90}
361
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 closer than 3 bonds. For non-bonded interactions, Lennard-Jones
370 potentials are used.
371
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, 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, 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} 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 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 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}}
415 \end{eqnarray}
416
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 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}
441 \begin{center}
442 \caption{Non-bonded interaction parameters (including cross
443 interactions with Au atoms) for both force fields used in this
444 work.}
445 \begin{tabular}{lllllll}
446 \hline\hline
447 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
448 $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
449 & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
450 \hline
451 United Atom (UA)
452 &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
453 &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
454 &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
455 &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
456 \hline
457 All Atom (AA)
458 &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
459 &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
460 &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
461 &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
462 &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
463 &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
464 \hline
465 Both UA and AA
466 & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
467 \hline\hline
468 \end{tabular}
469 \label{MnM}
470 \end{center}
471 \end{minipage}
472 \end{table*}
473
474 \subsection{Vibrational Power Spectrum}
475
476 To investigate the mechanism of interfacial thermal conductance, the
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 The power spectrum is constructed via a Fourier transform of the
489 symmetrized velocity autocorrelation function,
490 \begin{equation}
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}
497 In what follows, how the parameters and protocol of simulations would
498 affect the measurement of $G$'s is first discussed. With a reliable
499 protocol and set of parameters, the influence of capping agent
500 coverage on thermal conductance is investigated. Besides, different
501 force field models for both solvents and selected deuterated models
502 were tested and compared. Finally, a summary of the role of capping
503 agent in the interfacial thermal transport process is given.
504
505 \subsection{How Simulation Parameters Affects $G$}
506 We have varied our protocol or other parameters of the simulations in
507 order to investigate how these factors would affect the measurement of
508 $G$'s. It turned out that while some of these parameters would not
509 affect the results substantially, some other changes to the
510 simulations would have a significant impact on the measurement
511 results.
512
513 In some of our simulations, we allowed $L_x$ and $L_y$ to change
514 during equilibrating the liquid phase. Due to the stiffness of the
515 crystalline Au structure, $L_x$ and $L_y$ would not change noticeably
516 after equilibration. Although $L_z$ could fluctuate $\sim$1\% after a
517 system is fully equilibrated in the NPT ensemble, this fluctuation, as
518 well as those of $L_x$ and $L_y$ (which is significantly smaller),
519 would not be magnified on the calculated $G$'s, as shown in Table
520 \ref{AuThiolHexaneUA}. This insensivity to $L_i$ fluctuations allows
521 reliable measurement of $G$'s without the necessity of extremely
522 cautious equilibration process.
523
524 As stated in our computational details, the spacing filled with
525 solvent molecules can be chosen within a range. This allows some
526 change of solvent molecule numbers for the same Au-butanethiol
527 surfaces. We did this study on our Au-butanethiol/hexane
528 simulations. Nevertheless, the results obtained from systems of
529 different $N_{hexane}$ did not indicate that the measurement of $G$ is
530 susceptible to this parameter. For computational efficiency concern,
531 smaller system size would be preferable, given that the liquid phase
532 structure is not affected.
533
534 Our NIVS algorithm allows change of unphysical thermal flux both in
535 direction and in quantity. This feature extends our investigation of
536 interfacial thermal conductance. However, the magnitude of this
537 thermal flux is not arbitary if one aims to obtain a stable and
538 reliable thermal gradient. A temperature profile would be
539 substantially affected by noise when $|J_z|$ has a much too low
540 magnitude; while an excessively large $|J_z|$ that overwhelms the
541 conductance capacity of the interface would prevent a thermal gradient
542 to reach a stablized steady state. NIVS has the advantage of allowing
543 $J$ to vary in a wide range such that the optimal flux range for $G$
544 measurement can generally be simulated by the algorithm. Within the
545 optimal range, we were able to study how $G$ would change according to
546 the thermal flux across the interface. For our simulations, we denote
547 $J_z$ to be positive when the physical thermal flux is from the liquid
548 to metal, and negative vice versa. The $G$'s measured under different
549 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and
550 \ref{AuThiolToluene}. These results do not suggest that $G$ is
551 dependent on $J_z$ within this flux range. The linear response of flux
552 to thermal gradient simplifies our investigations in that we can rely
553 on $G$ measurement with only a couple $J_z$'s and do not need to test
554 a large series of fluxes.
555
556 \begin{table*}
557 \begin{minipage}{\linewidth}
558 \begin{center}
559 \caption{Computed interfacial thermal conductivity ($G$ and
560 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
561 interfaces with UA model and different hexane molecule numbers
562 at different temperatures using a range of energy
563 fluxes. Error estimates indicated in parenthesis.}
564
565 \begin{tabular}{ccccccc}
566 \hline\hline
567 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
568 $J_z$ & $G$ & $G^\prime$ \\
569 (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
570 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
571 \hline
572 200 & 266 & No & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
573 & 200 & Yes & 0.694 & 1.92 & 129(11) & 87.3(0.3) \\
574 & & Yes & 0.672 & 1.93 & 131(16) & 78(13) \\
575 & & No & 0.688 & 0.96 & 125(16) & 90.2(15) \\
576 & & & & 1.91 & 139(10) & 101(10) \\
577 & & & & 2.83 & 141(6) & 89.9(9.8) \\
578 & 166 & Yes & 0.679 & 0.97 & 115(19) & 69(18) \\
579 & & & & 1.94 & 125(9) & 87.1(0.2) \\
580 & & No & 0.681 & 0.97 & 141(30) & 78(22) \\
581 & & & & 1.92 & 138(4) & 98.9(9.5) \\
582 \hline
583 250 & 200 & No & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
584 & & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
585 & 166 & Yes & 0.570 & 0.98 & 79.0(3.5) & 62.9(3.0) \\
586 & & No & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
587 & & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
588 & & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
589 & & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
590 \hline\hline
591 \end{tabular}
592 \label{AuThiolHexaneUA}
593 \end{center}
594 \end{minipage}
595 \end{table*}
596
597 Furthermore, we also attempted to increase system average temperatures
598 to above 200K. These simulations are first equilibrated in the NPT
599 ensemble under normal pressure. As stated above, the TraPPE-UA model
600 for hexane tends to predict a lower boiling point. In our simulations,
601 hexane had diffculty to remain in liquid phase when NPT equilibration
602 temperature is higher than 250K. Additionally, the equilibrated liquid
603 hexane density under 250K becomes lower than experimental value. This
604 expanded liquid phase leads to lower contact between hexane and
605 butanethiol as well.[MAY NEED SLAB DENSITY FIGURE]
606 And this reduced contact would
607 probably be accountable for a lower interfacial thermal conductance,
608 as shown in Table \ref{AuThiolHexaneUA}.
609
610 A similar study for TraPPE-UA toluene agrees with the above result as
611 well. Having a higher boiling point, toluene tends to remain liquid in
612 our simulations even equilibrated under 300K in NPT
613 ensembles. Furthermore, the expansion of the toluene liquid phase is
614 not as significant as that of the hexane. This prevents severe
615 decrease of liquid-capping agent contact and the results (Table
616 \ref{AuThiolToluene}) show only a slightly decreased interface
617 conductance. Therefore, solvent-capping agent contact should play an
618 important role in the thermal transport process across the interface
619 in that higher degree of contact could yield increased conductance.
620
621 \begin{table*}
622 \begin{minipage}{\linewidth}
623 \begin{center}
624 \caption{Computed interfacial thermal conductivity ($G$ and
625 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
626 interface at different temperatures using a range of energy
627 fluxes. Error estimates indicated in parenthesis.}
628
629 \begin{tabular}{ccccc}
630 \hline\hline
631 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
632 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
633 \hline
634 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
635 & & -1.86 & 180(3) & 135(21) \\
636 & & -3.93 & 176(5) & 113(12) \\
637 \hline
638 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
639 & & -4.19 & 135(9) & 113(12) \\
640 \hline\hline
641 \end{tabular}
642 \label{AuThiolToluene}
643 \end{center}
644 \end{minipage}
645 \end{table*}
646
647 Besides lower interfacial thermal conductance, surfaces in relatively
648 high temperatures are susceptible to reconstructions, when
649 butanethiols have a full coverage on the Au(111) surface. These
650 reconstructions include surface Au atoms migrated outward to the S
651 atom layer, and butanethiol molecules embedded into the original
652 surface Au layer. The driving force for this behavior is the strong
653 Au-S interactions in our simulations. And these reconstructions lead
654 to higher ratio of Au-S attraction and thus is energetically
655 favorable. Furthermore, this phenomenon agrees with experimental
656 results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
657 {\it et al.} had kept their Au(111) slab rigid so that their
658 simulations can reach 300K without surface reconstructions. Without
659 this practice, simulating 100\% thiol covered interfaces under higher
660 temperatures could hardly avoid surface reconstructions. However, our
661 measurement is based on assuming homogeneity on $x$ and $y$ dimensions
662 so that measurement of $T$ at particular $z$ would be an effective
663 average of the particles of the same type. Since surface
664 reconstructions could eliminate the original $x$ and $y$ dimensional
665 homogeneity, measurement of $G$ is more difficult to conduct under
666 higher temperatures. Therefore, most of our measurements are
667 undertaken at $\langle T\rangle\sim$200K.
668
669 However, when the surface is not completely covered by butanethiols,
670 the simulated system is more resistent to the reconstruction
671 above. Our Au-butanethiol/toluene system had the Au(111) surfaces 90\%
672 covered by butanethiols, but did not see this above phenomena even at
673 $\langle T\rangle\sim$300K. The empty three-fold sites not occupied by
674 capping agents could help prevent surface reconstruction in that they
675 provide other means of capping agent relaxation. It is observed that
676 butanethiols can migrate to their neighbor empty sites during a
677 simulation. Therefore, we were able to obtain $G$'s for these
678 interfaces even at a relatively high temperature without being
679 affected by surface reconstructions.
680
681 \subsection{Influence of Capping Agent Coverage on $G$}
682 To investigate the influence of butanethiol coverage on interfacial
683 thermal conductance, a series of different coverage Au-butanethiol
684 surfaces is prepared and solvated with various organic
685 molecules. These systems are then equilibrated and their interfacial
686 thermal conductivity are measured with our NIVS algorithm. Figure
687 \ref{coverage} demonstrates the trend of conductance change with
688 respect to different coverages of butanethiol. To study the isotope
689 effect in interfacial thermal conductance, deuterated UA-hexane is
690 included as well.
691
692 \begin{figure}
693 \includegraphics[width=\linewidth]{coverage}
694 \caption{Comparison of interfacial thermal conductivity ($G$) values
695 for the Au-butanethiol/solvent interface with various UA models and
696 different capping agent coverages at $\langle T\rangle\sim$200K
697 using certain energy flux respectively.}
698 \label{coverage}
699 \end{figure}
700
701 It turned out that with partial covered butanethiol on the Au(111)
702 surface, the derivative definition for $G^\prime$
703 (Eq. \ref{derivativeG}) was difficult to apply, due to the difficulty
704 in locating the maximum of change of $\lambda$. Instead, the discrete
705 definition (Eq. \ref{discreteG}) is easier to apply, as the Gibbs
706 deviding surface can still be well-defined. Therefore, $G$ (not
707 $G^\prime$) was used for this section.
708
709 From Figure \ref{coverage}, one can see the significance of the
710 presence of capping agents. Even when a fraction of the Au(111)
711 surface sites are covered with butanethiols, the conductivity would
712 see an enhancement by at least a factor of 3. This indicates the
713 important role cappping agent is playing for thermal transport
714 phenomena on metal / organic solvent surfaces.
715
716 Interestingly, as one could observe from our results, the maximum
717 conductance enhancement (largest $G$) happens while the surfaces are
718 about 75\% covered with butanethiols. This again indicates that
719 solvent-capping agent contact has an important role of the thermal
720 transport process. Slightly lower butanethiol coverage allows small
721 gaps between butanethiols to form. And these gaps could be filled with
722 solvent molecules, which acts like ``heat conductors'' on the
723 surface. The higher degree of interaction between these solvent
724 molecules and capping agents increases the enhancement effect and thus
725 produces a higher $G$ than densely packed butanethiol arrays. However,
726 once this maximum conductance enhancement is reached, $G$ decreases
727 when butanethiol coverage continues to decrease. Each capping agent
728 molecule reaches its maximum capacity for thermal
729 conductance. Therefore, even higher solvent-capping agent contact
730 would not offset this effect. Eventually, when butanethiol coverage
731 continues to decrease, solvent-capping agent contact actually
732 decreases with the disappearing of butanethiol molecules. In this
733 case, $G$ decrease could not be offset but instead accelerated. [NEED
734 SNAPSHOT SHOWING THE PHENOMENA / SLAB DENSITY ANALYSIS]
735
736 A comparison of the results obtained from differenet organic solvents
737 can also provide useful information of the interfacial thermal
738 transport process. The deuterated hexane (UA) results do not appear to
739 be much different from those of normal hexane (UA), given that
740 butanethiol (UA) is non-deuterated for both solvents. These UA model
741 studies, even though eliminating C-H vibration samplings, still have
742 C-C vibrational frequencies different from each other. However, these
743 differences in the infrared range do not seem to produce an observable
744 difference for the results of $G$ (Figure \ref{uahxnua}).
745
746 \begin{figure}
747 \includegraphics[width=\linewidth]{uahxnua}
748 \caption{Vibrational spectra obtained for normal (upper) and
749 deuterated (lower) hexane in Au-butanethiol/hexane
750 systems. Butanethiol spectra are shown as reference. Both hexane and
751 butanethiol were using United-Atom models.}
752 \label{uahxnua}
753 \end{figure}
754
755 Furthermore, results for rigid body toluene solvent, as well as other
756 UA-hexane solvents, are reasonable within the general experimental
757 ranges\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406}. This
758 suggests that explicit hydrogen might not be a required factor for
759 modeling thermal transport phenomena of systems such as
760 Au-thiol/organic solvent.
761
762 However, results for Au-butanethiol/toluene do not show an identical
763 trend with those for Au-butanethiol/hexane in that $G$ remains at
764 approximately the same magnitue when butanethiol coverage differs from
765 25\% to 75\%. This might be rooted in the molecule shape difference
766 for planar toluene and chain-like {\it n}-hexane. Due to this
767 difference, toluene molecules have more difficulty in occupying
768 relatively small gaps among capping agents when their coverage is not
769 too low. Therefore, the solvent-capping agent contact may keep
770 increasing until the capping agent coverage reaches a relatively low
771 level. This becomes an offset for decreasing butanethiol molecules on
772 its effect to the process of interfacial thermal transport. Thus, one
773 can see a plateau of $G$ vs. butanethiol coverage in our results.
774
775 \subsection{Influence of Chosen Molecule Model on $G$}
776 In addition to UA solvent/capping agent models, AA models are included
777 in our simulations as well. Besides simulations of the same (UA or AA)
778 model for solvent and capping agent, different models can be applied
779 to different components. Furthermore, regardless of models chosen,
780 either the solvent or the capping agent can be deuterated, similar to
781 the previous section. Table \ref{modelTest} summarizes the results of
782 these studies.
783
784 \begin{table*}
785 \begin{minipage}{\linewidth}
786 \begin{center}
787
788 \caption{Computed interfacial thermal conductivity ($G$ and
789 $G^\prime$) values for interfaces using various models for
790 solvent and capping agent (or without capping agent) at
791 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
792 or capping agent molecules; ``Avg.'' denotes results that are
793 averages of simulations under different $J_z$'s. Error
794 estimates indicated in parenthesis.)}
795
796 \begin{tabular}{llccc}
797 \hline\hline
798 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
799 (or bare surface) & model & (GW/m$^2$) &
800 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
801 \hline
802 UA & UA hexane & Avg. & 131(9) & 87(10) \\
803 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
804 & AA hexane & Avg. & 131(6) & 122(10) \\
805 & UA toluene & 1.96 & 187(16) & 151(11) \\
806 & AA toluene & 1.89 & 200(36) & 149(53) \\
807 \hline
808 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
809 & AA hexane & Avg. & 442(14) & 356(31) \\
810 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
811 & UA toluene & 1.98 & 125(25) & 97(60) \\
812 & AA toluene & 3.79 & 487(56) & 290(42) \\
813 \hline
814 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
815 & AA hexane & 1.92 & 243(29) & 191(11) \\
816 & AA toluene & 1.93 & 364(36) & 322(67) \\
817 \hline
818 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
819 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
820 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
821 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
822 \hline\hline
823 \end{tabular}
824 \label{modelTest}
825 \end{center}
826 \end{minipage}
827 \end{table*}
828
829 To facilitate direct comparison, the same system with differnt models
830 for different components uses the same length scale for their
831 simulation cells. Without the presence of capping agent, using
832 different models for hexane yields similar results for both $G$ and
833 $G^\prime$, and these two definitions agree with eath other very
834 well. This indicates very weak interaction between the metal and the
835 solvent, and is a typical case for acoustic impedance mismatch between
836 these two phases.
837
838 As for Au(111) surfaces completely covered by butanethiols, the choice
839 of models for capping agent and solvent could impact the measurement
840 of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
841 interfaces, using AA model for both butanethiol and hexane yields
842 substantially higher conductivity values than using UA model for at
843 least one component of the solvent and capping agent, which exceeds
844 the general range of experimental measurement results. This is
845 probably due to the classically treated C-H vibrations in the AA
846 model, which should not be appreciably populated at normal
847 temperatures. In comparison, once either the hexanes or the
848 butanethiols are deuterated, one can see a significantly lower $G$ and
849 $G^\prime$. In either of these cases, the C-H(D) vibrational overlap
850 between the solvent and the capping agent is removed (Figure
851 \ref{aahxntln}). Conclusively, the improperly treated C-H vibration in
852 the AA model produced over-predicted results accordingly. Compared to
853 the AA model, the UA model yields more reasonable results with higher
854 computational efficiency.
855
856 \begin{figure}
857 \includegraphics[width=\linewidth]{aahxntln}
858 \caption{Spectra obtained for All-Atom model Au-butanethil/solvent
859 systems. When butanethiol is deuterated (lower left), its
860 vibrational overlap with hexane would decrease significantly,
861 compared with normal butanethiol (upper left). However, this
862 dramatic change does not apply to toluene as much (right).}
863 \label{aahxntln}
864 \end{figure}
865
866 However, for Au-butanethiol/toluene interfaces, having the AA
867 butanethiol deuterated did not yield a significant change in the
868 measurement results. Compared to the C-H vibrational overlap between
869 hexane and butanethiol, both of which have alkyl chains, that overlap
870 between toluene and butanethiol is not so significant and thus does
871 not have as much contribution to the heat exchange
872 process. Conversely, extra degrees of freedom such as the C-H
873 vibrations could yield higher heat exchange rate between these two
874 phases and result in a much higher conductivity.
875
876 Although the QSC model for Au is known to predict an overly low value
877 for bulk metal gold conductivity\cite{kuang:164101}, our computational
878 results for $G$ and $G^\prime$ do not seem to be affected by this
879 drawback of the model for metal. Instead, our results suggest that the
880 modeling of interfacial thermal transport behavior relies mainly on
881 the accuracy of the interaction descriptions between components
882 occupying the interfaces.
883
884 \subsection{Role of Capping Agent in Interfacial Thermal Conductance}
885 The vibrational spectra for gold slabs in different environments are
886 shown as in Figure \ref{specAu}. Regardless of the presence of
887 solvent, the gold surfaces covered by butanethiol molecules, compared
888 to bare gold surfaces, exhibit an additional peak observed at the
889 frequency of $\sim$170cm$^{-1}$, which is attributed to the S-Au
890 bonding vibration. This vibration enables efficient thermal transport
891 from surface Au layer to the capping agents. Therefore, in our
892 simulations, the Au/S interfaces do not appear major heat barriers
893 compared to the butanethiol / solvent interfaces.
894
895 Simultaneously, the vibrational overlap between butanethiol and
896 organic solvents suggests higher thermal exchange efficiency between
897 these two components. Even exessively high heat transport was observed
898 when All-Atom models were used and C-H vibrations were treated
899 classically. Compared to metal and organic liquid phase, the heat
900 transfer efficiency between butanethiol and organic solvents is closer
901 to that within bulk liquid phase.
902
903 Furthermore, our observation validated previous
904 results\cite{hase:2010} that the intramolecular heat transport of
905 alkylthiols is highly effecient. As a combinational effects of these
906 phenomena, butanethiol acts as a channel to expedite thermal transport
907 process. The acoustic impedance mismatch between the metal and the
908 liquid phase can be effectively reduced with the presence of suitable
909 capping agents.
910
911 \begin{figure}
912 \includegraphics[width=\linewidth]{vibration}
913 \caption{Vibrational spectra obtained for gold in different
914 environments.}
915 \label{specAu}
916 \end{figure}
917
918 [MAY ADD COMPARISON OF AU SLAB WIDTHS]
919
920 \section{Conclusions}
921 The NIVS algorithm we developed has been applied to simulations of
922 Au-butanethiol surfaces with organic solvents. This algorithm allows
923 effective unphysical thermal flux transferred between the metal and
924 the liquid phase. With the flux applied, we were able to measure the
925 corresponding thermal gradient and to obtain interfacial thermal
926 conductivities. Under steady states, single trajectory simulation
927 would be enough for accurate measurement. This would be advantageous
928 compared to transient state simulations, which need multiple
929 trajectories to produce reliable average results.
930
931 Our simulations have seen significant conductance enhancement with the
932 presence of capping agent, compared to the bare gold / liquid
933 interfaces. The acoustic impedance mismatch between the metal and the
934 liquid phase is effectively eliminated by proper capping
935 agent. Furthermore, the coverage precentage of the capping agent plays
936 an important role in the interfacial thermal transport
937 process. Moderately lower coverages allow higher contact between
938 capping agent and solvent, and thus could further enhance the heat
939 transfer process.
940
941 Our measurement results, particularly of the UA models, agree with
942 available experimental data. This indicates that our force field
943 parameters have a nice description of the interactions between the
944 particles at the interfaces. AA models tend to overestimate the
945 interfacial thermal conductance in that the classically treated C-H
946 vibration would be overly sampled. Compared to the AA models, the UA
947 models have higher computational efficiency with satisfactory
948 accuracy, and thus are preferable in interfacial thermal transport
949 modelings. Of the two definitions for $G$, the discrete form
950 (Eq. \ref{discreteG}) was easier to use and gives out relatively
951 consistent results, while the derivative form (Eq. \ref{derivativeG})
952 is not as versatile. Although $G^\prime$ gives out comparable results
953 and follows similar trend with $G$ when measuring close to fully
954 covered or bare surfaces, the spatial resolution of $T$ profile is
955 limited for accurate computation of derivatives data.
956
957 Vlugt {\it et al.} has investigated the surface thiol structures for
958 nanocrystal gold and pointed out that they differs from those of the
959 Au(111) surface\cite{vlugt:cpc2007154}. This difference might lead to
960 change of interfacial thermal transport behavior as well. To
961 investigate this problem, an effective means to introduce thermal flux
962 and measure the corresponding thermal gradient is desirable for
963 simulating structures with spherical symmetry.
964
965 \section{Acknowledgments}
966 Support for this project was provided by the National Science
967 Foundation under grant CHE-0848243. Computational time was provided by
968 the Center for Research Computing (CRC) at the University of Notre
969 Dame. \newpage
970
971 \bibliography{interfacial}
972
973 \end{doublespace}
974 \end{document}
975