ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3754
Committed: Thu Jul 28 20:25:08 2011 UTC (12 years, 11 months ago) by gezelter
Content type: application/x-tex
File size: 58521 byte(s)
Log Message:
yesterday's 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 common way is to assume the temperature is
194 discrete on the two sides of the interface. $G$ can be calculated
195 using the applied thermal flux $J$ and the maximum temperature
196 difference measured along the thermal gradient max($\Delta T$), which
197 occurs at the Gibbs deviding surface (Figure \ref{demoPic}):
198 \begin{equation}
199 G=\frac{J}{\Delta T}
200 \label{discreteG}
201 \end{equation}
202
203 \begin{figure}
204 \includegraphics[width=\linewidth]{method}
205 \caption{Interfacial conductance can be calculated by applying an
206 (unphysical) kinetic energy flux between two slabs, one located
207 within the metal and another on the edge of the periodic box. The
208 system responds by forming a thermal response or a gradient. In
209 bulk liquids, this gradient typically has a single slope, but in
210 interfacial systems, there are distinct thermal conductivity
211 domains. The interfacial conductance, $G$ is found by measuring the
212 temperature gap at the Gibbs dividing surface, or by using second
213 derivatives of the thermal profile.}
214 \label{demoPic}
215 \end{figure}
216
217 The other approach is to assume a continuous temperature profile along
218 the thermal gradient axis (e.g. $z$) and define $G$ at the point where
219 the magnitude of thermal conductivity ($\lambda$) change reaches its
220 maximum, given that $\lambda$ is well-defined throughout the space:
221 \begin{equation}
222 G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
223 = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
224 \left(\frac{\partial T}{\partial z}\right)\right)\Big|
225 = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
226 \Big/\left(\frac{\partial T}{\partial z}\right)^2
227 \label{derivativeG}
228 \end{equation}
229
230 With temperature profiles obtained from simulation, one is able to
231 approximate the first and second derivatives of $T$ with finite
232 difference methods and calculate $G^\prime$. In what follows, both
233 definitions have been used, and are compared in the results.
234
235 To investigate the interfacial conductivity at metal / solvent
236 interfaces, we have modeled a metal slab with its (111) surfaces
237 perpendicular to the $z$-axis of our simulation cells. The metal slab
238 has been prepared both with and without capping agents on the exposed
239 surface, and has been solvated with simple organic solvents, as
240 illustrated in Figure \ref{gradT}.
241
242 With the simulation cell described above, we are able to equilibrate
243 the system and impose an unphysical thermal flux between the liquid
244 and the metal phase using the NIVS algorithm. By periodically applying
245 the unphysical flux, we obtained a temperature profile and its spatial
246 derivatives. Figure \ref{gradT} shows how an applied thermal flux can
247 be used to obtain the 1st and 2nd derivatives of the temperature
248 profile.
249
250 \begin{figure}
251 \includegraphics[width=\linewidth]{gradT}
252 \caption{A sample of Au-butanethiol/hexane interfacial system and the
253 temperature profile after a kinetic energy flux is imposed to
254 it. The 1st and 2nd derivatives of the temperature profile can be
255 obtained with finite difference approximation (lower panel).}
256 \label{gradT}
257 \end{figure}
258
259 \section{Computational Details}
260 \subsection{Simulation Protocol}
261 The NIVS algorithm has been implemented in our MD simulation code,
262 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
263 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
264 under atmospheric pressure (1 atm) and 200K. After equilibration,
265 butanethiol capping agents were placed at three-fold hollow sites on
266 the Au(111) surfaces. These sites are either {\it fcc} or {\it
267 hcp} sites, although Hase {\it et al.} found that they are
268 equivalent in a heat transfer process,\cite{hase:2010} so we did not
269 distinguish between these sites in our study. The maximum butanethiol
270 capacity on Au surface is $1/3$ of the total number of surface Au
271 atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
272 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
273 series of lower coverages was also prepared by eliminating
274 butanethiols from the higher coverage surface in a regular manner. The
275 lower coverages were prepared in order to study the relation between
276 coverage and interfacial conductance.
277
278 The capping agent molecules were allowed to migrate during the
279 simulations. They distributed themselves uniformly and sampled a
280 number of three-fold sites throughout out study. Therefore, the
281 initial configuration does not noticeably affect the sampling of a
282 variety of configurations of the same coverage, and the final
283 conductance measurement would be an average effect of these
284 configurations explored in the simulations.
285
286 After the modified Au-butanethiol surface systems were equilibrated in
287 the canonical (NVT) ensemble, organic solvent molecules were packed in
288 the previously empty part of the simulation cells.\cite{packmol} Two
289 solvents were investigated, one which has little vibrational overlap
290 with the alkanethiol and which has a planar shape (toluene), and one
291 which has similar vibrational frequencies to the capping agent and
292 chain-like shape ({\it n}-hexane).
293
294 The simulation cells were not particularly extensive along the
295 $z$-axis, as a very long length scale for the thermal gradient may
296 cause excessively hot or cold temperatures in the middle of the
297 solvent region and lead to undesired phenomena such as solvent boiling
298 or freezing when a thermal flux is applied. Conversely, too few
299 solvent molecules would change the normal behavior of the liquid
300 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
301 these extreme cases did not happen to our simulations. The spacing
302 between periodic images of the gold interfaces is $45 \sim 75$\AA.
303
304 The initial configurations generated are further equilibrated with the
305 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
306 change. This is to ensure that the equilibration of liquid phase does
307 not affect the metal's crystalline structure. Comparisons were made
308 with simulations that allowed changes of $L_x$ and $L_y$ during NPT
309 equilibration. No substantial changes in the box geometry were noticed
310 in these simulations. After ensuring the liquid phase reaches
311 equilibrium at atmospheric pressure (1 atm), further equilibration was
312 carried out under canonical (NVT) and microcanonical (NVE) ensembles.
313
314 After the systems reach equilibrium, NIVS was used to impose an
315 unphysical thermal flux between the metal and the liquid phases. Most
316 of our simulations were done under an average temperature of
317 $\sim$200K. Therefore, thermal flux usually came from the metal to the
318 liquid so that the liquid has a higher temperature and would not
319 freeze due to lowered temperatures. After this induced temperature
320 gradient had stablized, the temperature profile of the simulation cell
321 was recorded. To do this, the simulation cell is devided evenly into
322 $N$ slabs along the $z$-axis. The average temperatures of each slab
323 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
324 the same, the derivatives of $T$ with respect to slab number $n$ can
325 be directly used for $G^\prime$ calculations: \begin{equation}
326 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
327 \Big/\left(\frac{\partial T}{\partial z}\right)^2
328 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
329 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
330 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
331 \Big/\left(\frac{\partial T}{\partial n}\right)^2
332 \label{derivativeG2}
333 \end{equation}
334
335 All of the above simulation procedures use a time step of 1 fs. Each
336 equilibration stage took a minimum of 100 ps, although in some cases,
337 longer equilibration stages were utilized.
338
339 \subsection{Force Field Parameters}
340 Our simulations include a number of chemically distinct components.
341 Figure \ref{demoMol} demonstrates the sites defined for both
342 United-Atom and All-Atom models of the organic solvent and capping
343 agents in our simulations. Force field parameters are needed for
344 interactions both between the same type of particles and between
345 particles of different species.
346
347 \begin{figure}
348 \includegraphics[width=\linewidth]{structures}
349 \caption{Structures of the capping agent and solvents utilized in
350 these simulations. The chemically-distinct sites (a-e) are expanded
351 in terms of constituent atoms for both United Atom (UA) and All Atom
352 (AA) force fields. Most parameters are from
353 Refs. \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols} (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au atoms are given in Table \ref{MnM}.}
354 \label{demoMol}
355 \end{figure}
356
357 The Au-Au interactions in metal lattice slab is described by the
358 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
359 potentials include zero-point quantum corrections and are
360 reparametrized for accurate surface energies compared to the
361 Sutton-Chen potentials.\cite{Chen90}
362
363 For the two solvent molecules, {\it n}-hexane and toluene, two
364 different atomistic models were utilized. Both solvents were modeled
365 using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
366 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
367 for our UA solvent molecules. In these models, sites are located at
368 the carbon centers for alkyl groups. Bonding interactions, including
369 bond stretches and bends and torsions, were used for intra-molecular
370 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
371 potentials are used.
372
373 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
374 simple and computationally efficient, while maintaining good accuracy.
375 However, the TraPPE-UA model for alkanes is known to predict a slighly
376 lower boiling point than experimental values. This is one of the
377 reasons we used a lower average temperature (200K) for our
378 simulations. If heat is transferred to the liquid phase during the
379 NIVS simulation, the liquid in the hot slab can actually be
380 substantially warmer than the mean temperature in the simulation. The
381 lower mean temperatures therefore prevent solvent boiling.
382
383 For UA-toluene, the non-bonded potentials between intermolecular sites
384 have a similar Lennard-Jones formulation. The toluene molecules were
385 treated as a single rigid body, so there was no need for
386 intramolecular interactions (including bonds, bends, or torsions) in
387 this solvent model.
388
389 Besides the TraPPE-UA models, AA models for both organic solvents are
390 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
391 were used. For hexane, additional explicit hydrogen sites were
392 included. Besides bonding and non-bonded site-site interactions,
393 partial charges and the electrostatic interactions were added to each
394 CT and HC site. For toluene, a flexible model for the toluene molecule
395 was utilized which included bond, bend, torsion, and inversion
396 potentials to enforce ring planarity.
397
398 The butanethiol capping agent in our simulations, were also modeled
399 with both UA and AA model. The TraPPE-UA force field includes
400 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
401 UA butanethiol model in our simulations. The OPLS-AA also provides
402 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
403 surfaces do not have the hydrogen atom bonded to sulfur. To derive
404 suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
405 adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
406 modify the parameters for the CTS atom to maintain charge neutrality
407 in the molecule. Note that the model choice (UA or AA) for the capping
408 agent can be different from the solvent. Regardless of model choice,
409 the force field parameters for interactions between capping agent and
410 solvent can be derived using Lorentz-Berthelot Mixing Rule:
411 \begin{eqnarray}
412 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
413 \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
414 \end{eqnarray}
415
416 To describe the interactions between metal (Au) and non-metal atoms,
417 we refer to an adsorption study of alkyl thiols on gold surfaces by
418 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
419 Lennard-Jones form of potential parameters for the interaction between
420 Au and pseudo-atoms CH$_x$ and S based on a well-established and
421 widely-used effective potential of Hautman and Klein for the Au(111)
422 surface.\cite{hautman:4994} As our simulations require the gold slab
423 to be flexible to accommodate thermal excitation, the pair-wise form
424 of potentials they developed was used for our study.
425
426 The potentials developed from {\it ab initio} calculations by Leng
427 {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
428 interactions between Au and aromatic C/H atoms in toluene. However,
429 the Lennard-Jones parameters between Au and other types of particles,
430 (e.g. AA alkanes) have not yet been established. For these
431 interactions, the Lorentz-Berthelot mixing rule can be used to derive
432 effective single-atom LJ parameters for the metal using the fit values
433 for toluene. These are then used to construct reasonable mixing
434 parameters for the interactions between the gold and other atoms.
435 Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
436 our simulations.
437
438 \begin{table*}
439 \begin{minipage}{\linewidth}
440 \begin{center}
441 \caption{Non-bonded interaction parameters (including cross
442 interactions with Au atoms) for both force fields used in this
443 work.}
444 \begin{tabular}{lllllll}
445 \hline\hline
446 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
447 $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
448 & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
449 \hline
450 United Atom (UA)
451 &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
452 &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
453 &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
454 &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
455 \hline
456 All Atom (AA)
457 &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
458 &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
459 &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
460 &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
461 &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
462 &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
463 \hline
464 Both UA and AA
465 & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
466 \hline\hline
467 \end{tabular}
468 \label{MnM}
469 \end{center}
470 \end{minipage}
471 \end{table*}
472
473
474 \section{Results}
475 There are many factors contributing to the measured interfacial
476 conductance; some of these factors are physically motivated
477 (e.g. coverage of the surface by the capping agent coverage and
478 solvent identity), while some are governed by parameters of the
479 methodology (e.g. applied flux and the formulas used to obtain the
480 conductance). In this section we discuss the major physical and
481 calculational effects on the computed conductivity.
482
483 \subsection{Effects due to capping agent coverage}
484
485 A series of different initial conditions with a range of surface
486 coverages was prepared and solvated with various with both of the
487 solvent molecules. These systems were then equilibrated and their
488 interfacial thermal conductivity was measured with our NIVS
489 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
490 with respect to surface coverage.
491
492 \begin{figure}
493 \includegraphics[width=\linewidth]{coverage}
494 \caption{Comparison of interfacial thermal conductivity ($G$) values
495 for the Au-butanethiol/solvent interface with various UA models and
496 different capping agent coverages at $\langle T\rangle\sim$200K.}
497 \label{coverage}
498 \end{figure}
499
500
501 In partially covered butanethiol on the Au(111) surface, the
502 derivative definition for $G^\prime$ (Eq. \ref{derivativeG}) becomes
503 difficult to apply, as the location of maximum change of $\lambda$
504 becomes washed out. The discrete definition (Eq. \ref{discreteG}) is
505 easier to apply, as the Gibbs dividing surface is still
506 well-defined. Therefore, $G$ (not $G^\prime$) was used in this
507 section.
508
509 From Figure \ref{coverage}, one can see the significance of the
510 presence of capping agents. When even a small fraction of the Au(111)
511 surface sites are covered with butanethiols, the conductivity exhibits
512 an enhancement by at least a factor of 3. This indicates the important
513 role cappping agents are playing for thermal transport at metal /
514 organic solvent surfaces.
515
516 We note a non-monotonic behavior in the interfacial conductance as a
517 function of surface coverage. The maximum conductance (largest $G$)
518 happens when the surfaces are about 75\% covered with butanethiol
519 caps. The reason for this behavior is not entirely clear. One
520 explanation is that incomplete butanethiol coverage allows small gaps
521 between butanethiols to form. These gaps can be filled by transient
522 solvent molecules. These solvent molecules couple very strongly with
523 the hot capping agent molecules near the surface, and can then carry
524 (diffusively) the excess thermal energy away from the surface.
525
526 There appears to be a competition between the conduction of the
527 thermal energy away from the surface by the capping agents (enhanced
528 by greater coverage) and the coupling of the capping agents with the
529 solvent (enhanced by physical contact at lower coverages). This
530 competition would lead to the non-monotonic coverage behavior observed
531 here.
532
533 A comparison of the results obtained from the two different organic
534 solvents can also provide useful information of the interfacial
535 thermal transport process. The deuterated hexane (UA) results do not
536 appear to be substantially different from those of normal hexane (UA),
537 given that butanethiol (UA) is non-deuterated for both solvents. The
538 UA models, even though they have eliminated C-H vibrational overlap,
539 still have significant overlap in the infrared spectra. Because
540 differences in the infrared range do not seem to produce an observable
541 difference for the results of $G$ (Figure \ref{uahxnua}).
542
543 \begin{figure}
544 \includegraphics[width=\linewidth]{uahxnua}
545 \caption{Vibrational spectra obtained for normal (upper) and
546 deuterated (lower) hexane in Au-butanethiol/hexane
547 systems. Butanethiol spectra are shown as reference. Both hexane and
548 butanethiol were using United-Atom models.}
549 \label{uahxnua}
550 \end{figure}
551
552 Furthermore, results for rigid body toluene solvent, as well as other
553 UA-hexane solvents, are reasonable within the general experimental
554 ranges\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406}. This
555 suggests that explicit hydrogen might not be a required factor for
556 modeling thermal transport phenomena of systems such as
557 Au-thiol/organic solvent.
558
559 However, results for Au-butanethiol/toluene do not show an identical
560 trend with those for Au-butanethiol/hexane in that $G$ remains at
561 approximately the same magnitue when butanethiol coverage differs from
562 25\% to 75\%. This might be rooted in the molecule shape difference
563 for planar toluene and chain-like {\it n}-hexane. Due to this
564 difference, toluene molecules have more difficulty in occupying
565 relatively small gaps among capping agents when their coverage is not
566 too low. Therefore, the solvent-capping agent contact may keep
567 increasing until the capping agent coverage reaches a relatively low
568 level. This becomes an offset for decreasing butanethiol molecules on
569 its effect to the process of interfacial thermal transport. Thus, one
570 can see a plateau of $G$ vs. butanethiol coverage in our results.
571
572 \subsection{Effects due to Solvent \& Solvent Models}
573 In addition to UA solvent/capping agent models, AA models are included
574 in our simulations as well. Besides simulations of the same (UA or AA)
575 model for solvent and capping agent, different models can be applied
576 to different components. Furthermore, regardless of models chosen,
577 either the solvent or the capping agent can be deuterated, similar to
578 the previous section. Table \ref{modelTest} summarizes the results of
579 these studies.
580
581 \begin{table*}
582 \begin{minipage}{\linewidth}
583 \begin{center}
584
585 \caption{Computed interfacial thermal conductivity ($G$ and
586 $G^\prime$) values for interfaces using various models for
587 solvent and capping agent (or without capping agent) at
588 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
589 or capping agent molecules; ``Avg.'' denotes results that are
590 averages of simulations under different $J_z$'s. Error
591 estimates indicated in parenthesis.)}
592
593 \begin{tabular}{llccc}
594 \hline\hline
595 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
596 (or bare surface) & model & (GW/m$^2$) &
597 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
598 \hline
599 UA & UA hexane & Avg. & 131(9) & 87(10) \\
600 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
601 & AA hexane & Avg. & 131(6) & 122(10) \\
602 & UA toluene & 1.96 & 187(16) & 151(11) \\
603 & AA toluene & 1.89 & 200(36) & 149(53) \\
604 \hline
605 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
606 & AA hexane & Avg. & 442(14) & 356(31) \\
607 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
608 & UA toluene & 1.98 & 125(25) & 97(60) \\
609 & AA toluene & 3.79 & 487(56) & 290(42) \\
610 \hline
611 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
612 & AA hexane & 1.92 & 243(29) & 191(11) \\
613 & AA toluene & 1.93 & 364(36) & 322(67) \\
614 \hline
615 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
616 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
617 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
618 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
619 \hline\hline
620 \end{tabular}
621 \label{modelTest}
622 \end{center}
623 \end{minipage}
624 \end{table*}
625
626 To facilitate direct comparison, the same system with differnt models
627 for different components uses the same length scale for their
628 simulation cells. Without the presence of capping agent, using
629 different models for hexane yields similar results for both $G$ and
630 $G^\prime$, and these two definitions agree with eath other very
631 well. This indicates very weak interaction between the metal and the
632 solvent, and is a typical case for acoustic impedance mismatch between
633 these two phases.
634
635 As for Au(111) surfaces completely covered by butanethiols, the choice
636 of models for capping agent and solvent could impact the measurement
637 of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
638 interfaces, using AA model for both butanethiol and hexane yields
639 substantially higher conductivity values than using UA model for at
640 least one component of the solvent and capping agent, which exceeds
641 the general range of experimental measurement results. This is
642 probably due to the classically treated C-H vibrations in the AA
643 model, which should not be appreciably populated at normal
644 temperatures. In comparison, once either the hexanes or the
645 butanethiols are deuterated, one can see a significantly lower $G$ and
646 $G^\prime$. In either of these cases, the C-H(D) vibrational overlap
647 between the solvent and the capping agent is removed (Figure
648 \ref{aahxntln}). Conclusively, the improperly treated C-H vibration in
649 the AA model produced over-predicted results accordingly. Compared to
650 the AA model, the UA model yields more reasonable results with higher
651 computational efficiency.
652
653 \begin{figure}
654 \includegraphics[width=\linewidth]{aahxntln}
655 \caption{Spectra obtained for All-Atom model Au-butanethil/solvent
656 systems. When butanethiol is deuterated (lower left), its
657 vibrational overlap with hexane would decrease significantly,
658 compared with normal butanethiol (upper left). However, this
659 dramatic change does not apply to toluene as much (right).}
660 \label{aahxntln}
661 \end{figure}
662
663 However, for Au-butanethiol/toluene interfaces, having the AA
664 butanethiol deuterated did not yield a significant change in the
665 measurement results. Compared to the C-H vibrational overlap between
666 hexane and butanethiol, both of which have alkyl chains, that overlap
667 between toluene and butanethiol is not so significant and thus does
668 not have as much contribution to the heat exchange
669 process. Conversely, extra degrees of freedom such as the C-H
670 vibrations could yield higher heat exchange rate between these two
671 phases and result in a much higher conductivity.
672
673 Although the QSC model for Au is known to predict an overly low value
674 for bulk metal gold conductivity\cite{kuang:164101}, our computational
675 results for $G$ and $G^\prime$ do not seem to be affected by this
676 drawback of the model for metal. Instead, our results suggest that the
677 modeling of interfacial thermal transport behavior relies mainly on
678 the accuracy of the interaction descriptions between components
679 occupying the interfaces.
680
681 \subsection{Effects due to methodology and simulation parameters}
682
683 We have varied our protocol or other parameters of the simulations in
684 order to investigate how these factors would affect the measurement of
685 $G$'s. It turned out that while some of these parameters would not
686 affect the results substantially, some other changes to the
687 simulations would have a significant impact on the measurement
688 results.
689
690 In some of our simulations, we allowed $L_x$ and $L_y$ to change
691 during equilibrating the liquid phase. Due to the stiffness of the
692 crystalline Au structure, $L_x$ and $L_y$ would not change noticeably
693 after equilibration. Although $L_z$ could fluctuate $\sim$1\% after a
694 system is fully equilibrated in the NPT ensemble, this fluctuation, as
695 well as those of $L_x$ and $L_y$ (which is significantly smaller),
696 would not be magnified on the calculated $G$'s, as shown in Table
697 \ref{AuThiolHexaneUA}. This insensivity to $L_i$ fluctuations allows
698 reliable measurement of $G$'s without the necessity of extremely
699 cautious equilibration process.
700
701 As stated in our computational details, the spacing filled with
702 solvent molecules can be chosen within a range. This allows some
703 change of solvent molecule numbers for the same Au-butanethiol
704 surfaces. We did this study on our Au-butanethiol/hexane
705 simulations. Nevertheless, the results obtained from systems of
706 different $N_{hexane}$ did not indicate that the measurement of $G$ is
707 susceptible to this parameter. For computational efficiency concern,
708 smaller system size would be preferable, given that the liquid phase
709 structure is not affected.
710
711 \subsubsection{Effects of applied flux}
712 Our NIVS algorithm allows change of unphysical thermal flux both in
713 direction and in quantity. This feature extends our investigation of
714 interfacial thermal conductance. However, the magnitude of this
715 thermal flux is not arbitary if one aims to obtain a stable and
716 reliable thermal gradient. A temperature profile would be
717 substantially affected by noise when $|J_z|$ has a much too low
718 magnitude; while an excessively large $|J_z|$ that overwhelms the
719 conductance capacity of the interface would prevent a thermal gradient
720 to reach a stablized steady state. NIVS has the advantage of allowing
721 $J$ to vary in a wide range such that the optimal flux range for $G$
722 measurement can generally be simulated by the algorithm. Within the
723 optimal range, we were able to study how $G$ would change according to
724 the thermal flux across the interface. For our simulations, we denote
725 $J_z$ to be positive when the physical thermal flux is from the liquid
726 to metal, and negative vice versa. The $G$'s measured under different
727 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and
728 \ref{AuThiolToluene}. These results do not suggest that $G$ is
729 dependent on $J_z$ within this flux range. The linear response of flux
730 to thermal gradient simplifies our investigations in that we can rely
731 on $G$ measurement with only a couple $J_z$'s and do not need to test
732 a large series of fluxes.
733
734 \begin{table*}
735 \begin{minipage}{\linewidth}
736 \begin{center}
737 \caption{Computed interfacial thermal conductivity ($G$ and
738 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
739 interfaces with UA model and different hexane molecule numbers
740 at different temperatures using a range of energy
741 fluxes. Error estimates indicated in parenthesis.}
742
743 \begin{tabular}{ccccccc}
744 \hline\hline
745 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
746 $J_z$ & $G$ & $G^\prime$ \\
747 (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
748 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
749 \hline
750 200 & 266 & No & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
751 & 200 & Yes & 0.694 & 1.92 & 129(11) & 87.3(0.3) \\
752 & & Yes & 0.672 & 1.93 & 131(16) & 78(13) \\
753 & & No & 0.688 & 0.96 & 125(16) & 90.2(15) \\
754 & & & & 1.91 & 139(10) & 101(10) \\
755 & & & & 2.83 & 141(6) & 89.9(9.8) \\
756 & 166 & Yes & 0.679 & 0.97 & 115(19) & 69(18) \\
757 & & & & 1.94 & 125(9) & 87.1(0.2) \\
758 & & No & 0.681 & 0.97 & 141(30) & 78(22) \\
759 & & & & 1.92 & 138(4) & 98.9(9.5) \\
760 \hline
761 250 & 200 & No & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
762 & & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
763 & 166 & Yes & 0.570 & 0.98 & 79.0(3.5) & 62.9(3.0) \\
764 & & No & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
765 & & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
766 & & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
767 & & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
768 \hline\hline
769 \end{tabular}
770 \label{AuThiolHexaneUA}
771 \end{center}
772 \end{minipage}
773 \end{table*}
774
775 \subsubsection{Effects due to average temperature}
776
777 Furthermore, we also attempted to increase system average temperatures
778 to above 200K. These simulations are first equilibrated in the NPT
779 ensemble under normal pressure. As stated above, the TraPPE-UA model
780 for hexane tends to predict a lower boiling point. In our simulations,
781 hexane had diffculty to remain in liquid phase when NPT equilibration
782 temperature is higher than 250K. Additionally, the equilibrated liquid
783 hexane density under 250K becomes lower than experimental value. This
784 expanded liquid phase leads to lower contact between hexane and
785 butanethiol as well.[MAY NEED SLAB DENSITY FIGURE]
786 And this reduced contact would
787 probably be accountable for a lower interfacial thermal conductance,
788 as shown in Table \ref{AuThiolHexaneUA}.
789
790 A similar study for TraPPE-UA toluene agrees with the above result as
791 well. Having a higher boiling point, toluene tends to remain liquid in
792 our simulations even equilibrated under 300K in NPT
793 ensembles. Furthermore, the expansion of the toluene liquid phase is
794 not as significant as that of the hexane. This prevents severe
795 decrease of liquid-capping agent contact and the results (Table
796 \ref{AuThiolToluene}) show only a slightly decreased interface
797 conductance. Therefore, solvent-capping agent contact should play an
798 important role in the thermal transport process across the interface
799 in that higher degree of contact could yield increased conductance.
800
801 \begin{table*}
802 \begin{minipage}{\linewidth}
803 \begin{center}
804 \caption{Computed interfacial thermal conductivity ($G$ and
805 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
806 interface at different temperatures using a range of energy
807 fluxes. Error estimates indicated in parenthesis.}
808
809 \begin{tabular}{ccccc}
810 \hline\hline
811 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
812 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
813 \hline
814 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
815 & & -1.86 & 180(3) & 135(21) \\
816 & & -3.93 & 176(5) & 113(12) \\
817 \hline
818 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
819 & & -4.19 & 135(9) & 113(12) \\
820 \hline\hline
821 \end{tabular}
822 \label{AuThiolToluene}
823 \end{center}
824 \end{minipage}
825 \end{table*}
826
827 Besides lower interfacial thermal conductance, surfaces in relatively
828 high temperatures are susceptible to reconstructions, when
829 butanethiols have a full coverage on the Au(111) surface. These
830 reconstructions include surface Au atoms migrated outward to the S
831 atom layer, and butanethiol molecules embedded into the original
832 surface Au layer. The driving force for this behavior is the strong
833 Au-S interactions in our simulations. And these reconstructions lead
834 to higher ratio of Au-S attraction and thus is energetically
835 favorable. Furthermore, this phenomenon agrees with experimental
836 results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
837 {\it et al.} had kept their Au(111) slab rigid so that their
838 simulations can reach 300K without surface reconstructions. Without
839 this practice, simulating 100\% thiol covered interfaces under higher
840 temperatures could hardly avoid surface reconstructions. However, our
841 measurement is based on assuming homogeneity on $x$ and $y$ dimensions
842 so that measurement of $T$ at particular $z$ would be an effective
843 average of the particles of the same type. Since surface
844 reconstructions could eliminate the original $x$ and $y$ dimensional
845 homogeneity, measurement of $G$ is more difficult to conduct under
846 higher temperatures. Therefore, most of our measurements are
847 undertaken at $\langle T\rangle\sim$200K.
848
849 However, when the surface is not completely covered by butanethiols,
850 the simulated system is more resistent to the reconstruction
851 above. Our Au-butanethiol/toluene system had the Au(111) surfaces 90\%
852 covered by butanethiols, but did not see this above phenomena even at
853 $\langle T\rangle\sim$300K. The empty three-fold sites not occupied by
854 capping agents could help prevent surface reconstruction in that they
855 provide other means of capping agent relaxation. It is observed that
856 butanethiols can migrate to their neighbor empty sites during a
857 simulation. Therefore, we were able to obtain $G$'s for these
858 interfaces even at a relatively high temperature without being
859 affected by surface reconstructions.
860
861
862 \section{Discussion}
863
864 \subsection{Capping agent acts as a vibrational coupler between solid
865 and solvent phases}
866 To investigate the mechanism of interfacial thermal conductance, the
867 vibrational power spectrum was computed. Power spectra were taken for
868 individual components in different simulations. To obtain these
869 spectra, simulations were run after equilibration, in the NVE
870 ensemble, and without a thermal gradient. Snapshots of configurations
871 were collected at a frequency that is higher than that of the fastest
872 vibrations occuring in the simulations. With these configurations, the
873 velocity auto-correlation functions can be computed:
874 \begin{equation}
875 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
876 \label{vCorr}
877 \end{equation}
878 The power spectrum is constructed via a Fourier transform of the
879 symmetrized velocity autocorrelation function,
880 \begin{equation}
881 \hat{f}(\omega) =
882 \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
883 \label{fourier}
884 \end{equation}
885
886 From Figure \ref{coverage}, one can see the significance of the
887 presence of capping agents. Even when a fraction of the Au(111)
888 surface sites are covered with butanethiols, the conductivity would
889 see an enhancement by at least a factor of 3. This indicates the
890 important role cappping agent is playing for thermal transport
891 phenomena on metal / organic solvent surfaces.
892
893 Interestingly, as one could observe from our results, the maximum
894 conductance enhancement (largest $G$) happens while the surfaces are
895 about 75\% covered with butanethiols. This again indicates that
896 solvent-capping agent contact has an important role of the thermal
897 transport process. Slightly lower butanethiol coverage allows small
898 gaps between butanethiols to form. And these gaps could be filled with
899 solvent molecules, which acts like ``heat conductors'' on the
900 surface. The higher degree of interaction between these solvent
901 molecules and capping agents increases the enhancement effect and thus
902 produces a higher $G$ than densely packed butanethiol arrays. However,
903 once this maximum conductance enhancement is reached, $G$ decreases
904 when butanethiol coverage continues to decrease. Each capping agent
905 molecule reaches its maximum capacity for thermal
906 conductance. Therefore, even higher solvent-capping agent contact
907 would not offset this effect. Eventually, when butanethiol coverage
908 continues to decrease, solvent-capping agent contact actually
909 decreases with the disappearing of butanethiol molecules. In this
910 case, $G$ decrease could not be offset but instead accelerated. [MAY NEED
911 SNAPSHOT SHOWING THE PHENOMENA / SLAB DENSITY ANALYSIS]
912
913 A comparison of the results obtained from differenet organic solvents
914 can also provide useful information of the interfacial thermal
915 transport process. The deuterated hexane (UA) results do not appear to
916 be much different from those of normal hexane (UA), given that
917 butanethiol (UA) is non-deuterated for both solvents. These UA model
918 studies, even though eliminating C-H vibration samplings, still have
919 C-C vibrational frequencies different from each other. However, these
920 differences in the infrared range do not seem to produce an observable
921 difference for the results of $G$ (Figure \ref{uahxnua}).
922
923 \begin{figure}
924 \includegraphics[width=\linewidth]{uahxnua}
925 \caption{Vibrational spectra obtained for normal (upper) and
926 deuterated (lower) hexane in Au-butanethiol/hexane
927 systems. Butanethiol spectra are shown as reference. Both hexane and
928 butanethiol were using United-Atom models.}
929 \label{uahxnua}
930 \end{figure}
931
932 Furthermore, results for rigid body toluene solvent, as well as other
933 UA-hexane solvents, are reasonable within the general experimental
934 ranges\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406}. This
935 suggests that explicit hydrogen might not be a required factor for
936 modeling thermal transport phenomena of systems such as
937 Au-thiol/organic solvent.
938
939 However, results for Au-butanethiol/toluene do not show an identical
940 trend with those for Au-butanethiol/hexane in that $G$ remains at
941 approximately the same magnitue when butanethiol coverage differs from
942 25\% to 75\%. This might be rooted in the molecule shape difference
943 for planar toluene and chain-like {\it n}-hexane. Due to this
944 difference, toluene molecules have more difficulty in occupying
945 relatively small gaps among capping agents when their coverage is not
946 too low. Therefore, the solvent-capping agent contact may keep
947 increasing until the capping agent coverage reaches a relatively low
948 level. This becomes an offset for decreasing butanethiol molecules on
949 its effect to the process of interfacial thermal transport. Thus, one
950 can see a plateau of $G$ vs. butanethiol coverage in our results.
951
952 \subsection{Influence of Chosen Molecule Model on $G$}
953 In addition to UA solvent/capping agent models, AA models are included
954 in our simulations as well. Besides simulations of the same (UA or AA)
955 model for solvent and capping agent, different models can be applied
956 to different components. Furthermore, regardless of models chosen,
957 either the solvent or the capping agent can be deuterated, similar to
958 the previous section. Table \ref{modelTest} summarizes the results of
959 these studies.
960
961 \begin{table*}
962 \begin{minipage}{\linewidth}
963 \begin{center}
964
965 \caption{Computed interfacial thermal conductivity ($G$ and
966 $G^\prime$) values for interfaces using various models for
967 solvent and capping agent (or without capping agent) at
968 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
969 or capping agent molecules; ``Avg.'' denotes results that are
970 averages of simulations under different $J_z$'s. Error
971 estimates indicated in parenthesis.)}
972
973 \begin{tabular}{llccc}
974 \hline\hline
975 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
976 (or bare surface) & model & (GW/m$^2$) &
977 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
978 \hline
979 UA & UA hexane & Avg. & 131(9) & 87(10) \\
980 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
981 & AA hexane & Avg. & 131(6) & 122(10) \\
982 & UA toluene & 1.96 & 187(16) & 151(11) \\
983 & AA toluene & 1.89 & 200(36) & 149(53) \\
984 \hline
985 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
986 & AA hexane & Avg. & 442(14) & 356(31) \\
987 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
988 & UA toluene & 1.98 & 125(25) & 97(60) \\
989 & AA toluene & 3.79 & 487(56) & 290(42) \\
990 \hline
991 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
992 & AA hexane & 1.92 & 243(29) & 191(11) \\
993 & AA toluene & 1.93 & 364(36) & 322(67) \\
994 \hline
995 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
996 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
997 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
998 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
999 \hline\hline
1000 \end{tabular}
1001 \label{modelTest}
1002 \end{center}
1003 \end{minipage}
1004 \end{table*}
1005
1006 To facilitate direct comparison, the same system with differnt models
1007 for different components uses the same length scale for their
1008 simulation cells. Without the presence of capping agent, using
1009 different models for hexane yields similar results for both $G$ and
1010 $G^\prime$, and these two definitions agree with eath other very
1011 well. This indicates very weak interaction between the metal and the
1012 solvent, and is a typical case for acoustic impedance mismatch between
1013 these two phases.
1014
1015 As for Au(111) surfaces completely covered by butanethiols, the choice
1016 of models for capping agent and solvent could impact the measurement
1017 of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
1018 interfaces, using AA model for both butanethiol and hexane yields
1019 substantially higher conductivity values than using UA model for at
1020 least one component of the solvent and capping agent, which exceeds
1021 the general range of experimental measurement results. This is
1022 probably due to the classically treated C-H vibrations in the AA
1023 model, which should not be appreciably populated at normal
1024 temperatures. In comparison, once either the hexanes or the
1025 butanethiols are deuterated, one can see a significantly lower $G$ and
1026 $G^\prime$. In either of these cases, the C-H(D) vibrational overlap
1027 between the solvent and the capping agent is removed (Figure
1028 \ref{aahxntln}). Conclusively, the improperly treated C-H vibration in
1029 the AA model produced over-predicted results accordingly. Compared to
1030 the AA model, the UA model yields more reasonable results with higher
1031 computational efficiency.
1032
1033 \begin{figure}
1034 \includegraphics[width=\linewidth]{aahxntln}
1035 \caption{Spectra obtained for All-Atom model Au-butanethil/solvent
1036 systems. When butanethiol is deuterated (lower left), its
1037 vibrational overlap with hexane would decrease significantly,
1038 compared with normal butanethiol (upper left). However, this
1039 dramatic change does not apply to toluene as much (right).}
1040 \label{aahxntln}
1041 \end{figure}
1042
1043 However, for Au-butanethiol/toluene interfaces, having the AA
1044 butanethiol deuterated did not yield a significant change in the
1045 measurement results. Compared to the C-H vibrational overlap between
1046 hexane and butanethiol, both of which have alkyl chains, that overlap
1047 between toluene and butanethiol is not so significant and thus does
1048 not have as much contribution to the heat exchange
1049 process. Conversely, extra degrees of freedom such as the C-H
1050 vibrations could yield higher heat exchange rate between these two
1051 phases and result in a much higher conductivity.
1052
1053 Although the QSC model for Au is known to predict an overly low value
1054 for bulk metal gold conductivity\cite{kuang:164101}, our computational
1055 results for $G$ and $G^\prime$ do not seem to be affected by this
1056 drawback of the model for metal. Instead, our results suggest that the
1057 modeling of interfacial thermal transport behavior relies mainly on
1058 the accuracy of the interaction descriptions between components
1059 occupying the interfaces.
1060
1061 \subsection{Role of Capping Agent in Interfacial Thermal Conductance}
1062 The vibrational spectra for gold slabs in different environments are
1063 shown as in Figure \ref{specAu}. Regardless of the presence of
1064 solvent, the gold surfaces covered by butanethiol molecules, compared
1065 to bare gold surfaces, exhibit an additional peak observed at the
1066 frequency of $\sim$170cm$^{-1}$, which is attributed to the S-Au
1067 bonding vibration. This vibration enables efficient thermal transport
1068 from surface Au layer to the capping agents. Therefore, in our
1069 simulations, the Au/S interfaces do not appear major heat barriers
1070 compared to the butanethiol / solvent interfaces.
1071
1072 \subsubsection{Overlap of power spectrum}
1073 Simultaneously, the vibrational overlap between butanethiol and
1074 organic solvents suggests higher thermal exchange efficiency between
1075 these two components. Even exessively high heat transport was observed
1076 when All-Atom models were used and C-H vibrations were treated
1077 classically. Compared to metal and organic liquid phase, the heat
1078 transfer efficiency between butanethiol and organic solvents is closer
1079 to that within bulk liquid phase.
1080
1081 Furthermore, our observation validated previous
1082 results\cite{hase:2010} that the intramolecular heat transport of
1083 alkylthiols is highly effecient. As a combinational effects of these
1084 phenomena, butanethiol acts as a channel to expedite thermal transport
1085 process. The acoustic impedance mismatch between the metal and the
1086 liquid phase can be effectively reduced with the presence of suitable
1087 capping agents.
1088
1089 \begin{figure}
1090 \includegraphics[width=\linewidth]{vibration}
1091 \caption{Vibrational spectra obtained for gold in different
1092 environments.}
1093 \label{specAu}
1094 \end{figure}
1095
1096 [MAY ADD COMPARISON OF AU SLAB WIDTHS BUT NOT MUCH TO TALK ABOUT...]
1097
1098 \section{Conclusions}
1099 The NIVS algorithm we developed has been applied to simulations of
1100 Au-butanethiol surfaces with organic solvents. This algorithm allows
1101 effective unphysical thermal flux transferred between the metal and
1102 the liquid phase. With the flux applied, we were able to measure the
1103 corresponding thermal gradient and to obtain interfacial thermal
1104 conductivities. Under steady states, single trajectory simulation
1105 would be enough for accurate measurement. This would be advantageous
1106 compared to transient state simulations, which need multiple
1107 trajectories to produce reliable average results.
1108
1109 Our simulations have seen significant conductance enhancement with the
1110 presence of capping agent, compared to the bare gold / liquid
1111 interfaces. The acoustic impedance mismatch between the metal and the
1112 liquid phase is effectively eliminated by proper capping
1113 agent. Furthermore, the coverage precentage of the capping agent plays
1114 an important role in the interfacial thermal transport
1115 process. Moderately lower coverages allow higher contact between
1116 capping agent and solvent, and thus could further enhance the heat
1117 transfer process.
1118
1119 Our measurement results, particularly of the UA models, agree with
1120 available experimental data. This indicates that our force field
1121 parameters have a nice description of the interactions between the
1122 particles at the interfaces. AA models tend to overestimate the
1123 interfacial thermal conductance in that the classically treated C-H
1124 vibration would be overly sampled. Compared to the AA models, the UA
1125 models have higher computational efficiency with satisfactory
1126 accuracy, and thus are preferable in interfacial thermal transport
1127 modelings. Of the two definitions for $G$, the discrete form
1128 (Eq. \ref{discreteG}) was easier to use and gives out relatively
1129 consistent results, while the derivative form (Eq. \ref{derivativeG})
1130 is not as versatile. Although $G^\prime$ gives out comparable results
1131 and follows similar trend with $G$ when measuring close to fully
1132 covered or bare surfaces, the spatial resolution of $T$ profile is
1133 limited for accurate computation of derivatives data.
1134
1135 Vlugt {\it et al.} has investigated the surface thiol structures for
1136 nanocrystal gold and pointed out that they differs from those of the
1137 Au(111) surface\cite{landman:1998,vlugt:cpc2007154}. This difference
1138 might lead to change of interfacial thermal transport behavior as
1139 well. To investigate this problem, an effective means to introduce
1140 thermal flux and measure the corresponding thermal gradient is
1141 desirable for simulating structures with spherical symmetry.
1142
1143 \section{Acknowledgments}
1144 Support for this project was provided by the National Science
1145 Foundation under grant CHE-0848243. Computational time was provided by
1146 the Center for Research Computing (CRC) at the University of Notre
1147 Dame.
1148 \newpage
1149
1150 \bibliography{interfacial}
1151
1152 \end{doublespace}
1153 \end{document}
1154