ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3764
Committed: Wed Sep 28 22:26:23 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 52926 byte(s)
Log Message:
more revisions according to reviewer comments.

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 With the Non-Isotropic Velocity Scaling (NIVS) approach to Reverse
48 Non-Equilibrium Molecular Dynamics (RNEMD) it is possible to impose
49 an unphysical thermal flux between different regions of
50 inhomogeneous systems such as solid / liquid interfaces. We have
51 applied NIVS to compute the interfacial thermal conductance at a
52 metal / organic solvent interface that has been chemically capped by
53 butanethiol molecules. Our calculations suggest that the acoustic
54 impedance mismatch between the metal and liquid phases is
55 effectively reduced by the capping agents, leading to a greatly
56 enhanced conductivity at the interface. Specifically, the chemical
57 bond between the metal and the capping agent introduces a
58 vibrational overlap that is not present without the capping agent,
59 and the overlap between the vibrational spectra (metal to cap, cap
60 to solvent) provides a mechanism for rapid thermal transport across
61 the interface. Our calculations also suggest that this is a
62 non-monotonic function of the fractional coverage of the surface, as
63 moderate coverages allow {\bf vibrational heat diffusion} of solvent
64 molecules that have been in close contact with the capping agent.
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 (and heat removal) in
77 nanotechnology, interfacial thermal conductance has been studied
78 extensively both experimentally and computationally.\cite{cahill:793}
79 Nanoscale materials have a significant fraction of their atoms at
80 interfaces, and the chemical details of these interfaces govern the
81 thermal transport properties. Furthermore, the interfaces are often
82 heterogeneous (e.g. solid - liquid), which provides a challenge to
83 computational methods which have been developed for homogeneous or
84 bulk systems.
85
86 Experimentally, the thermal properties of a number of interfaces have
87 been investigated. Cahill and coworkers studied nanoscale thermal
88 transport from metal nanoparticle/fluid interfaces, to epitaxial
89 TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
90 interfaces between water and solids with different self-assembled
91 monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
92 Wang {\it et al.} studied heat transport through long-chain
93 hydrocarbon monolayers on gold substrate at individual molecular
94 level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
95 cetyltrimethylammonium bromide (CTAB) on the thermal transport between
96 gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
97 et al.} studied the cooling dynamics, which is controlled by thermal
98 interface resistance of glass-embedded metal
99 nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
100 normally considered barriers for heat transport, Alper {\it et al.}
101 suggested that specific ligands (capping agents) could completely
102 eliminate this barrier
103 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
104
105 Theoretical and computational models have also been used to study the
106 interfacial thermal transport in order to gain an understanding of
107 this phenomena at the molecular level. Recently, Hase and coworkers
108 employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
109 study thermal transport from hot Au(111) substrate to a self-assembled
110 monolayer of alkylthiol with relatively long chain (8-20 carbon
111 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
112 measurements for heat conductance of interfaces between the capping
113 monolayer on Au and a solvent phase have yet to be studied with their
114 approach. The comparatively low thermal flux through interfaces is
115 difficult to measure with Equilibrium
116 MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
117 methods. Therefore, the Reverse NEMD (RNEMD)
118 methods\cite{MullerPlathe:1997xw,kuang:164101} would be advantageous
119 in that they {\it apply} the difficult to measure quantity (flux),
120 while {\it measuring} the easily-computed quantity (the thermal
121 gradient). This is particularly true for inhomogeneous interfaces
122 where it would not be clear how to apply a gradient {\it a priori}.
123 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
124 this approach to various liquid interfaces and studied how thermal
125 conductance (or resistance) is dependent on chemical details of a
126 number of hydrophobic and hydrophilic aqueous interfaces. {\bf And
127 Luo {\it et al.} studied the thermal conductance of Au-SAM-Au
128 junctions using the same approach, with comparison to a constant
129 temperature difference method\cite{Luo20101}. While this latter
130 approach establishes more thermal distributions compared to the
131 former RNEMD methods, it does not guarantee momentum or kinetic
132 energy conservations.}
133
134 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
135 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
136 retains the desirable features of RNEMD (conservation of linear
137 momentum and total energy, compatibility with periodic boundary
138 conditions) while establishing true thermal distributions in each of
139 the two slabs. Furthermore, it allows effective thermal exchange
140 between particles of different identities, and thus makes the study of
141 interfacial conductance much simpler.
142
143 The work presented here deals with the Au(111) surface covered to
144 varying degrees by butanethiol, a capping agent with short carbon
145 chain, and solvated with organic solvents of different molecular
146 properties. {\bf To our knowledge, few previous MD inverstigations
147 have been found to address to these systems yet.} Different models
148 were used for both the capping agent and the solvent force field
149 parameters. Using the NIVS algorithm, the thermal transport across
150 these interfaces was studied and the underlying mechanism for the
151 phenomena was investigated.
152
153 \section{Methodology}
154 \subsection{Imposed-Flux Methods in MD Simulations}
155 Steady state MD simulations have an advantage in that not many
156 trajectories are needed to study the relationship between thermal flux
157 and thermal gradients. For systems with low interfacial conductance,
158 one must have a method capable of generating or measuring relatively
159 small fluxes, compared to those required for bulk conductivity. This
160 requirement makes the calculation even more difficult for
161 slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
162 NEMD methods impose a gradient (and measure a flux), but at interfaces
163 it is not clear what behavior should be imposed at the boundaries
164 between materials. Imposed-flux reverse non-equilibrium
165 methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
166 the thermal response becomes an easy-to-measure quantity. Although
167 M\"{u}ller-Plathe's original momentum swapping approach can be used
168 for exchanging energy between particles of different identity, the
169 kinetic energy transfer efficiency is affected by the mass difference
170 between the particles, which limits its application on heterogeneous
171 interfacial systems.
172
173 The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
174 to non-equilibrium MD simulations is able to impose a wide range of
175 kinetic energy fluxes without obvious perturbation to the velocity
176 distributions of the simulated systems. Furthermore, this approach has
177 the advantage in heterogeneous interfaces in that kinetic energy flux
178 can be applied between regions of particles of arbitrary identity, and
179 the flux will not be restricted by difference in particle mass.
180
181 The NIVS algorithm scales the velocity vectors in two separate regions
182 of a simulation system with respective diagonal scaling matrices. To
183 determine these scaling factors in the matrices, a set of equations
184 including linear momentum conservation and kinetic energy conservation
185 constraints and target energy flux satisfaction is solved. With the
186 scaling operation applied to the system in a set frequency, bulk
187 temperature gradients can be easily established, and these can be used
188 for computing thermal conductivities. The NIVS algorithm conserves
189 momenta and energy and does not depend on an external thermostat.
190
191 \subsection{Defining Interfacial Thermal Conductivity ($G$)}
192
193 For an interface with relatively low interfacial conductance, and a
194 thermal flux between two distinct bulk regions, the regions on either
195 side of the interface rapidly come to a state in which the two phases
196 have relatively homogeneous (but distinct) temperatures. The
197 interfacial thermal conductivity $G$ can therefore be approximated as:
198 \begin{equation}
199 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
200 \langle T_\mathrm{cold}\rangle \right)}
201 \label{lowG}
202 \end{equation}
203 where ${E_{total}}$ is the total imposed non-physical kinetic energy
204 transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
205 and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
206 temperature of the two separated phases. For an applied flux $J_z$
207 operating over a simulation time $t$ on a periodically-replicated slab
208 of dimensions $L_x \times L_y$, $E_{total} = J_z *(t)*(2 L_x L_y)$.
209
210 When the interfacial conductance is {\it not} small, there are two
211 ways to define $G$. One common way is to assume the temperature is
212 discrete on the two sides of the interface. $G$ can be calculated
213 using the applied thermal flux $J$ and the maximum temperature
214 difference measured along the thermal gradient max($\Delta T$), which
215 occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
216 known as the Kapitza conductance, which is the inverse of the Kapitza
217 resistance.
218 \begin{equation}
219 G=\frac{J}{\Delta T}
220 \label{discreteG}
221 \end{equation}
222
223 \begin{figure}
224 \includegraphics[width=\linewidth]{method}
225 \caption{Interfacial conductance can be calculated by applying an
226 (unphysical) kinetic energy flux between two slabs, one located
227 within the metal and another on the edge of the periodic box. The
228 system responds by forming a thermal gradient. In bulk liquids,
229 this gradient typically has a single slope, but in interfacial
230 systems, there are distinct thermal conductivity domains. The
231 interfacial conductance, $G$ is found by measuring the temperature
232 gap at the Gibbs dividing surface, or by using second derivatives of
233 the thermal profile.}
234 \label{demoPic}
235 \end{figure}
236
237 {\bf We attempt another approach by assuming that temperature is
238 continuous and differentiable throughout the space. Given that
239 $\lambda$ is also differentiable, $G$ can be defined as its
240 gradient. This quantity has the same unit as the commonly known $G$,
241 and the maximum of its magnitude denotes where thermal conductivity
242 has the largest change, i.e. the interface. And vector
243 $\nabla\lambda$ is normal to the interface. In a simplified
244 condition here, we have both $\vec{J}$ and the thermal gradient
245 paralell to the $z$ axis and yield the formula used in our
246 computations.}
247 (original text)
248 The other approach is to assume a continuous temperature profile along
249 the thermal gradient axis (e.g. $z$) and define $G$ at the point where
250 the magnitude of thermal conductivity ($\lambda$) change reaches its
251 maximum, given that $\lambda$ is well-defined throughout the space:
252 \begin{equation}
253 G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
254 = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
255 \left(\frac{\partial T}{\partial z}\right)\right)\Big|
256 = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
257 \Big/\left(\frac{\partial T}{\partial z}\right)^2
258 \label{derivativeG}
259 \end{equation}
260
261 With temperature profiles obtained from simulation, one is able to
262 approximate the first and second derivatives of $T$ with finite
263 difference methods and calculate $G^\prime$. In what follows, both
264 definitions have been used, and are compared in the results.
265
266 To investigate the interfacial conductivity at metal / solvent
267 interfaces, we have modeled a metal slab with its (111) surfaces
268 perpendicular to the $z$-axis of our simulation cells. The metal slab
269 has been prepared both with and without capping agents on the exposed
270 surface, and has been solvated with simple organic solvents, as
271 illustrated in Figure \ref{gradT}.
272
273 With the simulation cell described above, we are able to equilibrate
274 the system and impose an unphysical thermal flux between the liquid
275 and the metal phase using the NIVS algorithm. By periodically applying
276 the unphysical flux, we obtained a temperature profile and its spatial
277 derivatives. Figure \ref{gradT} shows how an applied thermal flux can
278 be used to obtain the 1st and 2nd derivatives of the temperature
279 profile.
280
281 \begin{figure}
282 \includegraphics[width=\linewidth]{gradT}
283 \caption{A sample of Au (111) / butanethiol / hexane interfacial
284 system with the temperature profile after a kinetic energy flux has
285 been imposed. Note that the largest temperature jump in the thermal
286 profile (corresponding to the lowest interfacial conductance) is at
287 the interface between the butanethiol molecules (blue) and the
288 solvent (grey). First and second derivatives of the temperature
289 profile are obtained using a finite difference approximation (lower
290 panel).}
291 \label{gradT}
292 \end{figure}
293
294 \section{Computational Details}
295 \subsection{Simulation Protocol}
296 The NIVS algorithm has been implemented in our MD simulation code,
297 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
298 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
299 under atmospheric pressure (1 atm) and 200K. After equilibration,
300 butanethiol capping agents were placed at three-fold hollow sites on
301 the Au(111) surfaces. These sites are either {\it fcc} or {\it
302 hcp} sites, although Hase {\it et al.} found that they are
303 equivalent in a heat transfer process,\cite{hase:2010} so we did not
304 distinguish between these sites in our study. The maximum butanethiol
305 capacity on Au surface is $1/3$ of the total number of surface Au
306 atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
307 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
308 series of lower coverages was also prepared by eliminating
309 butanethiols from the higher coverage surface in a regular manner. The
310 lower coverages were prepared in order to study the relation between
311 coverage and interfacial conductance.
312
313 The capping agent molecules were allowed to migrate during the
314 simulations. They distributed themselves uniformly and sampled a
315 number of three-fold sites throughout out study. Therefore, the
316 initial configuration does not noticeably affect the sampling of a
317 variety of configurations of the same coverage, and the final
318 conductance measurement would be an average effect of these
319 configurations explored in the simulations.
320
321 After the modified Au-butanethiol surface systems were equilibrated in
322 the canonical (NVT) ensemble, organic solvent molecules were packed in
323 the previously empty part of the simulation cells.\cite{packmol} Two
324 solvents were investigated, one which has little vibrational overlap
325 with the alkanethiol and which has a planar shape (toluene), and one
326 which has similar vibrational frequencies to the capping agent and
327 chain-like shape ({\it n}-hexane).
328
329 The simulation cells were not particularly extensive along the
330 $z$-axis, as a very long length scale for the thermal gradient may
331 cause excessively hot or cold temperatures in the middle of the
332 solvent region and lead to undesired phenomena such as solvent boiling
333 or freezing when a thermal flux is applied. Conversely, too few
334 solvent molecules would change the normal behavior of the liquid
335 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
336 these extreme cases did not happen to our simulations. The spacing
337 between periodic images of the gold interfaces is $45 \sim 75$\AA in
338 our simulations.
339
340 The initial configurations generated are further equilibrated with the
341 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
342 change. This is to ensure that the equilibration of liquid phase does
343 not affect the metal's crystalline structure. Comparisons were made
344 with simulations that allowed changes of $L_x$ and $L_y$ during NPT
345 equilibration. No substantial changes in the box geometry were noticed
346 in these simulations. After ensuring the liquid phase reaches
347 equilibrium at atmospheric pressure (1 atm), further equilibration was
348 carried out under canonical (NVT) and microcanonical (NVE) ensembles.
349
350 After the systems reach equilibrium, NIVS was used to impose an
351 unphysical thermal flux between the metal and the liquid phases. Most
352 of our simulations were done under an average temperature of
353 $\sim$200K. Therefore, thermal flux usually came from the metal to the
354 liquid so that the liquid has a higher temperature and would not
355 freeze due to lowered temperatures. After this induced temperature
356 gradient had stabilized, the temperature profile of the simulation cell
357 was recorded. To do this, the simulation cell is divided evenly into
358 $N$ slabs along the $z$-axis. The average temperatures of each slab
359 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
360 the same, the derivatives of $T$ with respect to slab number $n$ can
361 be directly used for $G^\prime$ calculations: \begin{equation}
362 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
363 \Big/\left(\frac{\partial T}{\partial z}\right)^2
364 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
365 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
366 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
367 \Big/\left(\frac{\partial T}{\partial n}\right)^2
368 \label{derivativeG2}
369 \end{equation}
370
371 All of the above simulation procedures use a time step of 1 fs. Each
372 equilibration stage took a minimum of 100 ps, although in some cases,
373 longer equilibration stages were utilized.
374
375 \subsection{Force Field Parameters}
376 Our simulations include a number of chemically distinct components.
377 Figure \ref{demoMol} demonstrates the sites defined for both
378 United-Atom and All-Atom models of the organic solvent and capping
379 agents in our simulations. Force field parameters are needed for
380 interactions both between the same type of particles and between
381 particles of different species.
382
383 \begin{figure}
384 \includegraphics[width=\linewidth]{structures}
385 \caption{Structures of the capping agent and solvents utilized in
386 these simulations. The chemically-distinct sites (a-e) are expanded
387 in terms of constituent atoms for both United Atom (UA) and All Atom
388 (AA) force fields. Most parameters are from References
389 \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
390 (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
391 atoms are given in Table \ref{MnM}.}
392 \label{demoMol}
393 \end{figure}
394
395 The Au-Au interactions in metal lattice slab is described by the
396 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
397 potentials include zero-point quantum corrections and are
398 reparametrized for accurate surface energies compared to the
399 Sutton-Chen potentials.\cite{Chen90}
400
401 For the two solvent molecules, {\it n}-hexane and toluene, two
402 different atomistic models were utilized. Both solvents were modeled
403 using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
404 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
405 for our UA solvent molecules. In these models, sites are located at
406 the carbon centers for alkyl groups. Bonding interactions, including
407 bond stretches and bends and torsions, were used for intra-molecular
408 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
409 potentials are used.
410
411 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
412 simple and computationally efficient, while maintaining good accuracy.
413 However, the TraPPE-UA model for alkanes is known to predict a slightly
414 lower boiling point than experimental values. This is one of the
415 reasons we used a lower average temperature (200K) for our
416 simulations. If heat is transferred to the liquid phase during the
417 NIVS simulation, the liquid in the hot slab can actually be
418 substantially warmer than the mean temperature in the simulation. The
419 lower mean temperatures therefore prevent solvent boiling.
420
421 For UA-toluene, the non-bonded potentials between intermolecular sites
422 have a similar Lennard-Jones formulation. The toluene molecules were
423 treated as a single rigid body, so there was no need for
424 intramolecular interactions (including bonds, bends, or torsions) in
425 this solvent model.
426
427 Besides the TraPPE-UA models, AA models for both organic solvents are
428 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
429 were used. For hexane, additional explicit hydrogen sites were
430 included. Besides bonding and non-bonded site-site interactions,
431 partial charges and the electrostatic interactions were added to each
432 CT and HC site. For toluene, a flexible model for the toluene molecule
433 was utilized which included bond, bend, torsion, and inversion
434 potentials to enforce ring planarity.
435
436 The butanethiol capping agent in our simulations, were also modeled
437 with both UA and AA model. The TraPPE-UA force field includes
438 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
439 UA butanethiol model in our simulations. The OPLS-AA also provides
440 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
441 surfaces do not have the hydrogen atom bonded to sulfur. To derive
442 suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
443 adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
444 modify the parameters for the CTS atom to maintain charge neutrality
445 in the molecule. Note that the model choice (UA or AA) for the capping
446 agent can be different from the solvent. Regardless of model choice,
447 the force field parameters for interactions between capping agent and
448 solvent can be derived using Lorentz-Berthelot Mixing Rule:
449 \begin{eqnarray}
450 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
451 \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
452 \end{eqnarray}
453
454 To describe the interactions between metal (Au) and non-metal atoms,
455 we refer to an adsorption study of alkyl thiols on gold surfaces by
456 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
457 Lennard-Jones form of potential parameters for the interaction between
458 Au and pseudo-atoms CH$_x$ and S based on a well-established and
459 widely-used effective potential of Hautman and Klein for the Au(111)
460 surface.\cite{hautman:4994} As our simulations require the gold slab
461 to be flexible to accommodate thermal excitation, the pair-wise form
462 of potentials they developed was used for our study.
463
464 The potentials developed from {\it ab initio} calculations by Leng
465 {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
466 interactions between Au and aromatic C/H atoms in toluene. However,
467 the Lennard-Jones parameters between Au and other types of particles,
468 (e.g. AA alkanes) have not yet been established. For these
469 interactions, the Lorentz-Berthelot mixing rule can be used to derive
470 effective single-atom LJ parameters for the metal using the fit values
471 for toluene. These are then used to construct reasonable mixing
472 parameters for the interactions between the gold and other atoms.
473 Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
474 our simulations.
475
476 \begin{table*}
477 \begin{minipage}{\linewidth}
478 \begin{center}
479 \caption{Non-bonded interaction parameters (including cross
480 interactions with Au atoms) for both force fields used in this
481 work.}
482 \begin{tabular}{lllllll}
483 \hline\hline
484 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
485 $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
486 & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
487 \hline
488 United Atom (UA)
489 &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
490 &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
491 &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
492 &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
493 \hline
494 All Atom (AA)
495 &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
496 &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
497 &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
498 &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
499 &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
500 &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
501 \hline
502 Both UA and AA
503 & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
504 \hline\hline
505 \end{tabular}
506 \label{MnM}
507 \end{center}
508 \end{minipage}
509 \end{table*}
510
511
512 \section{Results}
513 There are many factors contributing to the measured interfacial
514 conductance; some of these factors are physically motivated
515 (e.g. coverage of the surface by the capping agent coverage and
516 solvent identity), while some are governed by parameters of the
517 methodology (e.g. applied flux and the formulas used to obtain the
518 conductance). In this section we discuss the major physical and
519 calculational effects on the computed conductivity.
520
521 \subsection{Effects due to capping agent coverage}
522
523 A series of different initial conditions with a range of surface
524 coverages was prepared and solvated with various with both of the
525 solvent molecules. These systems were then equilibrated and their
526 interfacial thermal conductivity was measured with the NIVS
527 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
528 with respect to surface coverage.
529
530 \begin{figure}
531 \includegraphics[width=\linewidth]{coverage}
532 \caption{The interfacial thermal conductivity ($G$) has a
533 non-monotonic dependence on the degree of surface capping. This
534 data is for the Au(111) / butanethiol / solvent interface with
535 various UA force fields at $\langle T\rangle \sim $200K.}
536 \label{coverage}
537 \end{figure}
538
539 In partially covered surfaces, the derivative definition for
540 $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
541 location of maximum change of $\lambda$ becomes washed out. The
542 discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
543 Gibbs dividing surface is still well-defined. Therefore, $G$ (not
544 $G^\prime$) was used in this section.
545
546 From Figure \ref{coverage}, one can see the significance of the
547 presence of capping agents. When even a small fraction of the Au(111)
548 surface sites are covered with butanethiols, the conductivity exhibits
549 an enhancement by at least a factor of 3. Capping agents are clearly
550 playing a major role in thermal transport at metal / organic solvent
551 surfaces.
552
553 We note a non-monotonic behavior in the interfacial conductance as a
554 function of surface coverage. The maximum conductance (largest $G$)
555 happens when the surfaces are about 75\% covered with butanethiol
556 caps. The reason for this behavior is not entirely clear. One
557 explanation is that incomplete butanethiol coverage allows small gaps
558 between butanethiols to form. These gaps can be filled by transient
559 solvent molecules. These solvent molecules couple very strongly with
560 the hot capping agent molecules near the surface, and can then carry
561 away (diffusively) the excess thermal energy from the surface.
562
563 There appears to be a competition between the conduction of the
564 thermal energy away from the surface by the capping agents (enhanced
565 by greater coverage) and the coupling of the capping agents with the
566 solvent (enhanced by interdigitation at lower coverages). This
567 competition would lead to the non-monotonic coverage behavior observed
568 here.
569
570 Results for rigid body toluene solvent, as well as the UA hexane, are
571 within the ranges expected from prior experimental
572 work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
573 that explicit hydrogen atoms might not be required for modeling
574 thermal transport in these systems. C-H vibrational modes do not see
575 significant excited state population at low temperatures, and are not
576 likely to carry lower frequency excitations from the solid layer into
577 the bulk liquid.
578
579 The toluene solvent does not exhibit the same behavior as hexane in
580 that $G$ remains at approximately the same magnitude when the capping
581 coverage increases from 25\% to 75\%. Toluene, as a rigid planar
582 molecule, cannot occupy the relatively small gaps between the capping
583 agents as easily as the chain-like {\it n}-hexane. The effect of
584 solvent coupling to the capping agent is therefore weaker in toluene
585 except at the very lowest coverage levels. This effect counters the
586 coverage-dependent conduction of heat away from the metal surface,
587 leading to a much flatter $G$ vs. coverage trend than is observed in
588 {\it n}-hexane.
589
590 \subsection{Effects due to Solvent \& Solvent Models}
591 In addition to UA solvent and capping agent models, AA models have
592 also been included in our simulations. In most of this work, the same
593 (UA or AA) model for solvent and capping agent was used, but it is
594 also possible to utilize different models for different components.
595 We have also included isotopic substitutions (Hydrogen to Deuterium)
596 to decrease the explicit vibrational overlap between solvent and
597 capping agent. Table \ref{modelTest} summarizes the results of these
598 studies.
599
600 {\bf MAY NOT NEED $J_z$ IN TABLE}
601 \begin{table*}
602 \begin{minipage}{\linewidth}
603 \begin{center}
604
605 \caption{Computed interfacial thermal conductance ($G$ and
606 $G^\prime$) values for interfaces using various models for
607 solvent and capping agent (or without capping agent) at
608 $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
609 solvent or capping agent molecules; ``Avg.'' denotes results
610 that are averages of simulations under different applied
611 thermal flux $(J_z)$ values. Error estimates are indicated in
612 parentheses.}
613
614 \begin{tabular}{llccc}
615 \hline\hline
616 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
617 (or bare surface) & model & (GW/m$^2$) &
618 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
619 \hline
620 UA & UA hexane & Avg. & 131(9) & 87(10) \\
621 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
622 & AA hexane & Avg. & 131(6) & 122(10) \\
623 & UA toluene & 1.96 & 187(16) & 151(11) \\
624 & AA toluene & 1.89 & 200(36) & 149(53) \\
625 \hline
626 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
627 & AA hexane & Avg. & 442(14) & 356(31) \\
628 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
629 & UA toluene & 1.98 & 125(25) & 97(60) \\
630 & AA toluene & 3.79 & 487(56) & 290(42) \\
631 \hline
632 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
633 & AA hexane & 1.92 & 243(29) & 191(11) \\
634 & AA toluene & 1.93 & 364(36) & 322(67) \\
635 \hline
636 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
637 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
638 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
639 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
640 \hline\hline
641 \end{tabular}
642 \label{modelTest}
643 \end{center}
644 \end{minipage}
645 \end{table*}
646
647 To facilitate direct comparison between force fields, systems with the
648 same capping agent and solvent were prepared with the same length
649 scales for the simulation cells.
650
651 On bare metal / solvent surfaces, different force field models for
652 hexane yield similar results for both $G$ and $G^\prime$, and these
653 two definitions agree with each other very well. This is primarily an
654 indicator of weak interactions between the metal and the solvent, and
655 is a typical case for acoustic impedance mismatch between these two
656 phases.
657
658 For the fully-covered surfaces, the choice of force field for the
659 capping agent and solvent has a large impact on the calculated values
660 of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
661 much larger than their UA to UA counterparts, and these values exceed
662 the experimental estimates by a large measure. The AA force field
663 allows significant energy to go into C-H (or C-D) stretching modes,
664 and since these modes are high frequency, this non-quantum behavior is
665 likely responsible for the overestimate of the conductivity. Compared
666 to the AA model, the UA model yields more reasonable conductivity
667 values with much higher computational efficiency.
668
669 \subsubsection{Are electronic excitations in the metal important?}
670 Because they lack electronic excitations, the QSC and related embedded
671 atom method (EAM) models for gold are known to predict unreasonably
672 low values for bulk conductivity
673 ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
674 conductance between the phases ($G$) is governed primarily by phonon
675 excitation (and not electronic degrees of freedom), one would expect a
676 classical model to capture most of the interfacial thermal
677 conductance. Our results for $G$ and $G^\prime$ indicate that this is
678 indeed the case, and suggest that the modeling of interfacial thermal
679 transport depends primarily on the description of the interactions
680 between the various components at the interface. When the metal is
681 chemically capped, the primary barrier to thermal conductivity appears
682 to be the interface between the capping agent and the surrounding
683 solvent, so the excitations in the metal have little impact on the
684 value of $G$.
685
686 \subsection{Effects due to methodology and simulation parameters}
687
688 We have varied the parameters of the simulations in order to
689 investigate how these factors would affect the computation of $G$. Of
690 particular interest are: 1) the length scale for the applied thermal
691 gradient (modified by increasing the amount of solvent in the system),
692 2) the sign and magnitude of the applied thermal flux, 3) the average
693 temperature of the simulation (which alters the solvent density during
694 equilibration), and 4) the definition of the interfacial conductance
695 (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
696 calculation.
697
698 Systems of different lengths were prepared by altering the number of
699 solvent molecules and extending the length of the box along the $z$
700 axis to accomodate the extra solvent. Equilibration at the same
701 temperature and pressure conditions led to nearly identical surface
702 areas ($L_x$ and $L_y$) available to the metal and capping agent,
703 while the extra solvent served mainly to lengthen the axis that was
704 used to apply the thermal flux. For a given value of the applied
705 flux, the different $z$ length scale has only a weak effect on the
706 computed conductivities (Table \ref{AuThiolHexaneUA}).
707
708 \subsubsection{Effects of applied flux}
709 The NIVS algorithm allows changes in both the sign and magnitude of
710 the applied flux. It is possible to reverse the direction of heat
711 flow simply by changing the sign of the flux, and thermal gradients
712 which would be difficult to obtain experimentally ($5$ K/\AA) can be
713 easily simulated. However, the magnitude of the applied flux is not
714 arbitrary if one aims to obtain a stable and reliable thermal gradient.
715 A temperature gradient can be lost in the noise if $|J_z|$ is too
716 small, and excessive $|J_z|$ values can cause phase transitions if the
717 extremes of the simulation cell become widely separated in
718 temperature. Also, if $|J_z|$ is too large for the bulk conductivity
719 of the materials, the thermal gradient will never reach a stable
720 state.
721
722 Within a reasonable range of $J_z$ values, we were able to study how
723 $G$ changes as a function of this flux. In what follows, we use
724 positive $J_z$ values to denote the case where energy is being
725 transferred by the method from the metal phase and into the liquid.
726 The resulting gradient therefore has a higher temperature in the
727 liquid phase. Negative flux values reverse this transfer, and result
728 in higher temperature metal phases. The conductance measured under
729 different applied $J_z$ values is listed in Tables
730 \ref{AuThiolHexaneUA} and \ref{AuThiolToluene}. These results do not
731 indicate that $G$ depends strongly on $J_z$ within this flux
732 range. The linear response of flux to thermal gradient simplifies our
733 investigations in that we can rely on $G$ measurement with only a
734 small number $J_z$ values.
735
736 {\bf MAY MOVE TO SUPPORT INFO}
737 \begin{table*}
738 \begin{minipage}{\linewidth}
739 \begin{center}
740 \caption{In the hexane-solvated interfaces, the system size has
741 little effect on the calculated values for interfacial
742 conductance ($G$ and $G^\prime$), but the direction of heat
743 flow (i.e. the sign of $J_z$) can alter the average
744 temperature of the liquid phase and this can alter the
745 computed conductivity.}
746
747 \begin{tabular}{ccccccc}
748 \hline\hline
749 $\langle T\rangle$ & $N_{hexane}$ & $\rho_{hexane}$ &
750 $J_z$ & $G$ & $G^\prime$ \\
751 (K) & & (g/cm$^3$) & (GW/m$^2$) &
752 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
753 \hline
754 200 & 266 & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
755 & 200 & 0.688 & 0.96 & 125(16) & 90.2(15) \\
756 & & & 1.91 & 139(10) & 101(10) \\
757 & & & 2.83 & 141(6) & 89.9(9.8) \\
758 & 166 & 0.681 & 0.97 & 141(30) & 78(22) \\
759 & & & 1.92 & 138(4) & 98.9(9.5) \\
760 \hline
761 250 & 200 & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
762 & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
763 & 166 & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
764 & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
765 & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
766 & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
767 \hline\hline
768 \end{tabular}
769 \label{AuThiolHexaneUA}
770 \end{center}
771 \end{minipage}
772 \end{table*}
773
774 The sign of $J_z$ is a different matter, however, as this can alter
775 the temperature on the two sides of the interface. The average
776 temperature values reported are for the entire system, and not for the
777 liquid phase, so at a given $\langle T \rangle$, the system with
778 positive $J_z$ has a warmer liquid phase. This means that if the
779 liquid carries thermal energy via diffusive transport, {\it positive}
780 $J_z$ values will result in increased molecular motion on the liquid
781 side of the interface, and this will increase the measured
782 conductivity.
783
784 \subsubsection{Effects due to average temperature}
785
786 We also studied the effect of average system temperature on the
787 interfacial conductance. The simulations are first equilibrated in
788 the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
789 predict a lower boiling point (and liquid state density) than
790 experiments. This lower-density liquid phase leads to reduced contact
791 between the hexane and butanethiol, and this accounts for our
792 observation of lower conductance at higher temperatures as shown in
793 Table \ref{AuThiolHexaneUA}. In raising the average temperature from
794 200K to 250K, the density drop of $\sim$20\% in the solvent phase
795 leads to a $\sim$40\% drop in the conductance.
796
797 Similar behavior is observed in the TraPPE-UA model for toluene,
798 although this model has better agreement with the experimental
799 densities of toluene. The expansion of the toluene liquid phase is
800 not as significant as that of the hexane (8.3\% over 100K), and this
801 limits the effect to $\sim$20\% drop in thermal conductivity (Table
802 \ref{AuThiolToluene}).
803
804 Although we have not mapped out the behavior at a large number of
805 temperatures, is clear that there will be a strong temperature
806 dependence in the interfacial conductance when the physical properties
807 of one side of the interface (notably the density) change rapidly as a
808 function of temperature.
809
810 {\bf MAY MOVE TO SUPPORT INFO}
811 \begin{table*}
812 \begin{minipage}{\linewidth}
813 \begin{center}
814 \caption{When toluene is the solvent, the interfacial thermal
815 conductivity is less sensitive to temperature, but again, the
816 direction of the heat flow can alter the solvent temperature
817 and can change the computed conductance values.}
818
819 \begin{tabular}{ccccc}
820 \hline\hline
821 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
822 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
823 \hline
824 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
825 & & -1.86 & 180(3) & 135(21) \\
826 & & -3.93 & 176(5) & 113(12) \\
827 \hline
828 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
829 & & -4.19 & 135(9) & 113(12) \\
830 \hline\hline
831 \end{tabular}
832 \label{AuThiolToluene}
833 \end{center}
834 \end{minipage}
835 \end{table*}
836
837 Besides the lower interfacial thermal conductance, surfaces at
838 relatively high temperatures are susceptible to reconstructions,
839 particularly when butanethiols fully cover the Au(111) surface. These
840 reconstructions include surface Au atoms which migrate outward to the
841 S atom layer, and butanethiol molecules which embed into the surface
842 Au layer. The driving force for this behavior is the strong Au-S
843 interactions which are modeled here with a deep Lennard-Jones
844 potential. This phenomenon agrees with reconstructions that have been
845 experimentally
846 observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
847 {\it et al.} kept their Au(111) slab rigid so that their simulations
848 could reach 300K without surface
849 reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
850 blur the interface, the measurement of $G$ becomes more difficult to
851 conduct at higher temperatures. For this reason, most of our
852 measurements are undertaken at $\langle T\rangle\sim$200K where
853 reconstruction is minimized.
854
855 However, when the surface is not completely covered by butanethiols,
856 the simulated system appears to be more resistent to the
857 reconstruction. Our Au / butanethiol / toluene system had the Au(111)
858 surfaces 90\% covered by butanethiols, but did not see this above
859 phenomena even at $\langle T\rangle\sim$300K. That said, we did
860 observe butanethiols migrating to neighboring three-fold sites during
861 a simulation. Since the interface persisted in these simulations, we
862 were able to obtain $G$'s for these interfaces even at a relatively
863 high temperature without being affected by surface reconstructions.
864
865 \section{Discussion}
866
867 The primary result of this work is that the capping agent acts as an
868 efficient thermal coupler between solid and solvent phases. One of
869 the ways the capping agent can carry out this role is to down-shift
870 between the phonon vibrations in the solid (which carry the heat from
871 the gold) and the molecular vibrations in the liquid (which carry some
872 of the heat in the solvent).
873
874 To investigate the mechanism of interfacial thermal conductance, the
875 vibrational power spectrum was computed. Power spectra were taken for
876 individual components in different simulations. To obtain these
877 spectra, simulations were run after equilibration in the
878 microcanonical (NVE) ensemble and without a thermal
879 gradient. Snapshots of configurations were collected at a frequency
880 that is higher than that of the fastest vibrations occurring in the
881 simulations. With these configurations, the velocity auto-correlation
882 functions can be computed:
883 \begin{equation}
884 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
885 \label{vCorr}
886 \end{equation}
887 The power spectrum is constructed via a Fourier transform of the
888 symmetrized velocity autocorrelation function,
889 \begin{equation}
890 \hat{f}(\omega) =
891 \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
892 \label{fourier}
893 \end{equation}
894
895 \subsection{The role of specific vibrations}
896 The vibrational spectra for gold slabs in different environments are
897 shown as in Figure \ref{specAu}. Regardless of the presence of
898 solvent, the gold surfaces which are covered by butanethiol molecules
899 exhibit an additional peak observed at a frequency of
900 $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
901 vibration. This vibration enables efficient thermal coupling of the
902 surface Au layer to the capping agents. Therefore, in our simulations,
903 the Au / S interfaces do not appear to be the primary barrier to
904 thermal transport when compared with the butanethiol / solvent
905 interfaces. {\bf This confirms the results from Luo {\it et
906 al.}\cite{Luo20101}, which reported $G$ for Au-SAM junctions
907 generally twice larger than what we have computed for the
908 thiol-liquid interfaces.}
909
910 \begin{figure}
911 \includegraphics[width=\linewidth]{vibration}
912 \caption{The vibrational power spectrum for thiol-capped gold has an
913 additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
914 surfaces (both with and without a solvent over-layer) are missing
915 this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
916 the vibrational power spectrum for the butanethiol capping agents.}
917 \label{specAu}
918 \end{figure}
919
920 Also in this figure, we show the vibrational power spectrum for the
921 bound butanethiol molecules, which also exhibits the same
922 $\sim$165cm$^{-1}$ peak.
923
924 \subsection{Overlap of power spectra}
925 A comparison of the results obtained from the two different organic
926 solvents can also provide useful information of the interfacial
927 thermal transport process. In particular, the vibrational overlap
928 between the butanethiol and the organic solvents suggests a highly
929 efficient thermal exchange between these components. Very high
930 thermal conductivity was observed when AA models were used and C-H
931 vibrations were treated classically. The presence of extra degrees of
932 freedom in the AA force field yields higher heat exchange rates
933 between the two phases and results in a much higher conductivity than
934 in the UA force field. {\bf Due to the classical models used, this
935 even includes those high frequency modes which should be unpopulated
936 at our relatively low temperatures. This artifact causes high
937 frequency vibrations accountable for thermal transport in classical
938 MD simulations.}
939
940 The similarity in the vibrational modes available to solvent and
941 capping agent can be reduced by deuterating one of the two components
942 (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
943 are deuterated, one can observe a significantly lower $G$ and
944 $G^\prime$ values (Table \ref{modelTest}).
945
946 \begin{figure}
947 \includegraphics[width=\linewidth]{aahxntln}
948 \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
949 systems. When butanethiol is deuterated (lower left), its
950 vibrational overlap with hexane decreases significantly. Since
951 aromatic molecules and the butanethiol are vibrationally dissimilar,
952 the change is not as dramatic when toluene is the solvent (right).}
953 \label{aahxntln}
954 \end{figure}
955
956 For the Au / butanethiol / toluene interfaces, having the AA
957 butanethiol deuterated did not yield a significant change in the
958 measured conductance. Compared to the C-H vibrational overlap between
959 hexane and butanethiol, both of which have alkyl chains, the overlap
960 between toluene and butanethiol is not as significant and thus does
961 not contribute as much to the heat exchange process.
962
963 Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
964 that the {\it intra}molecular heat transport due to alkylthiols is
965 highly efficient. Combining our observations with those of Zhang {\it
966 et al.}, it appears that butanethiol acts as a channel to expedite
967 heat flow from the gold surface and into the alkyl chain. The
968 acoustic impedance mismatch between the metal and the liquid phase can
969 therefore be effectively reduced with the presence of suitable capping
970 agents.
971
972 Deuterated models in the UA force field did not decouple the thermal
973 transport as well as in the AA force field. The UA models, even
974 though they have eliminated the high frequency C-H vibrational
975 overlap, still have significant overlap in the lower-frequency
976 portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
977 the UA models did not decouple the low frequency region enough to
978 produce an observable difference for the results of $G$ (Table
979 \ref{modelTest}).
980
981 \begin{figure}
982 \includegraphics[width=\linewidth]{uahxnua}
983 \caption{Vibrational power spectra for UA models for the butanethiol
984 and hexane solvent (upper panel) show the high degree of overlap
985 between these two molecules, particularly at lower frequencies.
986 Deuterating a UA model for the solvent (lower panel) does not
987 decouple the two spectra to the same degree as in the AA force
988 field (see Fig \ref{aahxntln}).}
989 \label{uahxnua}
990 \end{figure}
991
992 \section{Conclusions}
993 The NIVS algorithm has been applied to simulations of
994 butanethiol-capped Au(111) surfaces in the presence of organic
995 solvents. This algorithm allows the application of unphysical thermal
996 flux to transfer heat between the metal and the liquid phase. With the
997 flux applied, we were able to measure the corresponding thermal
998 gradients and to obtain interfacial thermal conductivities. Under
999 steady states, 2-3 ns trajectory simulations are sufficient for
1000 computation of this quantity.
1001
1002 Our simulations have seen significant conductance enhancement in the
1003 presence of capping agent, compared with the bare gold / liquid
1004 interfaces. The acoustic impedance mismatch between the metal and the
1005 liquid phase is effectively eliminated by a chemically-bonded capping
1006 agent. Furthermore, the coverage percentage of the capping agent plays
1007 an important role in the interfacial thermal transport
1008 process. Moderately low coverages allow higher contact between capping
1009 agent and solvent, and thus could further enhance the heat transfer
1010 process, giving a non-monotonic behavior of conductance with
1011 increasing coverage.
1012
1013 Our results, particularly using the UA models, agree well with
1014 available experimental data. The AA models tend to overestimate the
1015 interfacial thermal conductance in that the classically treated C-H
1016 vibrations become too easily populated. Compared to the AA models, the
1017 UA models have higher computational efficiency with satisfactory
1018 accuracy, and thus are preferable in modeling interfacial thermal
1019 transport.
1020
1021 Of the two definitions for $G$, the discrete form
1022 (Eq. \ref{discreteG}) was easier to use and gives out relatively
1023 consistent results, while the derivative form (Eq. \ref{derivativeG})
1024 is not as versatile. Although $G^\prime$ gives out comparable results
1025 and follows similar trend with $G$ when measuring close to fully
1026 covered or bare surfaces, the spatial resolution of $T$ profile
1027 required for the use of a derivative form is limited by the number of
1028 bins and the sampling required to obtain thermal gradient information.
1029
1030 Vlugt {\it et al.} have investigated the surface thiol structures for
1031 nanocrystalline gold and pointed out that they differ from those of
1032 the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
1033 difference could also cause differences in the interfacial thermal
1034 transport behavior. To investigate this problem, one would need an
1035 effective method for applying thermal gradients in non-planar
1036 (i.e. spherical) geometries.
1037
1038 \section{Acknowledgments}
1039 Support for this project was provided by the National Science
1040 Foundation under grant CHE-0848243. Computational time was provided by
1041 the Center for Research Computing (CRC) at the University of Notre
1042 Dame.
1043
1044 \section{Supporting Information}
1045 This information is available free of charge via the Internet at
1046 http://pubs.acs.org.
1047
1048 \newpage
1049
1050 \bibliography{interfacial}
1051
1052 \end{doublespace}
1053 \end{document}
1054