ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3765
Committed: Thu Sep 29 14:09:15 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 53015 byte(s)
Log Message:
add supporting information file, add keywords.

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
66 Keywords: non-equilibrium, molecular dynamics, vibrational overlap,
67 coverage dependent.
68 \end{abstract}
69
70 \newpage
71
72 %\narrowtext
73
74 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
75 % BODY OF TEXT
76 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77
78 \section{Introduction}
79 Due to the importance of heat flow (and heat removal) in
80 nanotechnology, interfacial thermal conductance has been studied
81 extensively both experimentally and computationally.\cite{cahill:793}
82 Nanoscale materials have a significant fraction of their atoms at
83 interfaces, and the chemical details of these interfaces govern the
84 thermal transport properties. Furthermore, the interfaces are often
85 heterogeneous (e.g. solid - liquid), which provides a challenge to
86 computational methods which have been developed for homogeneous or
87 bulk systems.
88
89 Experimentally, the thermal properties of a number of interfaces have
90 been investigated. Cahill and coworkers studied nanoscale thermal
91 transport from metal nanoparticle/fluid interfaces, to epitaxial
92 TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
93 interfaces between water and solids with different self-assembled
94 monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
95 Wang {\it et al.} studied heat transport through long-chain
96 hydrocarbon monolayers on gold substrate at individual molecular
97 level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
98 cetyltrimethylammonium bromide (CTAB) on the thermal transport between
99 gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
100 et al.} studied the cooling dynamics, which is controlled by thermal
101 interface resistance of glass-embedded metal
102 nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
103 normally considered barriers for heat transport, Alper {\it et al.}
104 suggested that specific ligands (capping agents) could completely
105 eliminate this barrier
106 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
107
108 Theoretical and computational models have also been used to study the
109 interfacial thermal transport in order to gain an understanding of
110 this phenomena at the molecular level. Recently, Hase and coworkers
111 employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
112 study thermal transport from hot Au(111) substrate to a self-assembled
113 monolayer of alkylthiol with relatively long chain (8-20 carbon
114 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
115 measurements for heat conductance of interfaces between the capping
116 monolayer on Au and a solvent phase have yet to be studied with their
117 approach. The comparatively low thermal flux through interfaces is
118 difficult to measure with Equilibrium
119 MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
120 methods. Therefore, the Reverse NEMD (RNEMD)
121 methods\cite{MullerPlathe:1997xw,kuang:164101} would be advantageous
122 in that they {\it apply} the difficult to measure quantity (flux),
123 while {\it measuring} the easily-computed quantity (the thermal
124 gradient). This is particularly true for inhomogeneous interfaces
125 where it would not be clear how to apply a gradient {\it a priori}.
126 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
127 this approach to various liquid interfaces and studied how thermal
128 conductance (or resistance) is dependent on chemical details of a
129 number of hydrophobic and hydrophilic aqueous interfaces. {\bf And
130 Luo {\it et al.} studied the thermal conductance of Au-SAM-Au
131 junctions using the same approach, with comparison to a constant
132 temperature difference method\cite{Luo20101}. While this latter
133 approach establishes more thermal distributions compared to the
134 former RNEMD methods, it does not guarantee momentum or kinetic
135 energy conservations.}
136
137 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
138 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
139 retains the desirable features of RNEMD (conservation of linear
140 momentum and total energy, compatibility with periodic boundary
141 conditions) while establishing true thermal distributions in each of
142 the two slabs. Furthermore, it allows effective thermal exchange
143 between particles of different identities, and thus makes the study of
144 interfacial conductance much simpler.
145
146 The work presented here deals with the Au(111) surface covered to
147 varying degrees by butanethiol, a capping agent with short carbon
148 chain, and solvated with organic solvents of different molecular
149 properties. {\bf To our knowledge, few previous MD inverstigations
150 have been found to address to these systems yet.} Different models
151 were used for both the capping agent and the solvent force field
152 parameters. Using the NIVS algorithm, the thermal transport across
153 these interfaces was studied and the underlying mechanism for the
154 phenomena was investigated.
155
156 \section{Methodology}
157 \subsection{Imposed-Flux Methods in MD Simulations}
158 Steady state MD simulations have an advantage in that not many
159 trajectories are needed to study the relationship between thermal flux
160 and thermal gradients. For systems with low interfacial conductance,
161 one must have a method capable of generating or measuring relatively
162 small fluxes, compared to those required for bulk conductivity. This
163 requirement makes the calculation even more difficult for
164 slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
165 NEMD methods impose a gradient (and measure a flux), but at interfaces
166 it is not clear what behavior should be imposed at the boundaries
167 between materials. Imposed-flux reverse non-equilibrium
168 methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
169 the thermal response becomes an easy-to-measure quantity. Although
170 M\"{u}ller-Plathe's original momentum swapping approach can be used
171 for exchanging energy between particles of different identity, the
172 kinetic energy transfer efficiency is affected by the mass difference
173 between the particles, which limits its application on heterogeneous
174 interfacial systems.
175
176 The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
177 to non-equilibrium MD simulations is able to impose a wide range of
178 kinetic energy fluxes without obvious perturbation to the velocity
179 distributions of the simulated systems. Furthermore, this approach has
180 the advantage in heterogeneous interfaces in that kinetic energy flux
181 can be applied between regions of particles of arbitrary identity, and
182 the flux will not be restricted by difference in particle mass.
183
184 The NIVS algorithm scales the velocity vectors in two separate regions
185 of a simulation system with respective diagonal scaling matrices. To
186 determine these scaling factors in the matrices, a set of equations
187 including linear momentum conservation and kinetic energy conservation
188 constraints and target energy flux satisfaction is solved. With the
189 scaling operation applied to the system in a set frequency, bulk
190 temperature gradients can be easily established, and these can be used
191 for computing thermal conductivities. The NIVS algorithm conserves
192 momenta and energy and does not depend on an external thermostat.
193
194 \subsection{Defining Interfacial Thermal Conductivity ($G$)}
195
196 For an interface with relatively low interfacial conductance, and a
197 thermal flux between two distinct bulk regions, the regions on either
198 side of the interface rapidly come to a state in which the two phases
199 have relatively homogeneous (but distinct) temperatures. The
200 interfacial thermal conductivity $G$ can therefore be approximated as:
201 \begin{equation}
202 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
203 \langle T_\mathrm{cold}\rangle \right)}
204 \label{lowG}
205 \end{equation}
206 where ${E_{total}}$ is the total imposed non-physical kinetic energy
207 transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
208 and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
209 temperature of the two separated phases. For an applied flux $J_z$
210 operating over a simulation time $t$ on a periodically-replicated slab
211 of dimensions $L_x \times L_y$, $E_{total} = J_z *(t)*(2 L_x L_y)$.
212
213 When the interfacial conductance is {\it not} small, there are two
214 ways to define $G$. One common way is to assume the temperature is
215 discrete on the two sides of the interface. $G$ can be calculated
216 using the applied thermal flux $J$ and the maximum temperature
217 difference measured along the thermal gradient max($\Delta T$), which
218 occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
219 known as the Kapitza conductance, which is the inverse of the Kapitza
220 resistance.
221 \begin{equation}
222 G=\frac{J}{\Delta T}
223 \label{discreteG}
224 \end{equation}
225
226 \begin{figure}
227 \includegraphics[width=\linewidth]{method}
228 \caption{Interfacial conductance can be calculated by applying an
229 (unphysical) kinetic energy flux between two slabs, one located
230 within the metal and another on the edge of the periodic box. The
231 system responds by forming a thermal gradient. In bulk liquids,
232 this gradient typically has a single slope, but in interfacial
233 systems, there are distinct thermal conductivity domains. The
234 interfacial conductance, $G$ is found by measuring the temperature
235 gap at the Gibbs dividing surface, or by using second derivatives of
236 the thermal profile.}
237 \label{demoPic}
238 \end{figure}
239
240 {\bf We attempt another approach by assuming that temperature is
241 continuous and differentiable throughout the space. Given that
242 $\lambda$ is also differentiable, $G$ can be defined as its
243 gradient. This quantity has the same unit as the commonly known $G$,
244 and the maximum of its magnitude denotes where thermal conductivity
245 has the largest change, i.e. the interface. And vector
246 $\nabla\lambda$ is normal to the interface. In a simplified
247 condition here, we have both $\vec{J}$ and the thermal gradient
248 paralell to the $z$ axis and yield the formula used in our
249 computations.}
250 (original text)
251 The other approach is to assume a continuous temperature profile along
252 the thermal gradient axis (e.g. $z$) and define $G$ at the point where
253 the magnitude of thermal conductivity ($\lambda$) change reaches its
254 maximum, given that $\lambda$ is well-defined throughout the space:
255 \begin{equation}
256 G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
257 = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
258 \left(\frac{\partial T}{\partial z}\right)\right)\Big|
259 = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
260 \Big/\left(\frac{\partial T}{\partial z}\right)^2
261 \label{derivativeG}
262 \end{equation}
263
264 With temperature profiles obtained from simulation, one is able to
265 approximate the first and second derivatives of $T$ with finite
266 difference methods and calculate $G^\prime$. In what follows, both
267 definitions have been used, and are compared in the results.
268
269 To investigate the interfacial conductivity at metal / solvent
270 interfaces, we have modeled a metal slab with its (111) surfaces
271 perpendicular to the $z$-axis of our simulation cells. The metal slab
272 has been prepared both with and without capping agents on the exposed
273 surface, and has been solvated with simple organic solvents, as
274 illustrated in Figure \ref{gradT}.
275
276 With the simulation cell described above, we are able to equilibrate
277 the system and impose an unphysical thermal flux between the liquid
278 and the metal phase using the NIVS algorithm. By periodically applying
279 the unphysical flux, we obtained a temperature profile and its spatial
280 derivatives. Figure \ref{gradT} shows how an applied thermal flux can
281 be used to obtain the 1st and 2nd derivatives of the temperature
282 profile.
283
284 \begin{figure}
285 \includegraphics[width=\linewidth]{gradT}
286 \caption{A sample of Au (111) / butanethiol / hexane interfacial
287 system with the temperature profile after a kinetic energy flux has
288 been imposed. Note that the largest temperature jump in the thermal
289 profile (corresponding to the lowest interfacial conductance) is at
290 the interface between the butanethiol molecules (blue) and the
291 solvent (grey). First and second derivatives of the temperature
292 profile are obtained using a finite difference approximation (lower
293 panel).}
294 \label{gradT}
295 \end{figure}
296
297 \section{Computational Details}
298 \subsection{Simulation Protocol}
299 The NIVS algorithm has been implemented in our MD simulation code,
300 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
301 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
302 under atmospheric pressure (1 atm) and 200K. After equilibration,
303 butanethiol capping agents were placed at three-fold hollow sites on
304 the Au(111) surfaces. These sites are either {\it fcc} or {\it
305 hcp} sites, although Hase {\it et al.} found that they are
306 equivalent in a heat transfer process,\cite{hase:2010} so we did not
307 distinguish between these sites in our study. The maximum butanethiol
308 capacity on Au surface is $1/3$ of the total number of surface Au
309 atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
310 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
311 series of lower coverages was also prepared by eliminating
312 butanethiols from the higher coverage surface in a regular manner. The
313 lower coverages were prepared in order to study the relation between
314 coverage and interfacial conductance.
315
316 The capping agent molecules were allowed to migrate during the
317 simulations. They distributed themselves uniformly and sampled a
318 number of three-fold sites throughout out study. Therefore, the
319 initial configuration does not noticeably affect the sampling of a
320 variety of configurations of the same coverage, and the final
321 conductance measurement would be an average effect of these
322 configurations explored in the simulations.
323
324 After the modified Au-butanethiol surface systems were equilibrated in
325 the canonical (NVT) ensemble, organic solvent molecules were packed in
326 the previously empty part of the simulation cells.\cite{packmol} Two
327 solvents were investigated, one which has little vibrational overlap
328 with the alkanethiol and which has a planar shape (toluene), and one
329 which has similar vibrational frequencies to the capping agent and
330 chain-like shape ({\it n}-hexane).
331
332 The simulation cells were not particularly extensive along the
333 $z$-axis, as a very long length scale for the thermal gradient may
334 cause excessively hot or cold temperatures in the middle of the
335 solvent region and lead to undesired phenomena such as solvent boiling
336 or freezing when a thermal flux is applied. Conversely, too few
337 solvent molecules would change the normal behavior of the liquid
338 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
339 these extreme cases did not happen to our simulations. The spacing
340 between periodic images of the gold interfaces is $45 \sim 75$\AA in
341 our simulations.
342
343 The initial configurations generated are further equilibrated with the
344 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
345 change. This is to ensure that the equilibration of liquid phase does
346 not affect the metal's crystalline structure. Comparisons were made
347 with simulations that allowed changes of $L_x$ and $L_y$ during NPT
348 equilibration. No substantial changes in the box geometry were noticed
349 in these simulations. After ensuring the liquid phase reaches
350 equilibrium at atmospheric pressure (1 atm), further equilibration was
351 carried out under canonical (NVT) and microcanonical (NVE) ensembles.
352
353 After the systems reach equilibrium, NIVS was used to impose an
354 unphysical thermal flux between the metal and the liquid phases. Most
355 of our simulations were done under an average temperature of
356 $\sim$200K. Therefore, thermal flux usually came from the metal to the
357 liquid so that the liquid has a higher temperature and would not
358 freeze due to lowered temperatures. After this induced temperature
359 gradient had stabilized, the temperature profile of the simulation cell
360 was recorded. To do this, the simulation cell is divided evenly into
361 $N$ slabs along the $z$-axis. The average temperatures of each slab
362 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
363 the same, the derivatives of $T$ with respect to slab number $n$ can
364 be directly used for $G^\prime$ calculations: \begin{equation}
365 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
366 \Big/\left(\frac{\partial T}{\partial z}\right)^2
367 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
368 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
369 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
370 \Big/\left(\frac{\partial T}{\partial n}\right)^2
371 \label{derivativeG2}
372 \end{equation}
373
374 All of the above simulation procedures use a time step of 1 fs. Each
375 equilibration stage took a minimum of 100 ps, although in some cases,
376 longer equilibration stages were utilized.
377
378 \subsection{Force Field Parameters}
379 Our simulations include a number of chemically distinct components.
380 Figure \ref{demoMol} demonstrates the sites defined for both
381 United-Atom and All-Atom models of the organic solvent and capping
382 agents in our simulations. Force field parameters are needed for
383 interactions both between the same type of particles and between
384 particles of different species.
385
386 \begin{figure}
387 \includegraphics[width=\linewidth]{structures}
388 \caption{Structures of the capping agent and solvents utilized in
389 these simulations. The chemically-distinct sites (a-e) are expanded
390 in terms of constituent atoms for both United Atom (UA) and All Atom
391 (AA) force fields. Most parameters are from References
392 \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
393 (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
394 atoms are given in Table \ref{MnM}.}
395 \label{demoMol}
396 \end{figure}
397
398 The Au-Au interactions in metal lattice slab is described by the
399 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
400 potentials include zero-point quantum corrections and are
401 reparametrized for accurate surface energies compared to the
402 Sutton-Chen potentials.\cite{Chen90}
403
404 For the two solvent molecules, {\it n}-hexane and toluene, two
405 different atomistic models were utilized. Both solvents were modeled
406 using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
407 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
408 for our UA solvent molecules. In these models, sites are located at
409 the carbon centers for alkyl groups. Bonding interactions, including
410 bond stretches and bends and torsions, were used for intra-molecular
411 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
412 potentials are used.
413
414 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
415 simple and computationally efficient, while maintaining good accuracy.
416 However, the TraPPE-UA model for alkanes is known to predict a slightly
417 lower boiling point than experimental values. This is one of the
418 reasons we used a lower average temperature (200K) for our
419 simulations. If heat is transferred to the liquid phase during the
420 NIVS simulation, the liquid in the hot slab can actually be
421 substantially warmer than the mean temperature in the simulation. The
422 lower mean temperatures therefore prevent solvent boiling.
423
424 For UA-toluene, the non-bonded potentials between intermolecular sites
425 have a similar Lennard-Jones formulation. The toluene molecules were
426 treated as a single rigid body, so there was no need for
427 intramolecular interactions (including bonds, bends, or torsions) in
428 this solvent model.
429
430 Besides the TraPPE-UA models, AA models for both organic solvents are
431 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
432 were used. For hexane, additional explicit hydrogen sites were
433 included. Besides bonding and non-bonded site-site interactions,
434 partial charges and the electrostatic interactions were added to each
435 CT and HC site. For toluene, a flexible model for the toluene molecule
436 was utilized which included bond, bend, torsion, and inversion
437 potentials to enforce ring planarity.
438
439 The butanethiol capping agent in our simulations, were also modeled
440 with both UA and AA model. The TraPPE-UA force field includes
441 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
442 UA butanethiol model in our simulations. The OPLS-AA also provides
443 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
444 surfaces do not have the hydrogen atom bonded to sulfur. To derive
445 suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
446 adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
447 modify the parameters for the CTS atom to maintain charge neutrality
448 in the molecule. Note that the model choice (UA or AA) for the capping
449 agent can be different from the solvent. Regardless of model choice,
450 the force field parameters for interactions between capping agent and
451 solvent can be derived using Lorentz-Berthelot Mixing Rule:
452 \begin{eqnarray}
453 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
454 \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
455 \end{eqnarray}
456
457 To describe the interactions between metal (Au) and non-metal atoms,
458 we refer to an adsorption study of alkyl thiols on gold surfaces by
459 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
460 Lennard-Jones form of potential parameters for the interaction between
461 Au and pseudo-atoms CH$_x$ and S based on a well-established and
462 widely-used effective potential of Hautman and Klein for the Au(111)
463 surface.\cite{hautman:4994} As our simulations require the gold slab
464 to be flexible to accommodate thermal excitation, the pair-wise form
465 of potentials they developed was used for our study.
466
467 The potentials developed from {\it ab initio} calculations by Leng
468 {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
469 interactions between Au and aromatic C/H atoms in toluene. However,
470 the Lennard-Jones parameters between Au and other types of particles,
471 (e.g. AA alkanes) have not yet been established. For these
472 interactions, the Lorentz-Berthelot mixing rule can be used to derive
473 effective single-atom LJ parameters for the metal using the fit values
474 for toluene. These are then used to construct reasonable mixing
475 parameters for the interactions between the gold and other atoms.
476 Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
477 our simulations.
478
479 \begin{table*}
480 \begin{minipage}{\linewidth}
481 \begin{center}
482 \caption{Non-bonded interaction parameters (including cross
483 interactions with Au atoms) for both force fields used in this
484 work.}
485 \begin{tabular}{lllllll}
486 \hline\hline
487 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
488 $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
489 & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
490 \hline
491 United Atom (UA)
492 &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
493 &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
494 &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
495 &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
496 \hline
497 All Atom (AA)
498 &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
499 &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
500 &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
501 &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
502 &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
503 &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
504 \hline
505 Both UA and AA
506 & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
507 \hline\hline
508 \end{tabular}
509 \label{MnM}
510 \end{center}
511 \end{minipage}
512 \end{table*}
513
514
515 \section{Results}
516 There are many factors contributing to the measured interfacial
517 conductance; some of these factors are physically motivated
518 (e.g. coverage of the surface by the capping agent coverage and
519 solvent identity), while some are governed by parameters of the
520 methodology (e.g. applied flux and the formulas used to obtain the
521 conductance). In this section we discuss the major physical and
522 calculational effects on the computed conductivity.
523
524 \subsection{Effects due to capping agent coverage}
525
526 A series of different initial conditions with a range of surface
527 coverages was prepared and solvated with various with both of the
528 solvent molecules. These systems were then equilibrated and their
529 interfacial thermal conductivity was measured with the NIVS
530 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
531 with respect to surface coverage.
532
533 \begin{figure}
534 \includegraphics[width=\linewidth]{coverage}
535 \caption{The interfacial thermal conductivity ($G$) has a
536 non-monotonic dependence on the degree of surface capping. This
537 data is for the Au(111) / butanethiol / solvent interface with
538 various UA force fields at $\langle T\rangle \sim $200K.}
539 \label{coverage}
540 \end{figure}
541
542 In partially covered surfaces, the derivative definition for
543 $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
544 location of maximum change of $\lambda$ becomes washed out. The
545 discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
546 Gibbs dividing surface is still well-defined. Therefore, $G$ (not
547 $G^\prime$) was used in this section.
548
549 From Figure \ref{coverage}, one can see the significance of the
550 presence of capping agents. When even a small fraction of the Au(111)
551 surface sites are covered with butanethiols, the conductivity exhibits
552 an enhancement by at least a factor of 3. Capping agents are clearly
553 playing a major role in thermal transport at metal / organic solvent
554 surfaces.
555
556 We note a non-monotonic behavior in the interfacial conductance as a
557 function of surface coverage. The maximum conductance (largest $G$)
558 happens when the surfaces are about 75\% covered with butanethiol
559 caps. The reason for this behavior is not entirely clear. One
560 explanation is that incomplete butanethiol coverage allows small gaps
561 between butanethiols to form. These gaps can be filled by transient
562 solvent molecules. These solvent molecules couple very strongly with
563 the hot capping agent molecules near the surface, and can then carry
564 away (diffusively) the excess thermal energy from the surface.
565
566 There appears to be a competition between the conduction of the
567 thermal energy away from the surface by the capping agents (enhanced
568 by greater coverage) and the coupling of the capping agents with the
569 solvent (enhanced by interdigitation at lower coverages). This
570 competition would lead to the non-monotonic coverage behavior observed
571 here.
572
573 Results for rigid body toluene solvent, as well as the UA hexane, are
574 within the ranges expected from prior experimental
575 work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
576 that explicit hydrogen atoms might not be required for modeling
577 thermal transport in these systems. C-H vibrational modes do not see
578 significant excited state population at low temperatures, and are not
579 likely to carry lower frequency excitations from the solid layer into
580 the bulk liquid.
581
582 The toluene solvent does not exhibit the same behavior as hexane in
583 that $G$ remains at approximately the same magnitude when the capping
584 coverage increases from 25\% to 75\%. Toluene, as a rigid planar
585 molecule, cannot occupy the relatively small gaps between the capping
586 agents as easily as the chain-like {\it n}-hexane. The effect of
587 solvent coupling to the capping agent is therefore weaker in toluene
588 except at the very lowest coverage levels. This effect counters the
589 coverage-dependent conduction of heat away from the metal surface,
590 leading to a much flatter $G$ vs. coverage trend than is observed in
591 {\it n}-hexane.
592
593 \subsection{Effects due to Solvent \& Solvent Models}
594 In addition to UA solvent and capping agent models, AA models have
595 also been included in our simulations. In most of this work, the same
596 (UA or AA) model for solvent and capping agent was used, but it is
597 also possible to utilize different models for different components.
598 We have also included isotopic substitutions (Hydrogen to Deuterium)
599 to decrease the explicit vibrational overlap between solvent and
600 capping agent. Table \ref{modelTest} summarizes the results of these
601 studies.
602
603 {\bf MAY NOT NEED $J_z$ IN TABLE}
604 \begin{table*}
605 \begin{minipage}{\linewidth}
606 \begin{center}
607
608 \caption{Computed interfacial thermal conductance ($G$ and
609 $G^\prime$) values for interfaces using various models for
610 solvent and capping agent (or without capping agent) at
611 $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
612 solvent or capping agent molecules; ``Avg.'' denotes results
613 that are averages of simulations under different applied
614 thermal flux $(J_z)$ values. Error estimates are indicated in
615 parentheses.}
616
617 \begin{tabular}{llccc}
618 \hline\hline
619 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
620 (or bare surface) & model & (GW/m$^2$) &
621 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
622 \hline
623 UA & UA hexane & Avg. & 131(9) & 87(10) \\
624 & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
625 & AA hexane & Avg. & 131(6) & 122(10) \\
626 & UA toluene & 1.96 & 187(16) & 151(11) \\
627 & AA toluene & 1.89 & 200(36) & 149(53) \\
628 \hline
629 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
630 & AA hexane & Avg. & 442(14) & 356(31) \\
631 & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
632 & UA toluene & 1.98 & 125(25) & 97(60) \\
633 & AA toluene & 3.79 & 487(56) & 290(42) \\
634 \hline
635 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
636 & AA hexane & 1.92 & 243(29) & 191(11) \\
637 & AA toluene & 1.93 & 364(36) & 322(67) \\
638 \hline
639 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
640 & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
641 & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
642 & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
643 \hline\hline
644 \end{tabular}
645 \label{modelTest}
646 \end{center}
647 \end{minipage}
648 \end{table*}
649
650 To facilitate direct comparison between force fields, systems with the
651 same capping agent and solvent were prepared with the same length
652 scales for the simulation cells.
653
654 On bare metal / solvent surfaces, different force field models for
655 hexane yield similar results for both $G$ and $G^\prime$, and these
656 two definitions agree with each other very well. This is primarily an
657 indicator of weak interactions between the metal and the solvent, and
658 is a typical case for acoustic impedance mismatch between these two
659 phases.
660
661 For the fully-covered surfaces, the choice of force field for the
662 capping agent and solvent has a large impact on the calculated values
663 of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
664 much larger than their UA to UA counterparts, and these values exceed
665 the experimental estimates by a large measure. The AA force field
666 allows significant energy to go into C-H (or C-D) stretching modes,
667 and since these modes are high frequency, this non-quantum behavior is
668 likely responsible for the overestimate of the conductivity. Compared
669 to the AA model, the UA model yields more reasonable conductivity
670 values with much higher computational efficiency.
671
672 \subsubsection{Are electronic excitations in the metal important?}
673 Because they lack electronic excitations, the QSC and related embedded
674 atom method (EAM) models for gold are known to predict unreasonably
675 low values for bulk conductivity
676 ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
677 conductance between the phases ($G$) is governed primarily by phonon
678 excitation (and not electronic degrees of freedom), one would expect a
679 classical model to capture most of the interfacial thermal
680 conductance. Our results for $G$ and $G^\prime$ indicate that this is
681 indeed the case, and suggest that the modeling of interfacial thermal
682 transport depends primarily on the description of the interactions
683 between the various components at the interface. When the metal is
684 chemically capped, the primary barrier to thermal conductivity appears
685 to be the interface between the capping agent and the surrounding
686 solvent, so the excitations in the metal have little impact on the
687 value of $G$.
688
689 \subsection{Effects due to methodology and simulation parameters}
690
691 We have varied the parameters of the simulations in order to
692 investigate how these factors would affect the computation of $G$. Of
693 particular interest are: 1) the length scale for the applied thermal
694 gradient (modified by increasing the amount of solvent in the system),
695 2) the sign and magnitude of the applied thermal flux, 3) the average
696 temperature of the simulation (which alters the solvent density during
697 equilibration), and 4) the definition of the interfacial conductance
698 (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
699 calculation.
700
701 Systems of different lengths were prepared by altering the number of
702 solvent molecules and extending the length of the box along the $z$
703 axis to accomodate the extra solvent. Equilibration at the same
704 temperature and pressure conditions led to nearly identical surface
705 areas ($L_x$ and $L_y$) available to the metal and capping agent,
706 while the extra solvent served mainly to lengthen the axis that was
707 used to apply the thermal flux. For a given value of the applied
708 flux, the different $z$ length scale has only a weak effect on the
709 computed conductivities (Table \ref{AuThiolHexaneUA}).
710
711 \subsubsection{Effects of applied flux}
712 The NIVS algorithm allows changes in both the sign and magnitude of
713 the applied flux. It is possible to reverse the direction of heat
714 flow simply by changing the sign of the flux, and thermal gradients
715 which would be difficult to obtain experimentally ($5$ K/\AA) can be
716 easily simulated. However, the magnitude of the applied flux is not
717 arbitrary if one aims to obtain a stable and reliable thermal gradient.
718 A temperature gradient can be lost in the noise if $|J_z|$ is too
719 small, and excessive $|J_z|$ values can cause phase transitions if the
720 extremes of the simulation cell become widely separated in
721 temperature. Also, if $|J_z|$ is too large for the bulk conductivity
722 of the materials, the thermal gradient will never reach a stable
723 state.
724
725 Within a reasonable range of $J_z$ values, we were able to study how
726 $G$ changes as a function of this flux. In what follows, we use
727 positive $J_z$ values to denote the case where energy is being
728 transferred by the method from the metal phase and into the liquid.
729 The resulting gradient therefore has a higher temperature in the
730 liquid phase. Negative flux values reverse this transfer, and result
731 in higher temperature metal phases. The conductance measured under
732 different applied $J_z$ values is listed in Tables
733 \ref{AuThiolHexaneUA} and \ref{AuThiolToluene}. These results do not
734 indicate that $G$ depends strongly on $J_z$ within this flux
735 range. The linear response of flux to thermal gradient simplifies our
736 investigations in that we can rely on $G$ measurement with only a
737 small number $J_z$ values.
738
739 {\bf MAY MOVE TO SUPPORT INFO}
740 \begin{table*}
741 \begin{minipage}{\linewidth}
742 \begin{center}
743 \caption{In the hexane-solvated interfaces, the system size has
744 little effect on the calculated values for interfacial
745 conductance ($G$ and $G^\prime$), but the direction of heat
746 flow (i.e. the sign of $J_z$) can alter the average
747 temperature of the liquid phase and this can alter the
748 computed conductivity.}
749
750 \begin{tabular}{ccccccc}
751 \hline\hline
752 $\langle T\rangle$ & $N_{hexane}$ & $\rho_{hexane}$ &
753 $J_z$ & $G$ & $G^\prime$ \\
754 (K) & & (g/cm$^3$) & (GW/m$^2$) &
755 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
756 \hline
757 200 & 266 & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
758 & 200 & 0.688 & 0.96 & 125(16) & 90.2(15) \\
759 & & & 1.91 & 139(10) & 101(10) \\
760 & & & 2.83 & 141(6) & 89.9(9.8) \\
761 & 166 & 0.681 & 0.97 & 141(30) & 78(22) \\
762 & & & 1.92 & 138(4) & 98.9(9.5) \\
763 \hline
764 250 & 200 & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
765 & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
766 & 166 & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
767 & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
768 & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
769 & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
770 \hline\hline
771 \end{tabular}
772 \label{AuThiolHexaneUA}
773 \end{center}
774 \end{minipage}
775 \end{table*}
776
777 The sign of $J_z$ is a different matter, however, as this can alter
778 the temperature on the two sides of the interface. The average
779 temperature values reported are for the entire system, and not for the
780 liquid phase, so at a given $\langle T \rangle$, the system with
781 positive $J_z$ has a warmer liquid phase. This means that if the
782 liquid carries thermal energy via diffusive transport, {\it positive}
783 $J_z$ values will result in increased molecular motion on the liquid
784 side of the interface, and this will increase the measured
785 conductivity.
786
787 \subsubsection{Effects due to average temperature}
788
789 We also studied the effect of average system temperature on the
790 interfacial conductance. The simulations are first equilibrated in
791 the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
792 predict a lower boiling point (and liquid state density) than
793 experiments. This lower-density liquid phase leads to reduced contact
794 between the hexane and butanethiol, and this accounts for our
795 observation of lower conductance at higher temperatures as shown in
796 Table \ref{AuThiolHexaneUA}. In raising the average temperature from
797 200K to 250K, the density drop of $\sim$20\% in the solvent phase
798 leads to a $\sim$40\% drop in the conductance.
799
800 Similar behavior is observed in the TraPPE-UA model for toluene,
801 although this model has better agreement with the experimental
802 densities of toluene. The expansion of the toluene liquid phase is
803 not as significant as that of the hexane (8.3\% over 100K), and this
804 limits the effect to $\sim$20\% drop in thermal conductivity (Table
805 \ref{AuThiolToluene}).
806
807 Although we have not mapped out the behavior at a large number of
808 temperatures, is clear that there will be a strong temperature
809 dependence in the interfacial conductance when the physical properties
810 of one side of the interface (notably the density) change rapidly as a
811 function of temperature.
812
813 {\bf MAY MOVE TO SUPPORT INFO}
814 \begin{table*}
815 \begin{minipage}{\linewidth}
816 \begin{center}
817 \caption{When toluene is the solvent, the interfacial thermal
818 conductivity is less sensitive to temperature, but again, the
819 direction of the heat flow can alter the solvent temperature
820 and can change the computed conductance values.}
821
822 \begin{tabular}{ccccc}
823 \hline\hline
824 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
825 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
826 \hline
827 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
828 & & -1.86 & 180(3) & 135(21) \\
829 & & -3.93 & 176(5) & 113(12) \\
830 \hline
831 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
832 & & -4.19 & 135(9) & 113(12) \\
833 \hline\hline
834 \end{tabular}
835 \label{AuThiolToluene}
836 \end{center}
837 \end{minipage}
838 \end{table*}
839
840 Besides the lower interfacial thermal conductance, surfaces at
841 relatively high temperatures are susceptible to reconstructions,
842 particularly when butanethiols fully cover the Au(111) surface. These
843 reconstructions include surface Au atoms which migrate outward to the
844 S atom layer, and butanethiol molecules which embed into the surface
845 Au layer. The driving force for this behavior is the strong Au-S
846 interactions which are modeled here with a deep Lennard-Jones
847 potential. This phenomenon agrees with reconstructions that have been
848 experimentally
849 observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
850 {\it et al.} kept their Au(111) slab rigid so that their simulations
851 could reach 300K without surface
852 reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
853 blur the interface, the measurement of $G$ becomes more difficult to
854 conduct at higher temperatures. For this reason, most of our
855 measurements are undertaken at $\langle T\rangle\sim$200K where
856 reconstruction is minimized.
857
858 However, when the surface is not completely covered by butanethiols,
859 the simulated system appears to be more resistent to the
860 reconstruction. Our Au / butanethiol / toluene system had the Au(111)
861 surfaces 90\% covered by butanethiols, but did not see this above
862 phenomena even at $\langle T\rangle\sim$300K. That said, we did
863 observe butanethiols migrating to neighboring three-fold sites during
864 a simulation. Since the interface persisted in these simulations, we
865 were able to obtain $G$'s for these interfaces even at a relatively
866 high temperature without being affected by surface reconstructions.
867
868 \section{Discussion}
869
870 The primary result of this work is that the capping agent acts as an
871 efficient thermal coupler between solid and solvent phases. One of
872 the ways the capping agent can carry out this role is to down-shift
873 between the phonon vibrations in the solid (which carry the heat from
874 the gold) and the molecular vibrations in the liquid (which carry some
875 of the heat in the solvent).
876
877 To investigate the mechanism of interfacial thermal conductance, the
878 vibrational power spectrum was computed. Power spectra were taken for
879 individual components in different simulations. To obtain these
880 spectra, simulations were run after equilibration in the
881 microcanonical (NVE) ensemble and without a thermal
882 gradient. Snapshots of configurations were collected at a frequency
883 that is higher than that of the fastest vibrations occurring in the
884 simulations. With these configurations, the velocity auto-correlation
885 functions can be computed:
886 \begin{equation}
887 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
888 \label{vCorr}
889 \end{equation}
890 The power spectrum is constructed via a Fourier transform of the
891 symmetrized velocity autocorrelation function,
892 \begin{equation}
893 \hat{f}(\omega) =
894 \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
895 \label{fourier}
896 \end{equation}
897
898 \subsection{The role of specific vibrations}
899 The vibrational spectra for gold slabs in different environments are
900 shown as in Figure \ref{specAu}. Regardless of the presence of
901 solvent, the gold surfaces which are covered by butanethiol molecules
902 exhibit an additional peak observed at a frequency of
903 $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
904 vibration. This vibration enables efficient thermal coupling of the
905 surface Au layer to the capping agents. Therefore, in our simulations,
906 the Au / S interfaces do not appear to be the primary barrier to
907 thermal transport when compared with the butanethiol / solvent
908 interfaces. {\bf This confirms the results from Luo {\it et
909 al.}\cite{Luo20101}, which reported $G$ for Au-SAM junctions
910 generally twice larger than what we have computed for the
911 thiol-liquid interfaces.}
912
913 \begin{figure}
914 \includegraphics[width=\linewidth]{vibration}
915 \caption{The vibrational power spectrum for thiol-capped gold has an
916 additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
917 surfaces (both with and without a solvent over-layer) are missing
918 this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
919 the vibrational power spectrum for the butanethiol capping agents.}
920 \label{specAu}
921 \end{figure}
922
923 Also in this figure, we show the vibrational power spectrum for the
924 bound butanethiol molecules, which also exhibits the same
925 $\sim$165cm$^{-1}$ peak.
926
927 \subsection{Overlap of power spectra}
928 A comparison of the results obtained from the two different organic
929 solvents can also provide useful information of the interfacial
930 thermal transport process. In particular, the vibrational overlap
931 between the butanethiol and the organic solvents suggests a highly
932 efficient thermal exchange between these components. Very high
933 thermal conductivity was observed when AA models were used and C-H
934 vibrations were treated classically. The presence of extra degrees of
935 freedom in the AA force field yields higher heat exchange rates
936 between the two phases and results in a much higher conductivity than
937 in the UA force field. {\bf Due to the classical models used, this
938 even includes those high frequency modes which should be unpopulated
939 at our relatively low temperatures. This artifact causes high
940 frequency vibrations accountable for thermal transport in classical
941 MD simulations.}
942
943 The similarity in the vibrational modes available to solvent and
944 capping agent can be reduced by deuterating one of the two components
945 (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
946 are deuterated, one can observe a significantly lower $G$ and
947 $G^\prime$ values (Table \ref{modelTest}).
948
949 \begin{figure}
950 \includegraphics[width=\linewidth]{aahxntln}
951 \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
952 systems. When butanethiol is deuterated (lower left), its
953 vibrational overlap with hexane decreases significantly. Since
954 aromatic molecules and the butanethiol are vibrationally dissimilar,
955 the change is not as dramatic when toluene is the solvent (right).}
956 \label{aahxntln}
957 \end{figure}
958
959 For the Au / butanethiol / toluene interfaces, having the AA
960 butanethiol deuterated did not yield a significant change in the
961 measured conductance. Compared to the C-H vibrational overlap between
962 hexane and butanethiol, both of which have alkyl chains, the overlap
963 between toluene and butanethiol is not as significant and thus does
964 not contribute as much to the heat exchange process.
965
966 Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
967 that the {\it intra}molecular heat transport due to alkylthiols is
968 highly efficient. Combining our observations with those of Zhang {\it
969 et al.}, it appears that butanethiol acts as a channel to expedite
970 heat flow from the gold surface and into the alkyl chain. The
971 acoustic impedance mismatch between the metal and the liquid phase can
972 therefore be effectively reduced with the presence of suitable capping
973 agents.
974
975 Deuterated models in the UA force field did not decouple the thermal
976 transport as well as in the AA force field. The UA models, even
977 though they have eliminated the high frequency C-H vibrational
978 overlap, still have significant overlap in the lower-frequency
979 portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
980 the UA models did not decouple the low frequency region enough to
981 produce an observable difference for the results of $G$ (Table
982 \ref{modelTest}).
983
984 \begin{figure}
985 \includegraphics[width=\linewidth]{uahxnua}
986 \caption{Vibrational power spectra for UA models for the butanethiol
987 and hexane solvent (upper panel) show the high degree of overlap
988 between these two molecules, particularly at lower frequencies.
989 Deuterating a UA model for the solvent (lower panel) does not
990 decouple the two spectra to the same degree as in the AA force
991 field (see Fig \ref{aahxntln}).}
992 \label{uahxnua}
993 \end{figure}
994
995 \section{Conclusions}
996 The NIVS algorithm has been applied to simulations of
997 butanethiol-capped Au(111) surfaces in the presence of organic
998 solvents. This algorithm allows the application of unphysical thermal
999 flux to transfer heat between the metal and the liquid phase. With the
1000 flux applied, we were able to measure the corresponding thermal
1001 gradients and to obtain interfacial thermal conductivities. Under
1002 steady states, 2-3 ns trajectory simulations are sufficient for
1003 computation of this quantity.
1004
1005 Our simulations have seen significant conductance enhancement in the
1006 presence of capping agent, compared with the bare gold / liquid
1007 interfaces. The acoustic impedance mismatch between the metal and the
1008 liquid phase is effectively eliminated by a chemically-bonded capping
1009 agent. Furthermore, the coverage percentage of the capping agent plays
1010 an important role in the interfacial thermal transport
1011 process. Moderately low coverages allow higher contact between capping
1012 agent and solvent, and thus could further enhance the heat transfer
1013 process, giving a non-monotonic behavior of conductance with
1014 increasing coverage.
1015
1016 Our results, particularly using the UA models, agree well with
1017 available experimental data. The AA models tend to overestimate the
1018 interfacial thermal conductance in that the classically treated C-H
1019 vibrations become too easily populated. Compared to the AA models, the
1020 UA models have higher computational efficiency with satisfactory
1021 accuracy, and thus are preferable in modeling interfacial thermal
1022 transport.
1023
1024 Of the two definitions for $G$, the discrete form
1025 (Eq. \ref{discreteG}) was easier to use and gives out relatively
1026 consistent results, while the derivative form (Eq. \ref{derivativeG})
1027 is not as versatile. Although $G^\prime$ gives out comparable results
1028 and follows similar trend with $G$ when measuring close to fully
1029 covered or bare surfaces, the spatial resolution of $T$ profile
1030 required for the use of a derivative form is limited by the number of
1031 bins and the sampling required to obtain thermal gradient information.
1032
1033 Vlugt {\it et al.} have investigated the surface thiol structures for
1034 nanocrystalline gold and pointed out that they differ from those of
1035 the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
1036 difference could also cause differences in the interfacial thermal
1037 transport behavior. To investigate this problem, one would need an
1038 effective method for applying thermal gradients in non-planar
1039 (i.e. spherical) geometries.
1040
1041 \section{Acknowledgments}
1042 Support for this project was provided by the National Science
1043 Foundation under grant CHE-0848243. Computational time was provided by
1044 the Center for Research Computing (CRC) at the University of Notre
1045 Dame.
1046
1047 \section{Supporting Information}
1048 This information is available free of charge via the Internet at
1049 http://pubs.acs.org.
1050
1051 \newpage
1052
1053 \bibliography{interfacial}
1054
1055 \end{doublespace}
1056 \end{document}
1057