ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3761
Committed: Fri Sep 23 17:31:50 2011 UTC (12 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 51149 byte(s)
Log Message:
publication version

File Contents

# User Rev Content
1 gezelter 3717 \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 skuang 3727 %\renewcommand\citemid{\ } % no comma in optional reference note
26 gezelter 3740 \bibpunct{[}{]}{,}{n}{}{;}
27     \bibliographystyle{achemso}
28 gezelter 3717
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 gezelter 3761 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 convective heat transport of solvent
64     molecules that have been in close contact with the capping agent.
65 gezelter 3717 \end{abstract}
66    
67     \newpage
68    
69     %\narrowtext
70    
71     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72     % BODY OF TEXT
73     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74    
75     \section{Introduction}
76 gezelter 3761 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 gezelter 3751 heterogeneous (e.g. solid - liquid), which provides a challenge to
83 gezelter 3761 computational methods which have been developed for homogeneous or
84     bulk systems.
85 gezelter 3717
86 gezelter 3761 Experimentally, the thermal properties of a number of interfaces have
87     been investigated. Cahill and coworkers studied nanoscale thermal
88 skuang 3755 transport from metal nanoparticle/fluid interfaces, to epitaxial
89 gezelter 3761 TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
90 skuang 3755 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 gezelter 3761 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 gezelter 3751 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 skuang 3733
105 skuang 3737 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 skuang 3738 monolayer of alkylthiol with relatively long chain (8-20 carbon
111 gezelter 3751 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
112 skuang 3737 measurements for heat conductance of interfaces between the capping
113 gezelter 3751 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 skuang 3755 difficult to measure with Equilibrium
116     MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
117 skuang 3750 methods. Therefore, the Reverse NEMD (RNEMD)
118 gezelter 3761 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 gezelter 3751 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
124     this approach to various liquid interfaces and studied how thermal
125 gezelter 3761 conductance (or resistance) is dependent on chemical details of a
126     number of hydrophobic and hydrophilic aqueous interfaces.
127 skuang 3734
128 gezelter 3751 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
129 skuang 3725 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
130     retains the desirable features of RNEMD (conservation of linear
131     momentum and total energy, compatibility with periodic boundary
132     conditions) while establishing true thermal distributions in each of
133 skuang 3737 the two slabs. Furthermore, it allows effective thermal exchange
134     between particles of different identities, and thus makes the study of
135     interfacial conductance much simpler.
136 skuang 3725
137 skuang 3737 The work presented here deals with the Au(111) surface covered to
138     varying degrees by butanethiol, a capping agent with short carbon
139     chain, and solvated with organic solvents of different molecular
140     properties. Different models were used for both the capping agent and
141     the solvent force field parameters. Using the NIVS algorithm, the
142     thermal transport across these interfaces was studied and the
143 skuang 3747 underlying mechanism for the phenomena was investigated.
144 skuang 3733
145 skuang 3721 \section{Methodology}
146 gezelter 3761 \subsection{Imposed-Flux Methods in MD Simulations}
147 gezelter 3751 Steady state MD simulations have an advantage in that not many
148 skuang 3749 trajectories are needed to study the relationship between thermal flux
149 gezelter 3751 and thermal gradients. For systems with low interfacial conductance,
150     one must have a method capable of generating or measuring relatively
151     small fluxes, compared to those required for bulk conductivity. This
152     requirement makes the calculation even more difficult for
153     slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
154     NEMD methods impose a gradient (and measure a flux), but at interfaces
155     it is not clear what behavior should be imposed at the boundaries
156     between materials. Imposed-flux reverse non-equilibrium
157     methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
158     the thermal response becomes an easy-to-measure quantity. Although
159 skuang 3749 M\"{u}ller-Plathe's original momentum swapping approach can be used
160     for exchanging energy between particles of different identity, the
161     kinetic energy transfer efficiency is affected by the mass difference
162     between the particles, which limits its application on heterogeneous
163     interfacial systems.
164 skuang 3721
165 gezelter 3751 The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
166     to non-equilibrium MD simulations is able to impose a wide range of
167 skuang 3737 kinetic energy fluxes without obvious perturbation to the velocity
168     distributions of the simulated systems. Furthermore, this approach has
169 skuang 3721 the advantage in heterogeneous interfaces in that kinetic energy flux
170 gezelter 3761 can be applied between regions of particles of arbitrary identity, and
171 skuang 3737 the flux will not be restricted by difference in particle mass.
172 skuang 3721
173     The NIVS algorithm scales the velocity vectors in two separate regions
174 gezelter 3761 of a simulation system with respective diagonal scaling matrices. To
175     determine these scaling factors in the matrices, a set of equations
176 skuang 3721 including linear momentum conservation and kinetic energy conservation
177 skuang 3737 constraints and target energy flux satisfaction is solved. With the
178     scaling operation applied to the system in a set frequency, bulk
179     temperature gradients can be easily established, and these can be used
180     for computing thermal conductivities. The NIVS algorithm conserves
181     momenta and energy and does not depend on an external thermostat.
182 skuang 3721
183 gezelter 3751 \subsection{Defining Interfacial Thermal Conductivity ($G$)}
184    
185     For an interface with relatively low interfacial conductance, and a
186     thermal flux between two distinct bulk regions, the regions on either
187     side of the interface rapidly come to a state in which the two phases
188     have relatively homogeneous (but distinct) temperatures. The
189     interfacial thermal conductivity $G$ can therefore be approximated as:
190 skuang 3727 \begin{equation}
191 gezelter 3751 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
192 skuang 3727 \langle T_\mathrm{cold}\rangle \right)}
193     \label{lowG}
194     \end{equation}
195 gezelter 3751 where ${E_{total}}$ is the total imposed non-physical kinetic energy
196     transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
197     and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
198 gezelter 3756 temperature of the two separated phases. For an applied flux $J_z$
199     operating over a simulation time $t$ on a periodically-replicated slab
200     of dimensions $L_x \times L_y$, $E_{total} = J_z *(t)*(2 L_x L_y)$.
201 skuang 3721
202 skuang 3737 When the interfacial conductance is {\it not} small, there are two
203 skuang 3752 ways to define $G$. One common way is to assume the temperature is
204     discrete on the two sides of the interface. $G$ can be calculated
205     using the applied thermal flux $J$ and the maximum temperature
206     difference measured along the thermal gradient max($\Delta T$), which
207 gezelter 3761 occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
208 skuang 3755 known as the Kapitza conductance, which is the inverse of the Kapitza
209     resistance.
210 skuang 3752 \begin{equation}
211     G=\frac{J}{\Delta T}
212     \label{discreteG}
213     \end{equation}
214 skuang 3727
215 skuang 3745 \begin{figure}
216     \includegraphics[width=\linewidth]{method}
217     \caption{Interfacial conductance can be calculated by applying an
218     (unphysical) kinetic energy flux between two slabs, one located
219     within the metal and another on the edge of the periodic box. The
220 gezelter 3761 system responds by forming a thermal gradient. In bulk liquids,
221     this gradient typically has a single slope, but in interfacial
222     systems, there are distinct thermal conductivity domains. The
223     interfacial conductance, $G$ is found by measuring the temperature
224     gap at the Gibbs dividing surface, or by using second derivatives of
225     the thermal profile.}
226 skuang 3745 \label{demoPic}
227     \end{figure}
228    
229 skuang 3727 The other approach is to assume a continuous temperature profile along
230     the thermal gradient axis (e.g. $z$) and define $G$ at the point where
231 gezelter 3751 the magnitude of thermal conductivity ($\lambda$) change reaches its
232 skuang 3727 maximum, given that $\lambda$ is well-defined throughout the space:
233     \begin{equation}
234     G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
235     = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
236     \left(\frac{\partial T}{\partial z}\right)\right)\Big|
237     = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
238     \Big/\left(\frac{\partial T}{\partial z}\right)^2
239     \label{derivativeG}
240     \end{equation}
241    
242 gezelter 3751 With temperature profiles obtained from simulation, one is able to
243 skuang 3727 approximate the first and second derivatives of $T$ with finite
244 gezelter 3751 difference methods and calculate $G^\prime$. In what follows, both
245     definitions have been used, and are compared in the results.
246 skuang 3727
247 gezelter 3751 To investigate the interfacial conductivity at metal / solvent
248     interfaces, we have modeled a metal slab with its (111) surfaces
249     perpendicular to the $z$-axis of our simulation cells. The metal slab
250     has been prepared both with and without capping agents on the exposed
251     surface, and has been solvated with simple organic solvents, as
252 skuang 3746 illustrated in Figure \ref{gradT}.
253 skuang 3727
254 skuang 3737 With the simulation cell described above, we are able to equilibrate
255     the system and impose an unphysical thermal flux between the liquid
256     and the metal phase using the NIVS algorithm. By periodically applying
257 gezelter 3751 the unphysical flux, we obtained a temperature profile and its spatial
258     derivatives. Figure \ref{gradT} shows how an applied thermal flux can
259     be used to obtain the 1st and 2nd derivatives of the temperature
260     profile.
261 skuang 3727
262     \begin{figure}
263     \includegraphics[width=\linewidth]{gradT}
264 gezelter 3761 \caption{A sample of Au (111) / butanethiol / hexane interfacial
265     system with the temperature profile after a kinetic energy flux has
266     been imposed. Note that the largest temperature jump in the thermal
267     profile (corresponding to the lowest interfacial conductance) is at
268     the interface between the butanethiol molecules (blue) and the
269     solvent (grey). First and second derivatives of the temperature
270     profile are obtained using a finite difference approximation (lower
271     panel).}
272 skuang 3727 \label{gradT}
273     \end{figure}
274    
275     \section{Computational Details}
276 skuang 3730 \subsection{Simulation Protocol}
277 skuang 3737 The NIVS algorithm has been implemented in our MD simulation code,
278 gezelter 3751 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
279     Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
280     under atmospheric pressure (1 atm) and 200K. After equilibration,
281     butanethiol capping agents were placed at three-fold hollow sites on
282     the Au(111) surfaces. These sites are either {\it fcc} or {\it
283     hcp} sites, although Hase {\it et al.} found that they are
284     equivalent in a heat transfer process,\cite{hase:2010} so we did not
285     distinguish between these sites in our study. The maximum butanethiol
286 skuang 3747 capacity on Au surface is $1/3$ of the total number of surface Au
287     atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
288 skuang 3749 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
289 gezelter 3751 series of lower coverages was also prepared by eliminating
290     butanethiols from the higher coverage surface in a regular manner. The
291     lower coverages were prepared in order to study the relation between
292     coverage and interfacial conductance.
293 skuang 3727
294 skuang 3737 The capping agent molecules were allowed to migrate during the
295     simulations. They distributed themselves uniformly and sampled a
296     number of three-fold sites throughout out study. Therefore, the
297 gezelter 3751 initial configuration does not noticeably affect the sampling of a
298 skuang 3737 variety of configurations of the same coverage, and the final
299     conductance measurement would be an average effect of these
300 gezelter 3751 configurations explored in the simulations.
301 skuang 3727
302 gezelter 3751 After the modified Au-butanethiol surface systems were equilibrated in
303     the canonical (NVT) ensemble, organic solvent molecules were packed in
304     the previously empty part of the simulation cells.\cite{packmol} Two
305 skuang 3737 solvents were investigated, one which has little vibrational overlap
306 gezelter 3751 with the alkanethiol and which has a planar shape (toluene), and one
307     which has similar vibrational frequencies to the capping agent and
308     chain-like shape ({\it n}-hexane).
309 skuang 3727
310 gezelter 3751 The simulation cells were not particularly extensive along the
311     $z$-axis, as a very long length scale for the thermal gradient may
312     cause excessively hot or cold temperatures in the middle of the
313 skuang 3730 solvent region and lead to undesired phenomena such as solvent boiling
314     or freezing when a thermal flux is applied. Conversely, too few
315     solvent molecules would change the normal behavior of the liquid
316     phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
317 gezelter 3751 these extreme cases did not happen to our simulations. The spacing
318 skuang 3760 between periodic images of the gold interfaces is $45 \sim 75$\AA in
319     our simulations.
320 skuang 3730
321 skuang 3746 The initial configurations generated are further equilibrated with the
322 gezelter 3751 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
323     change. This is to ensure that the equilibration of liquid phase does
324     not affect the metal's crystalline structure. Comparisons were made
325     with simulations that allowed changes of $L_x$ and $L_y$ during NPT
326     equilibration. No substantial changes in the box geometry were noticed
327     in these simulations. After ensuring the liquid phase reaches
328     equilibrium at atmospheric pressure (1 atm), further equilibration was
329     carried out under canonical (NVT) and microcanonical (NVE) ensembles.
330 skuang 3728
331 gezelter 3751 After the systems reach equilibrium, NIVS was used to impose an
332     unphysical thermal flux between the metal and the liquid phases. Most
333     of our simulations were done under an average temperature of
334     $\sim$200K. Therefore, thermal flux usually came from the metal to the
335 skuang 3727 liquid so that the liquid has a higher temperature and would not
336 gezelter 3751 freeze due to lowered temperatures. After this induced temperature
337 gezelter 3761 gradient had stabilized, the temperature profile of the simulation cell
338     was recorded. To do this, the simulation cell is divided evenly into
339 gezelter 3751 $N$ slabs along the $z$-axis. The average temperatures of each slab
340 skuang 3747 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
341     the same, the derivatives of $T$ with respect to slab number $n$ can
342 gezelter 3751 be directly used for $G^\prime$ calculations: \begin{equation}
343     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
344 skuang 3727 \Big/\left(\frac{\partial T}{\partial z}\right)^2
345     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
346     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
347     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
348     \Big/\left(\frac{\partial T}{\partial n}\right)^2
349     \label{derivativeG2}
350     \end{equation}
351    
352 gezelter 3751 All of the above simulation procedures use a time step of 1 fs. Each
353     equilibration stage took a minimum of 100 ps, although in some cases,
354     longer equilibration stages were utilized.
355 skuang 3747
356 skuang 3725 \subsection{Force Field Parameters}
357 gezelter 3751 Our simulations include a number of chemically distinct components.
358     Figure \ref{demoMol} demonstrates the sites defined for both
359     United-Atom and All-Atom models of the organic solvent and capping
360     agents in our simulations. Force field parameters are needed for
361 skuang 3744 interactions both between the same type of particles and between
362     particles of different species.
363 skuang 3721
364 skuang 3736 \begin{figure}
365 gezelter 3740 \includegraphics[width=\linewidth]{structures}
366     \caption{Structures of the capping agent and solvents utilized in
367     these simulations. The chemically-distinct sites (a-e) are expanded
368     in terms of constituent atoms for both United Atom (UA) and All Atom
369 gezelter 3761 (AA) force fields. Most parameters are from References
370     \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
371 skuang 3755 (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
372     atoms are given in Table \ref{MnM}.}
373 skuang 3736 \label{demoMol}
374     \end{figure}
375    
376 skuang 3744 The Au-Au interactions in metal lattice slab is described by the
377 gezelter 3751 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
378 skuang 3744 potentials include zero-point quantum corrections and are
379     reparametrized for accurate surface energies compared to the
380 gezelter 3751 Sutton-Chen potentials.\cite{Chen90}
381 skuang 3744
382 gezelter 3751 For the two solvent molecules, {\it n}-hexane and toluene, two
383     different atomistic models were utilized. Both solvents were modeled
384     using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
385 skuang 3728 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
386 skuang 3744 for our UA solvent molecules. In these models, sites are located at
387     the carbon centers for alkyl groups. Bonding interactions, including
388     bond stretches and bends and torsions, were used for intra-molecular
389 gezelter 3751 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
390     potentials are used.
391 skuang 3721
392 gezelter 3751 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
393     simple and computationally efficient, while maintaining good accuracy.
394 gezelter 3761 However, the TraPPE-UA model for alkanes is known to predict a slightly
395 gezelter 3751 lower boiling point than experimental values. This is one of the
396     reasons we used a lower average temperature (200K) for our
397     simulations. If heat is transferred to the liquid phase during the
398     NIVS simulation, the liquid in the hot slab can actually be
399     substantially warmer than the mean temperature in the simulation. The
400     lower mean temperatures therefore prevent solvent boiling.
401 skuang 3744
402 gezelter 3751 For UA-toluene, the non-bonded potentials between intermolecular sites
403     have a similar Lennard-Jones formulation. The toluene molecules were
404     treated as a single rigid body, so there was no need for
405     intramolecular interactions (including bonds, bends, or torsions) in
406     this solvent model.
407 skuang 3744
408 skuang 3729 Besides the TraPPE-UA models, AA models for both organic solvents are
409 skuang 3752 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
410     were used. For hexane, additional explicit hydrogen sites were
411 skuang 3744 included. Besides bonding and non-bonded site-site interactions,
412     partial charges and the electrostatic interactions were added to each
413 skuang 3752 CT and HC site. For toluene, a flexible model for the toluene molecule
414     was utilized which included bond, bend, torsion, and inversion
415     potentials to enforce ring planarity.
416 skuang 3728
417 gezelter 3751 The butanethiol capping agent in our simulations, were also modeled
418     with both UA and AA model. The TraPPE-UA force field includes
419 skuang 3730 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
420     UA butanethiol model in our simulations. The OPLS-AA also provides
421     parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
422 gezelter 3751 surfaces do not have the hydrogen atom bonded to sulfur. To derive
423     suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
424     adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
425     modify the parameters for the CTS atom to maintain charge neutrality
426     in the molecule. Note that the model choice (UA or AA) for the capping
427     agent can be different from the solvent. Regardless of model choice,
428     the force field parameters for interactions between capping agent and
429     solvent can be derived using Lorentz-Berthelot Mixing Rule:
430 skuang 3738 \begin{eqnarray}
431 gezelter 3751 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
432     \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
433 skuang 3738 \end{eqnarray}
434 skuang 3721
435 gezelter 3751 To describe the interactions between metal (Au) and non-metal atoms,
436     we refer to an adsorption study of alkyl thiols on gold surfaces by
437     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
438     Lennard-Jones form of potential parameters for the interaction between
439     Au and pseudo-atoms CH$_x$ and S based on a well-established and
440     widely-used effective potential of Hautman and Klein for the Au(111)
441     surface.\cite{hautman:4994} As our simulations require the gold slab
442     to be flexible to accommodate thermal excitation, the pair-wise form
443     of potentials they developed was used for our study.
444 skuang 3721
445 gezelter 3751 The potentials developed from {\it ab initio} calculations by Leng
446     {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
447     interactions between Au and aromatic C/H atoms in toluene. However,
448     the Lennard-Jones parameters between Au and other types of particles,
449     (e.g. AA alkanes) have not yet been established. For these
450     interactions, the Lorentz-Berthelot mixing rule can be used to derive
451     effective single-atom LJ parameters for the metal using the fit values
452     for toluene. These are then used to construct reasonable mixing
453     parameters for the interactions between the gold and other atoms.
454     Table \ref{MnM} summarizes the ``metal/non-metal'' parameters used in
455     our simulations.
456 skuang 3725
457 skuang 3730 \begin{table*}
458     \begin{minipage}{\linewidth}
459     \begin{center}
460 gezelter 3741 \caption{Non-bonded interaction parameters (including cross
461     interactions with Au atoms) for both force fields used in this
462     work.}
463     \begin{tabular}{lllllll}
464 skuang 3730 \hline\hline
465 gezelter 3741 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
466     $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
467     & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
468 skuang 3730 \hline
469 gezelter 3741 United Atom (UA)
470     &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
471     &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
472     &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
473     &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
474     \hline
475     All Atom (AA)
476     &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
477     &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
478     &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
479     &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
480     &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
481     &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
482     \hline
483 skuang 3744 Both UA and AA
484     & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
485 skuang 3730 \hline\hline
486     \end{tabular}
487     \label{MnM}
488     \end{center}
489     \end{minipage}
490     \end{table*}
491 skuang 3729
492 gezelter 3751
493 gezelter 3754 \section{Results}
494     There are many factors contributing to the measured interfacial
495     conductance; some of these factors are physically motivated
496     (e.g. coverage of the surface by the capping agent coverage and
497     solvent identity), while some are governed by parameters of the
498     methodology (e.g. applied flux and the formulas used to obtain the
499     conductance). In this section we discuss the major physical and
500     calculational effects on the computed conductivity.
501 skuang 3746
502 gezelter 3754 \subsection{Effects due to capping agent coverage}
503 skuang 3747
504 gezelter 3754 A series of different initial conditions with a range of surface
505     coverages was prepared and solvated with various with both of the
506     solvent molecules. These systems were then equilibrated and their
507 skuang 3755 interfacial thermal conductivity was measured with the NIVS
508 gezelter 3754 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
509     with respect to surface coverage.
510    
511     \begin{figure}
512     \includegraphics[width=\linewidth]{coverage}
513 gezelter 3761 \caption{The interfacial thermal conductivity ($G$) has a
514     non-monotonic dependence on the degree of surface capping. This
515     data is for the Au(111) / butanethiol / solvent interface with
516     various UA force fields at $\langle T\rangle \sim $200K.}
517 gezelter 3754 \label{coverage}
518     \end{figure}
519    
520 gezelter 3756 In partially covered surfaces, the derivative definition for
521     $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
522     location of maximum change of $\lambda$ becomes washed out. The
523     discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
524     Gibbs dividing surface is still well-defined. Therefore, $G$ (not
525     $G^\prime$) was used in this section.
526 gezelter 3754
527 gezelter 3756 From Figure \ref{coverage}, one can see the significance of the
528     presence of capping agents. When even a small fraction of the Au(111)
529     surface sites are covered with butanethiols, the conductivity exhibits
530 gezelter 3761 an enhancement by at least a factor of 3. Capping agents are clearly
531 gezelter 3756 playing a major role in thermal transport at metal / organic solvent
532     surfaces.
533 gezelter 3754
534 gezelter 3756 We note a non-monotonic behavior in the interfacial conductance as a
535     function of surface coverage. The maximum conductance (largest $G$)
536     happens when the surfaces are about 75\% covered with butanethiol
537     caps. The reason for this behavior is not entirely clear. One
538     explanation is that incomplete butanethiol coverage allows small gaps
539     between butanethiols to form. These gaps can be filled by transient
540     solvent molecules. These solvent molecules couple very strongly with
541     the hot capping agent molecules near the surface, and can then carry
542     away (diffusively) the excess thermal energy from the surface.
543 gezelter 3754
544 gezelter 3756 There appears to be a competition between the conduction of the
545     thermal energy away from the surface by the capping agents (enhanced
546     by greater coverage) and the coupling of the capping agents with the
547     solvent (enhanced by interdigitation at lower coverages). This
548     competition would lead to the non-monotonic coverage behavior observed
549     here.
550 gezelter 3754
551 gezelter 3756 Results for rigid body toluene solvent, as well as the UA hexane, are
552     within the ranges expected from prior experimental
553     work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
554     that explicit hydrogen atoms might not be required for modeling
555     thermal transport in these systems. C-H vibrational modes do not see
556     significant excited state population at low temperatures, and are not
557     likely to carry lower frequency excitations from the solid layer into
558     the bulk liquid.
559 gezelter 3754
560 gezelter 3756 The toluene solvent does not exhibit the same behavior as hexane in
561     that $G$ remains at approximately the same magnitude when the capping
562     coverage increases from 25\% to 75\%. Toluene, as a rigid planar
563     molecule, cannot occupy the relatively small gaps between the capping
564     agents as easily as the chain-like {\it n}-hexane. The effect of
565     solvent coupling to the capping agent is therefore weaker in toluene
566     except at the very lowest coverage levels. This effect counters the
567     coverage-dependent conduction of heat away from the metal surface,
568     leading to a much flatter $G$ vs. coverage trend than is observed in
569     {\it n}-hexane.
570 gezelter 3754
571     \subsection{Effects due to Solvent \& Solvent Models}
572 gezelter 3756 In addition to UA solvent and capping agent models, AA models have
573     also been included in our simulations. In most of this work, the same
574     (UA or AA) model for solvent and capping agent was used, but it is
575     also possible to utilize different models for different components.
576     We have also included isotopic substitutions (Hydrogen to Deuterium)
577     to decrease the explicit vibrational overlap between solvent and
578     capping agent. Table \ref{modelTest} summarizes the results of these
579     studies.
580 gezelter 3754
581     \begin{table*}
582     \begin{minipage}{\linewidth}
583     \begin{center}
584    
585 skuang 3755 \caption{Computed interfacial thermal conductance ($G$ and
586 gezelter 3754 $G^\prime$) values for interfaces using various models for
587     solvent and capping agent (or without capping agent) at
588 gezelter 3761 $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
589     solvent or capping agent molecules; ``Avg.'' denotes results
590     that are averages of simulations under different applied
591     thermal flux $(J_z)$ values. Error estimates are indicated in
592     parentheses.}
593 gezelter 3754
594     \begin{tabular}{llccc}
595     \hline\hline
596     Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
597     (or bare surface) & model & (GW/m$^2$) &
598     \multicolumn{2}{c}{(MW/m$^2$/K)} \\
599     \hline
600     UA & UA hexane & Avg. & 131(9) & 87(10) \\
601     & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
602     & AA hexane & Avg. & 131(6) & 122(10) \\
603     & UA toluene & 1.96 & 187(16) & 151(11) \\
604     & AA toluene & 1.89 & 200(36) & 149(53) \\
605     \hline
606     AA & UA hexane & 1.94 & 116(9) & 129(8) \\
607     & AA hexane & Avg. & 442(14) & 356(31) \\
608     & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
609     & UA toluene & 1.98 & 125(25) & 97(60) \\
610     & AA toluene & 3.79 & 487(56) & 290(42) \\
611     \hline
612     AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
613     & AA hexane & 1.92 & 243(29) & 191(11) \\
614     & AA toluene & 1.93 & 364(36) & 322(67) \\
615     \hline
616     bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
617     & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
618     & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
619     & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
620     \hline\hline
621     \end{tabular}
622     \label{modelTest}
623     \end{center}
624     \end{minipage}
625     \end{table*}
626    
627 gezelter 3756 To facilitate direct comparison between force fields, systems with the
628     same capping agent and solvent were prepared with the same length
629     scales for the simulation cells.
630 gezelter 3754
631 gezelter 3756 On bare metal / solvent surfaces, different force field models for
632     hexane yield similar results for both $G$ and $G^\prime$, and these
633     two definitions agree with each other very well. This is primarily an
634     indicator of weak interactions between the metal and the solvent, and
635     is a typical case for acoustic impedance mismatch between these two
636     phases.
637 gezelter 3754
638 gezelter 3756 For the fully-covered surfaces, the choice of force field for the
639 gezelter 3761 capping agent and solvent has a large impact on the calculated values
640 gezelter 3756 of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
641     much larger than their UA to UA counterparts, and these values exceed
642     the experimental estimates by a large measure. The AA force field
643     allows significant energy to go into C-H (or C-D) stretching modes,
644     and since these modes are high frequency, this non-quantum behavior is
645     likely responsible for the overestimate of the conductivity. Compared
646     to the AA model, the UA model yields more reasonable conductivity
647     values with much higher computational efficiency.
648 skuang 3755
649     \subsubsection{Are electronic excitations in the metal important?}
650 gezelter 3756 Because they lack electronic excitations, the QSC and related embedded
651     atom method (EAM) models for gold are known to predict unreasonably
652     low values for bulk conductivity
653     ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
654     conductance between the phases ($G$) is governed primarily by phonon
655     excitation (and not electronic degrees of freedom), one would expect a
656     classical model to capture most of the interfacial thermal
657     conductance. Our results for $G$ and $G^\prime$ indicate that this is
658     indeed the case, and suggest that the modeling of interfacial thermal
659     transport depends primarily on the description of the interactions
660     between the various components at the interface. When the metal is
661     chemically capped, the primary barrier to thermal conductivity appears
662     to be the interface between the capping agent and the surrounding
663     solvent, so the excitations in the metal have little impact on the
664     value of $G$.
665 gezelter 3754
666     \subsection{Effects due to methodology and simulation parameters}
667    
668 gezelter 3756 We have varied the parameters of the simulations in order to
669     investigate how these factors would affect the computation of $G$. Of
670     particular interest are: 1) the length scale for the applied thermal
671     gradient (modified by increasing the amount of solvent in the system),
672     2) the sign and magnitude of the applied thermal flux, 3) the average
673     temperature of the simulation (which alters the solvent density during
674     equilibration), and 4) the definition of the interfacial conductance
675     (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
676     calculation.
677 skuang 3725
678 gezelter 3756 Systems of different lengths were prepared by altering the number of
679     solvent molecules and extending the length of the box along the $z$
680     axis to accomodate the extra solvent. Equilibration at the same
681     temperature and pressure conditions led to nearly identical surface
682     areas ($L_x$ and $L_y$) available to the metal and capping agent,
683     while the extra solvent served mainly to lengthen the axis that was
684     used to apply the thermal flux. For a given value of the applied
685     flux, the different $z$ length scale has only a weak effect on the
686     computed conductivities (Table \ref{AuThiolHexaneUA}).
687 skuang 3725
688 gezelter 3756 \subsubsection{Effects of applied flux}
689     The NIVS algorithm allows changes in both the sign and magnitude of
690     the applied flux. It is possible to reverse the direction of heat
691     flow simply by changing the sign of the flux, and thermal gradients
692     which would be difficult to obtain experimentally ($5$ K/\AA) can be
693     easily simulated. However, the magnitude of the applied flux is not
694 gezelter 3761 arbitrary if one aims to obtain a stable and reliable thermal gradient.
695 gezelter 3756 A temperature gradient can be lost in the noise if $|J_z|$ is too
696     small, and excessive $|J_z|$ values can cause phase transitions if the
697     extremes of the simulation cell become widely separated in
698     temperature. Also, if $|J_z|$ is too large for the bulk conductivity
699     of the materials, the thermal gradient will never reach a stable
700     state.
701 skuang 3755
702 gezelter 3756 Within a reasonable range of $J_z$ values, we were able to study how
703     $G$ changes as a function of this flux. In what follows, we use
704     positive $J_z$ values to denote the case where energy is being
705     transferred by the method from the metal phase and into the liquid.
706     The resulting gradient therefore has a higher temperature in the
707     liquid phase. Negative flux values reverse this transfer, and result
708     in higher temperature metal phases. The conductance measured under
709     different applied $J_z$ values is listed in Tables
710     \ref{AuThiolHexaneUA} and \ref{AuThiolToluene}. These results do not
711     indicate that $G$ depends strongly on $J_z$ within this flux
712     range. The linear response of flux to thermal gradient simplifies our
713     investigations in that we can rely on $G$ measurement with only a
714     small number $J_z$ values.
715 skuang 3730
716 skuang 3725 \begin{table*}
717     \begin{minipage}{\linewidth}
718     \begin{center}
719 gezelter 3761 \caption{In the hexane-solvated interfaces, the system size has
720     little effect on the calculated values for interfacial
721     conductance ($G$ and $G^\prime$), but the direction of heat
722     flow (i.e. the sign of $J_z$) can alter the average
723     temperature of the liquid phase and this can alter the
724     computed conductivity.}
725 skuang 3730
726 skuang 3738 \begin{tabular}{ccccccc}
727 skuang 3730 \hline\hline
728 gezelter 3756 $\langle T\rangle$ & $N_{hexane}$ & $\rho_{hexane}$ &
729 skuang 3738 $J_z$ & $G$ & $G^\prime$ \\
730 gezelter 3756 (K) & & (g/cm$^3$) & (GW/m$^2$) &
731 skuang 3730 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
732     \hline
733 gezelter 3756 200 & 266 & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
734     & 200 & 0.688 & 0.96 & 125(16) & 90.2(15) \\
735     & & & 1.91 & 139(10) & 101(10) \\
736     & & & 2.83 & 141(6) & 89.9(9.8) \\
737     & 166 & 0.681 & 0.97 & 141(30) & 78(22) \\
738     & & & 1.92 & 138(4) & 98.9(9.5) \\
739 skuang 3739 \hline
740 gezelter 3756 250 & 200 & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
741     & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
742     & 166 & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
743     & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
744     & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
745     & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
746 skuang 3730 \hline\hline
747     \end{tabular}
748     \label{AuThiolHexaneUA}
749     \end{center}
750     \end{minipage}
751     \end{table*}
752    
753 gezelter 3756 The sign of $J_z$ is a different matter, however, as this can alter
754     the temperature on the two sides of the interface. The average
755     temperature values reported are for the entire system, and not for the
756     liquid phase, so at a given $\langle T \rangle$, the system with
757     positive $J_z$ has a warmer liquid phase. This means that if the
758     liquid carries thermal energy via convective transport, {\it positive}
759     $J_z$ values will result in increased molecular motion on the liquid
760     side of the interface, and this will increase the measured
761     conductivity.
762    
763 gezelter 3754 \subsubsection{Effects due to average temperature}
764    
765 gezelter 3756 We also studied the effect of average system temperature on the
766     interfacial conductance. The simulations are first equilibrated in
767     the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
768     predict a lower boiling point (and liquid state density) than
769     experiments. This lower-density liquid phase leads to reduced contact
770     between the hexane and butanethiol, and this accounts for our
771     observation of lower conductance at higher temperatures as shown in
772     Table \ref{AuThiolHexaneUA}. In raising the average temperature from
773 skuang 3760 200K to 250K, the density drop of $\sim$20\% in the solvent phase
774 gezelter 3761 leads to a $\sim$40\% drop in the conductance.
775 skuang 3730
776 gezelter 3756 Similar behavior is observed in the TraPPE-UA model for toluene,
777     although this model has better agreement with the experimental
778     densities of toluene. The expansion of the toluene liquid phase is
779     not as significant as that of the hexane (8.3\% over 100K), and this
780 skuang 3760 limits the effect to $\sim$20\% drop in thermal conductivity (Table
781 gezelter 3756 \ref{AuThiolToluene}).
782 skuang 3730
783 gezelter 3756 Although we have not mapped out the behavior at a large number of
784     temperatures, is clear that there will be a strong temperature
785     dependence in the interfacial conductance when the physical properties
786     of one side of the interface (notably the density) change rapidly as a
787     function of temperature.
788    
789 skuang 3730 \begin{table*}
790     \begin{minipage}{\linewidth}
791     \begin{center}
792 gezelter 3761 \caption{When toluene is the solvent, the interfacial thermal
793     conductivity is less sensitive to temperature, but again, the
794     direction of the heat flow can alter the solvent temperature
795     and can change the computed conductance values.}
796 skuang 3725
797 skuang 3738 \begin{tabular}{ccccc}
798 skuang 3725 \hline\hline
799 skuang 3738 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
800     (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
801 skuang 3725 \hline
802 skuang 3745 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
803     & & -1.86 & 180(3) & 135(21) \\
804     & & -3.93 & 176(5) & 113(12) \\
805 skuang 3738 \hline
806 skuang 3745 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
807     & & -4.19 & 135(9) & 113(12) \\
808 skuang 3725 \hline\hline
809     \end{tabular}
810     \label{AuThiolToluene}
811     \end{center}
812     \end{minipage}
813     \end{table*}
814    
815 gezelter 3756 Besides the lower interfacial thermal conductance, surfaces at
816     relatively high temperatures are susceptible to reconstructions,
817     particularly when butanethiols fully cover the Au(111) surface. These
818     reconstructions include surface Au atoms which migrate outward to the
819     S atom layer, and butanethiol molecules which embed into the surface
820     Au layer. The driving force for this behavior is the strong Au-S
821     interactions which are modeled here with a deep Lennard-Jones
822 gezelter 3761 potential. This phenomenon agrees with reconstructions that have been
823 gezelter 3756 experimentally
824     observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
825     {\it et al.} kept their Au(111) slab rigid so that their simulations
826     could reach 300K without surface
827     reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
828     blur the interface, the measurement of $G$ becomes more difficult to
829     conduct at higher temperatures. For this reason, most of our
830     measurements are undertaken at $\langle T\rangle\sim$200K where
831     reconstruction is minimized.
832 skuang 3725
833 skuang 3730 However, when the surface is not completely covered by butanethiols,
834 gezelter 3756 the simulated system appears to be more resistent to the
835 skuang 3760 reconstruction. Our Au / butanethiol / toluene system had the Au(111)
836 gezelter 3756 surfaces 90\% covered by butanethiols, but did not see this above
837     phenomena even at $\langle T\rangle\sim$300K. That said, we did
838     observe butanethiols migrating to neighboring three-fold sites during
839     a simulation. Since the interface persisted in these simulations,
840     were able to obtain $G$'s for these interfaces even at a relatively
841     high temperature without being affected by surface reconstructions.
842 skuang 3725
843 gezelter 3754 \section{Discussion}
844 skuang 3748
845 gezelter 3756 The primary result of this work is that the capping agent acts as an
846     efficient thermal coupler between solid and solvent phases. One of
847     the ways the capping agent can carry out this role is to down-shift
848     between the phonon vibrations in the solid (which carry the heat from
849     the gold) and the molecular vibrations in the liquid (which carry some
850     of the heat in the solvent).
851    
852 gezelter 3754 To investigate the mechanism of interfacial thermal conductance, the
853     vibrational power spectrum was computed. Power spectra were taken for
854     individual components in different simulations. To obtain these
855 gezelter 3756 spectra, simulations were run after equilibration in the
856     microcanonical (NVE) ensemble and without a thermal
857     gradient. Snapshots of configurations were collected at a frequency
858 gezelter 3761 that is higher than that of the fastest vibrations occurring in the
859 gezelter 3756 simulations. With these configurations, the velocity auto-correlation
860     functions can be computed:
861 gezelter 3754 \begin{equation}
862     C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
863     \label{vCorr}
864     \end{equation}
865     The power spectrum is constructed via a Fourier transform of the
866     symmetrized velocity autocorrelation function,
867     \begin{equation}
868     \hat{f}(\omega) =
869     \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
870     \label{fourier}
871     \end{equation}
872 skuang 3725
873 gezelter 3756 \subsection{The role of specific vibrations}
874 skuang 3747 The vibrational spectra for gold slabs in different environments are
875     shown as in Figure \ref{specAu}. Regardless of the presence of
876 gezelter 3756 solvent, the gold surfaces which are covered by butanethiol molecules
877     exhibit an additional peak observed at a frequency of
878 skuang 3759 $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
879 gezelter 3756 vibration. This vibration enables efficient thermal coupling of the
880     surface Au layer to the capping agents. Therefore, in our simulations,
881     the Au / S interfaces do not appear to be the primary barrier to
882     thermal transport when compared with the butanethiol / solvent
883     interfaces.
884 skuang 3732
885 skuang 3725 \begin{figure}
886     \includegraphics[width=\linewidth]{vibration}
887 gezelter 3761 \caption{The vibrational power spectrum for thiol-capped gold has an
888     additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
889     surfaces (both with and without a solvent over-layer) are missing
890     this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
891     the vibrational power spectrum for the butanethiol capping agents.}
892 skuang 3747 \label{specAu}
893 skuang 3725 \end{figure}
894    
895 gezelter 3756 Also in this figure, we show the vibrational power spectrum for the
896     bound butanethiol molecules, which also exhibits the same
897 skuang 3759 $\sim$165cm$^{-1}$ peak.
898 gezelter 3756
899     \subsection{Overlap of power spectra}
900 skuang 3755 A comparison of the results obtained from the two different organic
901     solvents can also provide useful information of the interfacial
902 gezelter 3756 thermal transport process. In particular, the vibrational overlap
903     between the butanethiol and the organic solvents suggests a highly
904     efficient thermal exchange between these components. Very high
905     thermal conductivity was observed when AA models were used and C-H
906     vibrations were treated classically. The presence of extra degrees of
907     freedom in the AA force field yields higher heat exchange rates
908     between the two phases and results in a much higher conductivity than
909     in the UA force field.
910 skuang 3732
911 gezelter 3756 The similarity in the vibrational modes available to solvent and
912     capping agent can be reduced by deuterating one of the two components
913     (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
914     are deuterated, one can observe a significantly lower $G$ and
915     $G^\prime$ values (Table \ref{modelTest}).
916    
917 skuang 3755 \begin{figure}
918 gezelter 3756 \includegraphics[width=\linewidth]{aahxntln}
919     \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
920     systems. When butanethiol is deuterated (lower left), its
921     vibrational overlap with hexane decreases significantly. Since
922     aromatic molecules and the butanethiol are vibrationally dissimilar,
923     the change is not as dramatic when toluene is the solvent (right).}
924     \label{aahxntln}
925     \end{figure}
926    
927     For the Au / butanethiol / toluene interfaces, having the AA
928     butanethiol deuterated did not yield a significant change in the
929     measured conductance. Compared to the C-H vibrational overlap between
930     hexane and butanethiol, both of which have alkyl chains, the overlap
931     between toluene and butanethiol is not as significant and thus does
932     not contribute as much to the heat exchange process.
933    
934     Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
935     that the {\it intra}molecular heat transport due to alkylthiols is
936     highly efficient. Combining our observations with those of Zhang {\it
937     et al.}, it appears that butanethiol acts as a channel to expedite
938     heat flow from the gold surface and into the alkyl chain. The
939     acoustic impedance mismatch between the metal and the liquid phase can
940     therefore be effectively reduced with the presence of suitable capping
941     agents.
942    
943     Deuterated models in the UA force field did not decouple the thermal
944     transport as well as in the AA force field. The UA models, even
945     though they have eliminated the high frequency C-H vibrational
946     overlap, still have significant overlap in the lower-frequency
947     portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
948     the UA models did not decouple the low frequency region enough to
949     produce an observable difference for the results of $G$ (Table
950     \ref{modelTest}).
951    
952     \begin{figure}
953 skuang 3755 \includegraphics[width=\linewidth]{uahxnua}
954 gezelter 3761 \caption{Vibrational power spectra for UA models for the butanethiol
955     and hexane solvent (upper panel) show the high degree of overlap
956     between these two molecules, particularly at lower frequencies.
957     Deuterating a UA model for the solvent (lower panel) does not
958     decouple the two spectra to the same degree as in the AA force
959     field (see Fig \ref{aahxntln}).}
960 skuang 3755 \label{uahxnua}
961     \end{figure}
962    
963 skuang 3730 \section{Conclusions}
964 gezelter 3756 The NIVS algorithm has been applied to simulations of
965     butanethiol-capped Au(111) surfaces in the presence of organic
966     solvents. This algorithm allows the application of unphysical thermal
967     flux to transfer heat between the metal and the liquid phase. With the
968     flux applied, we were able to measure the corresponding thermal
969     gradients and to obtain interfacial thermal conductivities. Under
970     steady states, 2-3 ns trajectory simulations are sufficient for
971     computation of this quantity.
972 skuang 3747
973 gezelter 3756 Our simulations have seen significant conductance enhancement in the
974     presence of capping agent, compared with the bare gold / liquid
975 skuang 3747 interfaces. The acoustic impedance mismatch between the metal and the
976 gezelter 3756 liquid phase is effectively eliminated by a chemically-bonded capping
977 gezelter 3761 agent. Furthermore, the coverage percentage of the capping agent plays
978 skuang 3747 an important role in the interfacial thermal transport
979 gezelter 3756 process. Moderately low coverages allow higher contact between capping
980     agent and solvent, and thus could further enhance the heat transfer
981     process, giving a non-monotonic behavior of conductance with
982     increasing coverage.
983 skuang 3725
984 gezelter 3756 Our results, particularly using the UA models, agree well with
985     available experimental data. The AA models tend to overestimate the
986 skuang 3732 interfacial thermal conductance in that the classically treated C-H
987 gezelter 3756 vibrations become too easily populated. Compared to the AA models, the
988     UA models have higher computational efficiency with satisfactory
989     accuracy, and thus are preferable in modeling interfacial thermal
990     transport.
991    
992     Of the two definitions for $G$, the discrete form
993 skuang 3747 (Eq. \ref{discreteG}) was easier to use and gives out relatively
994     consistent results, while the derivative form (Eq. \ref{derivativeG})
995     is not as versatile. Although $G^\prime$ gives out comparable results
996     and follows similar trend with $G$ when measuring close to fully
997 gezelter 3756 covered or bare surfaces, the spatial resolution of $T$ profile
998     required for the use of a derivative form is limited by the number of
999     bins and the sampling required to obtain thermal gradient information.
1000 skuang 3730
1001 gezelter 3756 Vlugt {\it et al.} have investigated the surface thiol structures for
1002     nanocrystalline gold and pointed out that they differ from those of
1003     the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
1004     difference could also cause differences in the interfacial thermal
1005     transport behavior. To investigate this problem, one would need an
1006     effective method for applying thermal gradients in non-planar
1007     (i.e. spherical) geometries.
1008 skuang 3730
1009 gezelter 3717 \section{Acknowledgments}
1010     Support for this project was provided by the National Science
1011     Foundation under grant CHE-0848243. Computational time was provided by
1012     the Center for Research Computing (CRC) at the University of Notre
1013 gezelter 3754 Dame.
1014     \newpage
1015 gezelter 3717
1016     \bibliography{interfacial}
1017    
1018     \end{doublespace}
1019     \end{document}
1020