ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3745
Committed: Thu Jul 21 00:04:26 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 46273 byte(s)
Log Message:
edit figures. finish all error estimates.

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 skuang 3725
48 skuang 3732 With the Non-Isotropic Velocity Scaling algorithm (NIVS) we have
49     developed, an unphysical thermal flux can be effectively set up even
50     for non-homogeneous systems like interfaces in non-equilibrium
51     molecular dynamics simulations. In this work, this algorithm is
52     applied for simulating thermal conductance at metal / organic solvent
53     interfaces with various coverages of butanethiol capping
54     agents. Different solvents and force field models were tested. Our
55     results suggest that the United-Atom models are able to provide an
56     estimate of the interfacial thermal conductivity comparable to
57     experiments in our simulations with satisfactory computational
58     efficiency. From our results, the acoustic impedance mismatch between
59     metal and liquid phase is effectively reduced by the capping
60     agents, and thus leads to interfacial thermal conductance
61     enhancement. Furthermore, this effect is closely related to the
62     capping agent coverage on the metal surfaces and the type of solvent
63     molecules, and is affected by the models used in the simulations.
64 skuang 3725
65 gezelter 3717 \end{abstract}
66    
67     \newpage
68    
69     %\narrowtext
70    
71     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72     % BODY OF TEXT
73     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74    
75     \section{Introduction}
76 skuang 3725 Interfacial thermal conductance is extensively studied both
77 skuang 3737 experimentally and computationally\cite{cahill:793}, due to its
78     importance in nanoscale science and technology. Reliability of
79     nanoscale devices depends on their thermal transport
80     properties. Unlike bulk homogeneous materials, nanoscale materials
81     features significant presence of interfaces, and these interfaces
82     could dominate the heat transfer behavior of these
83 skuang 3733 materials. Furthermore, these materials are generally heterogeneous,
84 skuang 3737 which challenges traditional research methods for homogeneous
85     systems.
86 gezelter 3717
87 skuang 3733 Heat conductance of molecular and nano-scale interfaces will be
88     affected by the chemical details of the surface. Experimentally,
89     various interfaces have been investigated for their thermal
90     conductance properties. Wang {\it et al.} studied heat transport
91     through long-chain hydrocarbon monolayers on gold substrate at
92     individual molecular level\cite{Wang10082007}; Schmidt {\it et al.}
93     studied the role of CTAB on thermal transport between gold nanorods
94     and solvent\cite{doi:10.1021/jp8051888}; Juv\'e {\it et al.} studied
95     the cooling dynamics, which is controlled by thermal interface
96     resistence of glass-embedded metal
97     nanoparticles\cite{PhysRevB.80.195406}. Although interfaces are
98     commonly barriers for heat transport, Alper {\it et al.} suggested
99     that specific ligands (capping agents) could completely eliminate this
100     barrier ($G\rightarrow\infty$)\cite{doi:10.1021/la904855s}.
101    
102 skuang 3737 Theoretical and computational models have also been used to study the
103     interfacial thermal transport in order to gain an understanding of
104     this phenomena at the molecular level. Recently, Hase and coworkers
105     employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
106     study thermal transport from hot Au(111) substrate to a self-assembled
107 skuang 3738 monolayer of alkylthiol with relatively long chain (8-20 carbon
108 skuang 3737 atoms)\cite{hase:2010,hase:2011}. However, ensemble averaged
109     measurements for heat conductance of interfaces between the capping
110     monolayer on Au and a solvent phase has yet to be studied.
111 skuang 3738 The comparatively low thermal flux through interfaces is
112 skuang 3736 difficult to measure with Equilibrium MD or forward NEMD simulation
113     methods. Therefore, the Reverse NEMD (RNEMD) methods would have the
114     advantage of having this difficult to measure flux known when studying
115     the thermal transport across interfaces, given that the simulation
116 skuang 3734 methods being able to effectively apply an unphysical flux in
117     non-homogeneous systems.
118    
119 skuang 3725 Recently, we have developed the Non-Isotropic Velocity Scaling (NIVS)
120     algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
121     retains the desirable features of RNEMD (conservation of linear
122     momentum and total energy, compatibility with periodic boundary
123     conditions) while establishing true thermal distributions in each of
124 skuang 3737 the two slabs. Furthermore, it allows effective thermal exchange
125     between particles of different identities, and thus makes the study of
126     interfacial conductance much simpler.
127 skuang 3725
128 skuang 3737 The work presented here deals with the Au(111) surface covered to
129     varying degrees by butanethiol, a capping agent with short carbon
130     chain, and solvated with organic solvents of different molecular
131     properties. Different models were used for both the capping agent and
132     the solvent force field parameters. Using the NIVS algorithm, the
133     thermal transport across these interfaces was studied and the
134 skuang 3734 underlying mechanism for this phenomena was investigated.
135 skuang 3733
136 skuang 3737 [MAY ADD WHY STUDY AU-THIOL SURFACE; CITE SHAOYI JIANG]
137 skuang 3734
138 skuang 3721 \section{Methodology}
139 skuang 3737 \subsection{Imposd-Flux Methods in MD Simulations}
140     For systems with low interfacial conductivity one must have a method
141     capable of generating relatively small fluxes, compared to those
142     required for bulk conductivity. This requirement makes the calculation
143     even more difficult for those slowly-converging equilibrium
144     methods\cite{Viscardy:2007lq}.
145     Forward methods impose gradient, but in interfacail conditions it is
146     not clear what behavior to impose at the boundary...
147     Imposed-flux reverse non-equilibrium
148 skuang 3721 methods\cite{MullerPlathe:1997xw} have the flux set {\it a priori} and
149 skuang 3737 the thermal response becomes easier to
150     measure than the flux. Although M\"{u}ller-Plathe's original momentum
151     swapping approach can be used for exchanging energy between particles
152     of different identity, the kinetic energy transfer efficiency is
153     affected by the mass difference between the particles, which limits
154     its application on heterogeneous interfacial systems.
155 skuang 3721
156 skuang 3737 The non-isotropic velocity scaling (NIVS)\cite{kuang:164101} approach to
157     non-equilibrium MD simulations is able to impose a wide range of
158     kinetic energy fluxes without obvious perturbation to the velocity
159     distributions of the simulated systems. Furthermore, this approach has
160 skuang 3721 the advantage in heterogeneous interfaces in that kinetic energy flux
161     can be applied between regions of particles of arbitary identity, and
162 skuang 3737 the flux will not be restricted by difference in particle mass.
163 skuang 3721
164     The NIVS algorithm scales the velocity vectors in two separate regions
165     of a simulation system with respective diagonal scaling matricies. To
166     determine these scaling factors in the matricies, a set of equations
167     including linear momentum conservation and kinetic energy conservation
168 skuang 3737 constraints and target energy flux satisfaction is solved. With the
169     scaling operation applied to the system in a set frequency, bulk
170     temperature gradients can be easily established, and these can be used
171     for computing thermal conductivities. The NIVS algorithm conserves
172     momenta and energy and does not depend on an external thermostat.
173 skuang 3721
174 skuang 3727 \subsection{Defining Interfacial Thermal Conductivity $G$}
175     For interfaces with a relatively low interfacial conductance, the bulk
176     regions on either side of an interface rapidly come to a state in
177     which the two phases have relatively homogeneous (but distinct)
178     temperatures. The interfacial thermal conductivity $G$ can therefore
179     be approximated as:
180     \begin{equation}
181     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
182     \langle T_\mathrm{cold}\rangle \right)}
183     \label{lowG}
184     \end{equation}
185     where ${E_{total}}$ is the imposed non-physical kinetic energy
186     transfer and ${\langle T_\mathrm{hot}\rangle}$ and ${\langle
187     T_\mathrm{cold}\rangle}$ are the average observed temperature of the
188     two separated phases.
189 skuang 3721
190 skuang 3737 When the interfacial conductance is {\it not} small, there are two
191     ways to define $G$.
192 skuang 3727
193 skuang 3737 One way is to assume the temperature is discrete on the two sides of
194     the interface. $G$ can be calculated using the applied thermal flux
195     $J$ and the maximum temperature difference measured along the thermal
196 skuang 3745 gradient max($\Delta T$), which occurs at the Gibbs deviding surface
197     (Figure \ref{demoPic}):
198 skuang 3727 \begin{equation}
199     G=\frac{J}{\Delta T}
200     \label{discreteG}
201     \end{equation}
202    
203 skuang 3745 \begin{figure}
204     \includegraphics[width=\linewidth]{method}
205     \caption{Interfacial conductance can be calculated by applying an
206     (unphysical) kinetic energy flux between two slabs, one located
207     within the metal and another on the edge of the periodic box. The
208     system responds by forming a thermal response or a gradient. In
209     bulk liquids, this gradient typically has a single slope, but in
210     interfacial systems, there are distinct thermal conductivity
211     domains. The interfacial conductance, $G$ is found by measuring the
212     temperature gap at the Gibbs dividing surface, or by using second
213     derivatives of the thermal profile.}
214     \label{demoPic}
215     \end{figure}
216    
217 skuang 3727 The other approach is to assume a continuous temperature profile along
218     the thermal gradient axis (e.g. $z$) and define $G$ at the point where
219     the magnitude of thermal conductivity $\lambda$ change reach its
220     maximum, given that $\lambda$ is well-defined throughout the space:
221     \begin{equation}
222     G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
223     = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
224     \left(\frac{\partial T}{\partial z}\right)\right)\Big|
225     = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
226     \Big/\left(\frac{\partial T}{\partial z}\right)^2
227     \label{derivativeG}
228     \end{equation}
229    
230     With the temperature profile obtained from simulations, one is able to
231     approximate the first and second derivatives of $T$ with finite
232 skuang 3737 difference methods and thus calculate $G^\prime$.
233 skuang 3727
234 skuang 3737 In what follows, both definitions have been used for calculation and
235     are compared in the results.
236 skuang 3727
237 skuang 3737 To compare the above definitions ($G$ and $G^\prime$), we have modeled
238     a metal slab with its (111) surfaces perpendicular to the $z$-axis of
239 skuang 3744 our simulation cells. Both with and without capping agents on the
240 skuang 3737 surfaces, the metal slab is solvated with simple organic solvents, as
241     illustrated in Figure \ref{demoPic}.
242 skuang 3727
243 skuang 3737 With the simulation cell described above, we are able to equilibrate
244     the system and impose an unphysical thermal flux between the liquid
245     and the metal phase using the NIVS algorithm. By periodically applying
246     the unphysical flux, we are able to obtain a temperature profile and
247     its spatial derivatives. These quantities enable the evaluation of the
248     interfacial thermal conductance of a surface. Figure \ref{gradT} is an
249     example how those applied thermal fluxes can be used to obtain the 1st
250     and 2nd derivatives of the temperature profile.
251 skuang 3727
252     \begin{figure}
253     \includegraphics[width=\linewidth]{gradT}
254 skuang 3745 \caption{A sample of Au-butanethiol/hexane interfacial system and the
255     temperature profile after a kinetic energy flux is imposed to
256     it. The 1st and 2nd derivatives of the temperature profile can be
257     obtained with finite difference approximation (lower panel).}
258 skuang 3727 \label{gradT}
259     \end{figure}
260    
261     \section{Computational Details}
262 skuang 3730 \subsection{Simulation Protocol}
263 skuang 3737 The NIVS algorithm has been implemented in our MD simulation code,
264     OpenMD\cite{Meineke:2005gd,openmd}, and was used for our
265     simulations. Different slab thickness (layer numbers of Au) were
266     simulated. Metal slabs were first equilibrated under atmospheric
267     pressure (1 atm) and a desired temperature (e.g. 200K). After
268     equilibration, butanethiol capping agents were placed at three-fold
269     sites on the Au(111) surfaces. The maximum butanethiol capacity on Au
270     surface is $1/3$ of the total number of surface Au
271     atoms\cite{vlugt:cpc2007154}. A series of different coverages was
272     investigated in order to study the relation between coverage and
273     interfacial conductance.
274 skuang 3727
275 skuang 3737 The capping agent molecules were allowed to migrate during the
276     simulations. They distributed themselves uniformly and sampled a
277     number of three-fold sites throughout out study. Therefore, the
278     initial configuration would not noticeably affect the sampling of a
279     variety of configurations of the same coverage, and the final
280     conductance measurement would be an average effect of these
281     configurations explored in the simulations. [MAY NEED FIGURES]
282 skuang 3727
283 skuang 3737 After the modified Au-butanethiol surface systems were equilibrated
284     under canonical ensemble, organic solvent molecules were packed in the
285     previously empty part of the simulation cells\cite{packmol}. Two
286     solvents were investigated, one which has little vibrational overlap
287     with the alkanethiol and a planar shape (toluene), and one which has
288     similar vibrational frequencies and chain-like shape ({\it n}-hexane).
289 skuang 3727
290 skuang 3737 The space filled by solvent molecules, i.e. the gap between
291 skuang 3730 periodically repeated Au-butanethiol surfaces should be carefully
292     chosen. A very long length scale for the thermal gradient axis ($z$)
293     may cause excessively hot or cold temperatures in the middle of the
294     solvent region and lead to undesired phenomena such as solvent boiling
295     or freezing when a thermal flux is applied. Conversely, too few
296     solvent molecules would change the normal behavior of the liquid
297     phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
298     these extreme cases did not happen to our simulations. And the
299 skuang 3745 corresponding spacing is usually $35 \sim 75$\AA.
300 skuang 3730
301 skuang 3728 The initial configurations generated by Packmol are further
302     equilibrated with the $x$ and $y$ dimensions fixed, only allowing
303     length scale change in $z$ dimension. This is to ensure that the
304     equilibration of liquid phase does not affect the metal crystal
305     structure in $x$ and $y$ dimensions. Further equilibration are run
306     under NVT and then NVE ensembles.
307    
308 skuang 3727 After the systems reach equilibrium, NIVS is implemented to impose a
309     periodic unphysical thermal flux between the metal and the liquid
310 skuang 3728 phase. Most of our simulations are under an average temperature of
311     $\sim$200K. Therefore, this flux usually comes from the metal to the
312 skuang 3727 liquid so that the liquid has a higher temperature and would not
313     freeze due to excessively low temperature. This induced temperature
314     gradient is stablized and the simulation cell is devided evenly into
315     N slabs along the $z$-axis and the temperatures of each slab are
316     recorded. When the slab width $d$ of each slab is the same, the
317     derivatives of $T$ with respect to slab number $n$ can be directly
318     used for $G^\prime$ calculations:
319     \begin{equation}
320     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
321     \Big/\left(\frac{\partial T}{\partial z}\right)^2
322     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
323     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
324     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
325     \Big/\left(\frac{\partial T}{\partial n}\right)^2
326     \label{derivativeG2}
327     \end{equation}
328    
329 skuang 3725 \subsection{Force Field Parameters}
330 skuang 3744 Our simulations include various components. Figure \ref{demoMol}
331     demonstrates the sites defined for both United-Atom and All-Atom
332     models of the organic solvent and capping agent molecules in our
333     simulations. Force field parameter descriptions are needed for
334     interactions both between the same type of particles and between
335     particles of different species.
336 skuang 3721
337 skuang 3736 \begin{figure}
338 gezelter 3740 \includegraphics[width=\linewidth]{structures}
339     \caption{Structures of the capping agent and solvents utilized in
340     these simulations. The chemically-distinct sites (a-e) are expanded
341     in terms of constituent atoms for both United Atom (UA) and All Atom
342     (AA) force fields. Most parameters are from
343     Refs. \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} (UA) and
344     \protect\cite{OPLSAA} (AA). Cross-interactions with the Au atoms are given
345     in Table \ref{MnM}.}
346 skuang 3736 \label{demoMol}
347     \end{figure}
348    
349 skuang 3744 The Au-Au interactions in metal lattice slab is described by the
350     quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59.3527}. The QSC
351     potentials include zero-point quantum corrections and are
352     reparametrized for accurate surface energies compared to the
353     Sutton-Chen potentials\cite{Chen90}.
354    
355 skuang 3728 For both solvent molecules, straight chain {\it n}-hexane and aromatic
356     toluene, United-Atom (UA) and All-Atom (AA) models are used
357     respectively. The TraPPE-UA
358     parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
359 skuang 3744 for our UA solvent molecules. In these models, sites are located at
360     the carbon centers for alkyl groups. Bonding interactions, including
361     bond stretches and bends and torsions, were used for intra-molecular
362     sites not separated by more than 3 bonds. Otherwise, for non-bonded
363     interactions, Lennard-Jones potentials are used. [MORE CITATION?]
364 skuang 3721
365 skuang 3744 By eliminating explicit hydrogen atoms, these models are simple and
366     computationally efficient, while maintains good accuracy. However, the
367     TraPPE-UA for alkanes is known to predict a lower boiling point than
368     experimental values. Considering that after an unphysical thermal flux
369     is applied to a system, the temperature of ``hot'' area in the liquid
370     phase would be significantly higher than the average, to prevent over
371     heating and boiling of the liquid phase, the average temperature in
372     our simulations should be much lower than the liquid boiling point.
373    
374     For UA-toluene model, the non-bonded potentials between
375     inter-molecular sites have a similar Lennard-Jones formulation. For
376     intra-molecular interactions, considering the stiffness of the benzene
377     ring, rigid body constraints are applied for further computational
378     efficiency. All bonds in the benzene ring and between the ring and the
379     methyl group remain rigid during the progress of simulations.
380    
381 skuang 3729 Besides the TraPPE-UA models, AA models for both organic solvents are
382 skuang 3730 included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
383 skuang 3744 force field is used. Additional explicit hydrogen sites were
384     included. Besides bonding and non-bonded site-site interactions,
385     partial charges and the electrostatic interactions were added to each
386     CT and HC site. For toluene, the United Force Field developed by
387     Rapp\'{e} {\it et al.}\cite{doi:10.1021/ja00051a040} is
388     adopted. Without the rigid body constraints, bonding interactions were
389     included. For the aromatic ring, improper torsions (inversions) were
390     added as an extra potential for maintaining the planar shape.
391 skuang 3745 [MORE CITATION?]
392 skuang 3728
393 skuang 3729 The capping agent in our simulations, the butanethiol molecules can
394     either use UA or AA model. The TraPPE-UA force fields includes
395 skuang 3730 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
396     UA butanethiol model in our simulations. The OPLS-AA also provides
397     parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
398     surfaces do not have the hydrogen atom bonded to sulfur. To adapt this
399     change and derive suitable parameters for butanethiol adsorbed on
400 skuang 3736 Au(111) surfaces, we adopt the S parameters from Luedtke and
401     Landman\cite{landman:1998} and modify parameters for its neighbor C
402     atom for charge balance in the molecule. Note that the model choice
403     (UA or AA) of capping agent can be different from the
404     solvent. Regardless of model choice, the force field parameters for
405     interactions between capping agent and solvent can be derived using
406 skuang 3738 Lorentz-Berthelot Mixing Rule:
407     \begin{eqnarray}
408 skuang 3742 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
409     \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
410 skuang 3738 \end{eqnarray}
411 skuang 3721
412     To describe the interactions between metal Au and non-metal capping
413 skuang 3730 agent and solvent particles, we refer to an adsorption study of alkyl
414     thiols on gold surfaces by Vlugt {\it et
415     al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
416     form of potential parameters for the interaction between Au and
417     pseudo-atoms CH$_x$ and S based on a well-established and widely-used
418 skuang 3736 effective potential of Hautman and Klein\cite{hautman:4994} for the
419     Au(111) surface. As our simulations require the gold lattice slab to
420     be non-rigid so that it could accommodate kinetic energy for thermal
421 skuang 3730 transport study purpose, the pair-wise form of potentials is
422     preferred.
423 skuang 3721
424 skuang 3730 Besides, the potentials developed from {\it ab initio} calculations by
425     Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
426 skuang 3744 interactions between Au and aromatic C/H atoms in toluene. A set of
427     pseudo Lennard-Jones parameters were provided for Au in their force
428     fields. By using the Mixing Rule, this can be used to derive pair-wise
429     potentials for non-bonded interactions between Au and non-metal sites.
430 skuang 3725
431 skuang 3730 However, the Lennard-Jones parameters between Au and other types of
432 skuang 3744 particles, such as All-Atom normal alkanes in our simulations are not
433     yet well-established. For these interactions, we attempt to derive
434     their parameters using the Mixing Rule. To do this, Au pseudo
435     Lennard-Jones parameters for ``Metal-non-Metal'' (MnM) interactions
436     were first extracted from the Au-CH$_x$ parameters by applying the
437     Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
438 skuang 3730 parameters in our simulations.
439 skuang 3729
440 skuang 3730 \begin{table*}
441     \begin{minipage}{\linewidth}
442     \begin{center}
443 gezelter 3741 \caption{Non-bonded interaction parameters (including cross
444     interactions with Au atoms) for both force fields used in this
445     work.}
446     \begin{tabular}{lllllll}
447 skuang 3730 \hline\hline
448 gezelter 3741 & Site & $\sigma_{ii}$ & $\epsilon_{ii}$ & $q_i$ &
449     $\sigma_{Au-i}$ & $\epsilon_{Au-i}$ \\
450     & & (\AA) & (kcal/mol) & ($e$) & (\AA) & (kcal/mol) \\
451 skuang 3730 \hline
452 gezelter 3741 United Atom (UA)
453     &CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
454     &CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
455     &CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
456     &CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
457     \hline
458     All Atom (AA)
459     &CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
460     &CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
461     &CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
462     &HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
463     &CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
464     &HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
465     \hline
466 skuang 3744 Both UA and AA
467     & S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
468 skuang 3730 \hline\hline
469     \end{tabular}
470     \label{MnM}
471     \end{center}
472     \end{minipage}
473     \end{table*}
474 skuang 3729
475    
476 skuang 3730 \section{Results and Discussions}
477     [MAY HAVE A BRIEF SUMMARY]
478     \subsection{How Simulation Parameters Affects $G$}
479     [MAY NOT PUT AT FIRST]
480     We have varied our protocol or other parameters of the simulations in
481     order to investigate how these factors would affect the measurement of
482     $G$'s. It turned out that while some of these parameters would not
483     affect the results substantially, some other changes to the
484     simulations would have a significant impact on the measurement
485     results.
486 skuang 3725
487 skuang 3730 In some of our simulations, we allowed $L_x$ and $L_y$ to change
488 skuang 3744 during equilibrating the liquid phase. Due to the stiffness of the
489     crystalline Au structure, $L_x$ and $L_y$ would not change noticeably
490     after equilibration. Although $L_z$ could fluctuate $\sim$1\% after a
491     system is fully equilibrated in the NPT ensemble, this fluctuation, as
492     well as those of $L_x$ and $L_y$ (which is significantly smaller),
493     would not be magnified on the calculated $G$'s, as shown in Table
494     \ref{AuThiolHexaneUA}. This insensivity to $L_i$ fluctuations allows
495     reliable measurement of $G$'s without the necessity of extremely
496     cautious equilibration process.
497 skuang 3725
498 skuang 3730 As stated in our computational details, the spacing filled with
499     solvent molecules can be chosen within a range. This allows some
500     change of solvent molecule numbers for the same Au-butanethiol
501     surfaces. We did this study on our Au-butanethiol/hexane
502     simulations. Nevertheless, the results obtained from systems of
503     different $N_{hexane}$ did not indicate that the measurement of $G$ is
504     susceptible to this parameter. For computational efficiency concern,
505     smaller system size would be preferable, given that the liquid phase
506     structure is not affected.
507    
508     Our NIVS algorithm allows change of unphysical thermal flux both in
509     direction and in quantity. This feature extends our investigation of
510     interfacial thermal conductance. However, the magnitude of this
511     thermal flux is not arbitary if one aims to obtain a stable and
512     reliable thermal gradient. A temperature profile would be
513     substantially affected by noise when $|J_z|$ has a much too low
514     magnitude; while an excessively large $|J_z|$ that overwhelms the
515     conductance capacity of the interface would prevent a thermal gradient
516     to reach a stablized steady state. NIVS has the advantage of allowing
517     $J$ to vary in a wide range such that the optimal flux range for $G$
518     measurement can generally be simulated by the algorithm. Within the
519     optimal range, we were able to study how $G$ would change according to
520     the thermal flux across the interface. For our simulations, we denote
521     $J_z$ to be positive when the physical thermal flux is from the liquid
522     to metal, and negative vice versa. The $G$'s measured under different
523 skuang 3744 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and
524     \ref{AuThiolToluene}. These results do not suggest that $G$ is
525     dependent on $J_z$ within this flux range. The linear response of flux
526     to thermal gradient simplifies our investigations in that we can rely
527     on $G$ measurement with only a couple $J_z$'s and do not need to test
528     a large series of fluxes.
529 skuang 3730
530 skuang 3725 \begin{table*}
531     \begin{minipage}{\linewidth}
532     \begin{center}
533     \caption{Computed interfacial thermal conductivity ($G$ and
534 skuang 3731 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
535     interfaces with UA model and different hexane molecule numbers
536 skuang 3745 at different temperatures using a range of energy
537     fluxes. Error estimates indicated in parenthesis.}
538 skuang 3730
539 skuang 3738 \begin{tabular}{ccccccc}
540 skuang 3730 \hline\hline
541 skuang 3738 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
542     $J_z$ & $G$ & $G^\prime$ \\
543     (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
544 skuang 3730 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
545     \hline
546 skuang 3745 200 & 266 & No & 0.672 & -0.96 & 102(3) & 80.0(0.8) \\
547 skuang 3743 & 200 & Yes & 0.694 & 1.92 & 129(11) & 87.3(0.3) \\
548     & & Yes & 0.672 & 1.93 & 131(16) & 78(13) \\
549 skuang 3745 & & No & 0.688 & 0.96 & 125(16) & 90.2(15) \\
550 skuang 3743 & & & & 1.91 & 139(10) & 101(10) \\
551     & & & & 2.83 & 141(6) & 89.9(9.8) \\
552     & 166 & Yes & 0.679 & 0.97 & 115(19) & 69(18) \\
553     & & & & 1.94 & 125(9) & 87.1(0.2) \\
554     & & No & 0.681 & 0.97 & 141(30) & 78(22) \\
555     & & & & 1.92 & 138(4) & 98.9(9.5) \\
556 skuang 3739 \hline
557 skuang 3743 250 & 200 & No & 0.560 & 0.96 & 75(10) & 61.8(7.3) \\
558     & & & & -0.95 & 49.4(0.3) & 45.7(2.1) \\
559     & 166 & Yes & 0.570 & 0.98 & 79.0(3.5) & 62.9(3.0) \\
560     & & No & 0.569 & 0.97 & 80.3(0.6) & 67(11) \\
561     & & & & 1.44 & 76.2(5.0) & 64.8(3.8) \\
562     & & & & -0.95 & 56.4(2.5) & 54.4(1.1) \\
563     & & & & -1.85 & 47.8(1.1) & 53.5(1.5) \\
564 skuang 3730 \hline\hline
565     \end{tabular}
566     \label{AuThiolHexaneUA}
567     \end{center}
568     \end{minipage}
569     \end{table*}
570    
571     Furthermore, we also attempted to increase system average temperatures
572     to above 200K. These simulations are first equilibrated in the NPT
573     ensemble under normal pressure. As stated above, the TraPPE-UA model
574     for hexane tends to predict a lower boiling point. In our simulations,
575     hexane had diffculty to remain in liquid phase when NPT equilibration
576     temperature is higher than 250K. Additionally, the equilibrated liquid
577     hexane density under 250K becomes lower than experimental value. This
578     expanded liquid phase leads to lower contact between hexane and
579 skuang 3744 butanethiol as well.[MAY NEED SLAB DENSITY FIGURE]
580     And this reduced contact would
581 skuang 3730 probably be accountable for a lower interfacial thermal conductance,
582     as shown in Table \ref{AuThiolHexaneUA}.
583    
584     A similar study for TraPPE-UA toluene agrees with the above result as
585     well. Having a higher boiling point, toluene tends to remain liquid in
586     our simulations even equilibrated under 300K in NPT
587     ensembles. Furthermore, the expansion of the toluene liquid phase is
588     not as significant as that of the hexane. This prevents severe
589     decrease of liquid-capping agent contact and the results (Table
590     \ref{AuThiolToluene}) show only a slightly decreased interface
591     conductance. Therefore, solvent-capping agent contact should play an
592     important role in the thermal transport process across the interface
593     in that higher degree of contact could yield increased conductance.
594    
595     \begin{table*}
596     \begin{minipage}{\linewidth}
597     \begin{center}
598     \caption{Computed interfacial thermal conductivity ($G$ and
599 skuang 3731 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
600     interface at different temperatures using a range of energy
601 skuang 3745 fluxes. Error estimates indicated in parenthesis.}
602 skuang 3725
603 skuang 3738 \begin{tabular}{ccccc}
604 skuang 3725 \hline\hline
605 skuang 3738 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
606     (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
607 skuang 3725 \hline
608 skuang 3745 200 & 0.933 & 2.15 & 204(12) & 113(12) \\
609     & & -1.86 & 180(3) & 135(21) \\
610     & & -3.93 & 176(5) & 113(12) \\
611 skuang 3738 \hline
612 skuang 3745 300 & 0.855 & -1.91 & 143(5) & 125(2) \\
613     & & -4.19 & 135(9) & 113(12) \\
614 skuang 3725 \hline\hline
615     \end{tabular}
616     \label{AuThiolToluene}
617     \end{center}
618     \end{minipage}
619     \end{table*}
620    
621 skuang 3730 Besides lower interfacial thermal conductance, surfaces in relatively
622     high temperatures are susceptible to reconstructions, when
623     butanethiols have a full coverage on the Au(111) surface. These
624     reconstructions include surface Au atoms migrated outward to the S
625     atom layer, and butanethiol molecules embedded into the original
626     surface Au layer. The driving force for this behavior is the strong
627     Au-S interactions in our simulations. And these reconstructions lead
628     to higher ratio of Au-S attraction and thus is energetically
629     favorable. Furthermore, this phenomenon agrees with experimental
630     results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
631     {\it et al.} had kept their Au(111) slab rigid so that their
632     simulations can reach 300K without surface reconstructions. Without
633     this practice, simulating 100\% thiol covered interfaces under higher
634     temperatures could hardly avoid surface reconstructions. However, our
635     measurement is based on assuming homogeneity on $x$ and $y$ dimensions
636     so that measurement of $T$ at particular $z$ would be an effective
637     average of the particles of the same type. Since surface
638     reconstructions could eliminate the original $x$ and $y$ dimensional
639     homogeneity, measurement of $G$ is more difficult to conduct under
640     higher temperatures. Therefore, most of our measurements are
641 skuang 3732 undertaken at $\langle T\rangle\sim$200K.
642 skuang 3725
643 skuang 3730 However, when the surface is not completely covered by butanethiols,
644     the simulated system is more resistent to the reconstruction
645 skuang 3744 above. Our Au-butanethiol/toluene system had the Au(111) surfaces 90\%
646     covered by butanethiols, but did not see this above phenomena even at
647     $\langle T\rangle\sim$300K. The empty three-fold sites not occupied by
648     capping agents could help prevent surface reconstruction in that they
649     provide other means of capping agent relaxation. It is observed that
650 skuang 3738 butanethiols can migrate to their neighbor empty sites during a
651     simulation. Therefore, we were able to obtain $G$'s for these
652     interfaces even at a relatively high temperature without being
653     affected by surface reconstructions.
654 skuang 3725
655 skuang 3730 \subsection{Influence of Capping Agent Coverage on $G$}
656     To investigate the influence of butanethiol coverage on interfacial
657     thermal conductance, a series of different coverage Au-butanethiol
658     surfaces is prepared and solvated with various organic
659     molecules. These systems are then equilibrated and their interfacial
660 skuang 3744 thermal conductivity are measured with our NIVS algorithm. Figure
661     \ref{coverage} demonstrates the trend of conductance change with
662     respect to different coverages of butanethiol. To study the isotope
663     effect in interfacial thermal conductance, deuterated UA-hexane is
664     included as well.
665 skuang 3730
666 skuang 3731 It turned out that with partial covered butanethiol on the Au(111)
667 skuang 3744 surface, the derivative definition for $G^\prime$
668     (Eq. \ref{derivativeG}) was difficult to apply, due to the difficulty
669     in locating the maximum of change of $\lambda$. Instead, the discrete
670     definition (Eq. \ref{discreteG}) is easier to apply, as the Gibbs
671     deviding surface can still be well-defined. Therefore, $G$ (not
672     $G^\prime$) was used for this section.
673 skuang 3725
674 skuang 3744 From Figure \ref{coverage}, one can see the significance of the
675 skuang 3731 presence of capping agents. Even when a fraction of the Au(111)
676     surface sites are covered with butanethiols, the conductivity would
677     see an enhancement by at least a factor of 3. This indicates the
678     important role cappping agent is playing for thermal transport
679 skuang 3744 phenomena on metal / organic solvent surfaces.
680 skuang 3725
681 skuang 3731 Interestingly, as one could observe from our results, the maximum
682     conductance enhancement (largest $G$) happens while the surfaces are
683     about 75\% covered with butanethiols. This again indicates that
684     solvent-capping agent contact has an important role of the thermal
685     transport process. Slightly lower butanethiol coverage allows small
686     gaps between butanethiols to form. And these gaps could be filled with
687     solvent molecules, which acts like ``heat conductors'' on the
688     surface. The higher degree of interaction between these solvent
689     molecules and capping agents increases the enhancement effect and thus
690     produces a higher $G$ than densely packed butanethiol arrays. However,
691     once this maximum conductance enhancement is reached, $G$ decreases
692     when butanethiol coverage continues to decrease. Each capping agent
693     molecule reaches its maximum capacity for thermal
694     conductance. Therefore, even higher solvent-capping agent contact
695     would not offset this effect. Eventually, when butanethiol coverage
696     continues to decrease, solvent-capping agent contact actually
697     decreases with the disappearing of butanethiol molecules. In this
698 skuang 3744 case, $G$ decrease could not be offset but instead accelerated. [NEED
699     SNAPSHOT SHOWING THE PHENOMENA]
700 skuang 3725
701 skuang 3731 A comparison of the results obtained from differenet organic solvents
702     can also provide useful information of the interfacial thermal
703     transport process. The deuterated hexane (UA) results do not appear to
704     be much different from those of normal hexane (UA), given that
705     butanethiol (UA) is non-deuterated for both solvents. These UA model
706     studies, even though eliminating C-H vibration samplings, still have
707     C-C vibrational frequencies different from each other. However, these
708 skuang 3732 differences in the infrared range do not seem to produce an observable
709 skuang 3744 difference for the results of $G$. [MAY NEED SPECTRA FIGURE]
710 skuang 3730
711 skuang 3731 Furthermore, results for rigid body toluene solvent, as well as other
712     UA-hexane solvents, are reasonable within the general experimental
713     ranges[CITATIONS]. This suggests that explicit hydrogen might not be a
714     required factor for modeling thermal transport phenomena of systems
715     such as Au-thiol/organic solvent.
716    
717     However, results for Au-butanethiol/toluene do not show an identical
718 skuang 3744 trend with those for Au-butanethiol/hexane in that $G$ remains at
719 skuang 3731 approximately the same magnitue when butanethiol coverage differs from
720     25\% to 75\%. This might be rooted in the molecule shape difference
721 skuang 3744 for planar toluene and chain-like {\it n}-hexane. Due to this
722 skuang 3731 difference, toluene molecules have more difficulty in occupying
723     relatively small gaps among capping agents when their coverage is not
724     too low. Therefore, the solvent-capping agent contact may keep
725     increasing until the capping agent coverage reaches a relatively low
726     level. This becomes an offset for decreasing butanethiol molecules on
727     its effect to the process of interfacial thermal transport. Thus, one
728     can see a plateau of $G$ vs. butanethiol coverage in our results.
729    
730 skuang 3739 \begin{figure}
731     \includegraphics[width=\linewidth]{coverage}
732     \caption{Comparison of interfacial thermal conductivity ($G$) values
733     for the Au-butanethiol/solvent interface with various UA models and
734     different capping agent coverages at $\langle T\rangle\sim$200K
735     using certain energy flux respectively.}
736     \label{coverage}
737     \end{figure}
738 skuang 3725
739 skuang 3730 \subsection{Influence of Chosen Molecule Model on $G$}
740     [MAY COMBINE W MECHANISM STUDY]
741    
742 skuang 3732 In addition to UA solvent/capping agent models, AA models are included
743     in our simulations as well. Besides simulations of the same (UA or AA)
744     model for solvent and capping agent, different models can be applied
745     to different components. Furthermore, regardless of models chosen,
746     either the solvent or the capping agent can be deuterated, similar to
747     the previous section. Table \ref{modelTest} summarizes the results of
748     these studies.
749 skuang 3725
750     \begin{table*}
751     \begin{minipage}{\linewidth}
752     \begin{center}
753    
754     \caption{Computed interfacial thermal conductivity ($G$ and
755 skuang 3732 $G^\prime$) values for interfaces using various models for
756     solvent and capping agent (or without capping agent) at
757 skuang 3739 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
758     or capping agent molecules; ``Avg.'' denotes results that are
759 skuang 3742 averages of simulations under different $J_z$'s. Error
760     estimates indicated in parenthesis.)}
761 skuang 3725
762 skuang 3742 \begin{tabular}{llccc}
763 skuang 3725 \hline\hline
764 skuang 3732 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
765     (or bare surface) & model & (GW/m$^2$) &
766     \multicolumn{2}{c}{(MW/m$^2$/K)} \\
767 skuang 3725 \hline
768 skuang 3742 UA & UA hexane & Avg. & 131(9) & 87(10) \\
769     & UA hexane(D) & 1.95 & 153(5) & 136(13) \\
770     & AA hexane & Avg. & 131(6) & 122(10) \\
771     & UA toluene & 1.96 & 187(16) & 151(11) \\
772     & AA toluene & 1.89 & 200(36) & 149(53) \\
773 skuang 3739 \hline
774 skuang 3742 AA & UA hexane & 1.94 & 116(9) & 129(8) \\
775     & AA hexane & Avg. & 442(14) & 356(31) \\
776     & AA hexane(D) & 1.93 & 222(12) & 234(54) \\
777     & UA toluene & 1.98 & 125(25) & 97(60) \\
778     & AA toluene & 3.79 & 487(56) & 290(42) \\
779 skuang 3739 \hline
780 skuang 3742 AA(D) & UA hexane & 1.94 & 158(25) & 172(4) \\
781     & AA hexane & 1.92 & 243(29) & 191(11) \\
782     & AA toluene & 1.93 & 364(36) & 322(67) \\
783 skuang 3739 \hline
784 skuang 3742 bare & UA hexane & Avg. & 46.5(3.2) & 49.4(4.5) \\
785     & UA hexane(D) & 0.98 & 43.9(4.6) & 43.0(2.0) \\
786     & AA hexane & 0.96 & 31.0(1.4) & 29.4(1.3) \\
787     & UA toluene & 1.99 & 70.1(1.3) & 65.8(0.5) \\
788 skuang 3725 \hline\hline
789     \end{tabular}
790 skuang 3732 \label{modelTest}
791 skuang 3725 \end{center}
792     \end{minipage}
793     \end{table*}
794    
795 skuang 3732 To facilitate direct comparison, the same system with differnt models
796     for different components uses the same length scale for their
797     simulation cells. Without the presence of capping agent, using
798     different models for hexane yields similar results for both $G$ and
799     $G^\prime$, and these two definitions agree with eath other very
800     well. This indicates very weak interaction between the metal and the
801     solvent, and is a typical case for acoustic impedance mismatch between
802     these two phases.
803 skuang 3730
804 skuang 3732 As for Au(111) surfaces completely covered by butanethiols, the choice
805     of models for capping agent and solvent could impact the measurement
806     of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
807     interfaces, using AA model for both butanethiol and hexane yields
808     substantially higher conductivity values than using UA model for at
809     least one component of the solvent and capping agent, which exceeds
810 skuang 3744 the general range of experimental measurement results. This is
811     probably due to the classically treated C-H vibrations in the AA
812     model, which should not be appreciably populated at normal
813     temperatures. In comparison, once either the hexanes or the
814     butanethiols are deuterated, one can see a significantly lower $G$ and
815     $G^\prime$. In either of these cases, the C-H(D) vibrational overlap
816     between the solvent and the capping agent is removed.
817     [MAY NEED SPECTRA FIGURE] Conclusively, the
818 skuang 3732 improperly treated C-H vibration in the AA model produced
819     over-predicted results accordingly. Compared to the AA model, the UA
820     model yields more reasonable results with higher computational
821     efficiency.
822 skuang 3731
823 skuang 3732 However, for Au-butanethiol/toluene interfaces, having the AA
824     butanethiol deuterated did not yield a significant change in the
825 skuang 3739 measurement results. Compared to the C-H vibrational overlap between
826     hexane and butanethiol, both of which have alkyl chains, that overlap
827     between toluene and butanethiol is not so significant and thus does
828     not have as much contribution to the ``Intramolecular Vibration
829     Redistribution''[CITE HASE]. Conversely, extra degrees of freedom such
830     as the C-H vibrations could yield higher heat exchange rate between
831     these two phases and result in a much higher conductivity.
832 skuang 3731
833 skuang 3732 Although the QSC model for Au is known to predict an overly low value
834 skuang 3738 for bulk metal gold conductivity\cite{kuang:164101}, our computational
835 skuang 3732 results for $G$ and $G^\prime$ do not seem to be affected by this
836 skuang 3739 drawback of the model for metal. Instead, our results suggest that the
837     modeling of interfacial thermal transport behavior relies mainly on
838     the accuracy of the interaction descriptions between components
839     occupying the interfaces.
840 skuang 3732
841 skuang 3730 \subsection{Mechanism of Interfacial Thermal Conductance Enhancement
842     by Capping Agent}
843 skuang 3744 [OR: Vibrational Spectrum Study on Conductance Mechanism]
844 skuang 3730
845 skuang 3732 [MAY INTRODUCE PROTOCOL IN METHODOLOGY/COMPUTATIONAL DETAIL, EQN'S]
846 skuang 3730
847 skuang 3725 To investigate the mechanism of this interfacial thermal conductance,
848     the vibrational spectra of various gold systems were obtained and are
849     shown as in the upper panel of Fig. \ref{vibration}. To obtain these
850     spectra, one first runs a simulation in the NVE ensemble and collects
851     snapshots of configurations; these configurations are used to compute
852     the velocity auto-correlation functions, which is used to construct a
853 skuang 3732 power spectrum via a Fourier transform.
854 skuang 3725
855 skuang 3739 [MAY RELATE TO HASE'S]
856 skuang 3745 The gold surfaces covered by butanethiol molecules, compared to bare
857     gold surfaces, exhibit an additional peak observed at the frequency of
858     $\sim$170cm$^{-1}$, which is attributed to the S-Au bonding
859     vibration. This vibration enables efficient thermal transport from
860     surface Au layer to the capping agents.
861     [MAY PUT IN OTHER SECTION] Simultaneously, as shown in
862     the lower panel of Fig. \ref{vibration}, the large overlap of the
863     vibration spectra of butanethiol and hexane in the All-Atom model,
864     including the C-H vibration, also suggests high thermal exchange
865     efficiency. The combination of these two effects produces the drastic
866     interfacial thermal conductance enhancement in the All-Atom model.
867 skuang 3732
868 skuang 3745 [NEED SEPARATE FIGURE. MAY NEED TO CONVERT TO JPEG]
869 skuang 3725 \begin{figure}
870     \includegraphics[width=\linewidth]{vibration}
871     \caption{Vibrational spectra obtained for gold in different
872 skuang 3745 environments.}
873 skuang 3725 \label{vibration}
874     \end{figure}
875    
876 skuang 3744 [MAY ADD COMPARISON OF G AND G', AU SLAB WIDTHS, ETC]
877 skuang 3732 % The results show that the two definitions used for $G$ yield
878     % comparable values, though $G^\prime$ tends to be smaller.
879    
880 skuang 3730 \section{Conclusions}
881 skuang 3732 The NIVS algorithm we developed has been applied to simulations of
882     Au-butanethiol surfaces with organic solvents. This algorithm allows
883     effective unphysical thermal flux transferred between the metal and
884     the liquid phase. With the flux applied, we were able to measure the
885     corresponding thermal gradient and to obtain interfacial thermal
886     conductivities. Our simulations have seen significant conductance
887     enhancement with the presence of capping agent, compared to the bare
888 skuang 3744 gold / liquid interfaces. The acoustic impedance mismatch between the
889 skuang 3732 metal and the liquid phase is effectively eliminated by proper capping
890     agent. Furthermore, the coverage precentage of the capping agent plays
891     an important role in the interfacial thermal transport process.
892 skuang 3725
893 skuang 3732 Our measurement results, particularly of the UA models, agree with
894     available experimental data. This indicates that our force field
895     parameters have a nice description of the interactions between the
896     particles at the interfaces. AA models tend to overestimate the
897     interfacial thermal conductance in that the classically treated C-H
898     vibration would be overly sampled. Compared to the AA models, the UA
899     models have higher computational efficiency with satisfactory
900     accuracy, and thus are preferable in interfacial thermal transport
901     modelings.
902 skuang 3730
903 skuang 3732 Vlugt {\it et al.} has investigated the surface thiol structures for
904     nanocrystal gold and pointed out that they differs from those of the
905     Au(111) surface\cite{vlugt:cpc2007154}. This difference might lead to
906     change of interfacial thermal transport behavior as well. To
907     investigate this problem, an effective means to introduce thermal flux
908     and measure the corresponding thermal gradient is desirable for
909     simulating structures with spherical symmetry.
910 skuang 3730
911 skuang 3732
912 gezelter 3717 \section{Acknowledgments}
913     Support for this project was provided by the National Science
914     Foundation under grant CHE-0848243. Computational time was provided by
915     the Center for Research Computing (CRC) at the University of Notre
916 skuang 3730 Dame. \newpage
917 gezelter 3717
918     \bibliography{interfacial}
919    
920     \end{doublespace}
921     \end{document}
922