ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3744
Committed: Wed Jul 20 19:02:46 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 46070 byte(s)
Log Message:
text finished except introduction and abstract

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