ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3738
Committed: Wed Jul 13 23:01:52 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 42746 byte(s)
Log Message:
add solvent density data. modified tables.

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 3717 \bibpunct{[}{]}{,}{s}{}{;}
27     \bibliographystyle{aip}
28    
29     \begin{document}
30    
31     \title{Simulating interfacial thermal conductance at metal-solvent
32     interfaces: the role of chemical capping agents}
33    
34     \author{Shenyu Kuang and J. Daniel
35     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
36     Department of Chemistry and Biochemistry,\\
37     University of Notre Dame\\
38     Notre Dame, Indiana 46556}
39    
40     \date{\today}
41    
42     \maketitle
43    
44     \begin{doublespace}
45    
46     \begin{abstract}
47 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     our simulation cells. Both with and withour capping agents on the
226     surfaces, the metal slab is solvated with simple organic solvents, as
227     illustrated in Figure \ref{demoPic}.
228 skuang 3727
229     \begin{figure}
230     \includegraphics[width=\linewidth]{demoPic}
231     \caption{A sample showing how a metal slab has its (111) surface
232     covered by capping agent molecules and solvated by hexane.}
233     \label{demoPic}
234     \end{figure}
235    
236 skuang 3737 With the simulation cell described above, we are able to equilibrate
237     the system and impose an unphysical thermal flux between the liquid
238     and the metal phase using the NIVS algorithm. By periodically applying
239     the unphysical flux, we are able to obtain a temperature profile and
240     its spatial derivatives. These quantities enable the evaluation of the
241     interfacial thermal conductance of a surface. Figure \ref{gradT} is an
242     example how those applied thermal fluxes can be used to obtain the 1st
243     and 2nd derivatives of the temperature profile.
244 skuang 3727
245     \begin{figure}
246     \includegraphics[width=\linewidth]{gradT}
247     \caption{The 1st and 2nd derivatives of temperature profile can be
248     obtained with finite difference approximation.}
249     \label{gradT}
250     \end{figure}
251    
252     \section{Computational Details}
253 skuang 3730 \subsection{Simulation Protocol}
254 skuang 3737 The NIVS algorithm has been implemented in our MD simulation code,
255     OpenMD\cite{Meineke:2005gd,openmd}, and was used for our
256     simulations. Different slab thickness (layer numbers of Au) were
257     simulated. Metal slabs were first equilibrated under atmospheric
258     pressure (1 atm) and a desired temperature (e.g. 200K). After
259     equilibration, butanethiol capping agents were placed at three-fold
260     sites on the Au(111) surfaces. The maximum butanethiol capacity on Au
261     surface is $1/3$ of the total number of surface Au
262     atoms\cite{vlugt:cpc2007154}. A series of different coverages was
263     investigated in order to study the relation between coverage and
264     interfacial conductance.
265 skuang 3727
266 skuang 3737 The capping agent molecules were allowed to migrate during the
267     simulations. They distributed themselves uniformly and sampled a
268     number of three-fold sites throughout out study. Therefore, the
269     initial configuration would not noticeably affect the sampling of a
270     variety of configurations of the same coverage, and the final
271     conductance measurement would be an average effect of these
272     configurations explored in the simulations. [MAY NEED FIGURES]
273 skuang 3727
274 skuang 3737 After the modified Au-butanethiol surface systems were equilibrated
275     under canonical ensemble, organic solvent molecules were packed in the
276     previously empty part of the simulation cells\cite{packmol}. Two
277     solvents were investigated, one which has little vibrational overlap
278     with the alkanethiol and a planar shape (toluene), and one which has
279     similar vibrational frequencies and chain-like shape ({\it n}-hexane).
280 skuang 3727
281 skuang 3737 The space filled by solvent molecules, i.e. the gap between
282 skuang 3730 periodically repeated Au-butanethiol surfaces should be carefully
283     chosen. A very long length scale for the thermal gradient axis ($z$)
284     may cause excessively hot or cold temperatures in the middle of the
285     solvent region and lead to undesired phenomena such as solvent boiling
286     or freezing when a thermal flux is applied. Conversely, too few
287     solvent molecules would change the normal behavior of the liquid
288     phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
289     these extreme cases did not happen to our simulations. And the
290     corresponding spacing is usually $35 \sim 60$\AA.
291    
292 skuang 3728 The initial configurations generated by Packmol are further
293     equilibrated with the $x$ and $y$ dimensions fixed, only allowing
294     length scale change in $z$ dimension. This is to ensure that the
295     equilibration of liquid phase does not affect the metal crystal
296     structure in $x$ and $y$ dimensions. Further equilibration are run
297     under NVT and then NVE ensembles.
298    
299 skuang 3727 After the systems reach equilibrium, NIVS is implemented to impose a
300     periodic unphysical thermal flux between the metal and the liquid
301 skuang 3728 phase. Most of our simulations are under an average temperature of
302     $\sim$200K. Therefore, this flux usually comes from the metal to the
303 skuang 3727 liquid so that the liquid has a higher temperature and would not
304     freeze due to excessively low temperature. This induced temperature
305     gradient is stablized and the simulation cell is devided evenly into
306     N slabs along the $z$-axis and the temperatures of each slab are
307     recorded. When the slab width $d$ of each slab is the same, the
308     derivatives of $T$ with respect to slab number $n$ can be directly
309     used for $G^\prime$ calculations:
310     \begin{equation}
311     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
312     \Big/\left(\frac{\partial T}{\partial z}\right)^2
313     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
314     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
315     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
316     \Big/\left(\frac{\partial T}{\partial n}\right)^2
317     \label{derivativeG2}
318     \end{equation}
319    
320 skuang 3725 \subsection{Force Field Parameters}
321 skuang 3728 Our simulations include various components. Therefore, force field
322     parameter descriptions are needed for interactions both between the
323     same type of particles and between particles of different species.
324 skuang 3721
325     The Au-Au interactions in metal lattice slab is described by the
326 skuang 3736 quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59.3527}. The QSC
327 skuang 3721 potentials include zero-point quantum corrections and are
328     reparametrized for accurate surface energies compared to the
329     Sutton-Chen potentials\cite{Chen90}.
330    
331 skuang 3736 Figure \ref{demoMol} demonstrates how we name our pseudo-atoms of the
332     organic solvent molecules in our simulations.
333 skuang 3730
334 skuang 3736 \begin{figure}
335     \includegraphics[width=\linewidth]{demoMol}
336     \caption{Denomination of atoms or pseudo-atoms in our simulations: a)
337     UA-hexane; b) AA-hexane; c) UA-toluene; d) AA-toluene.}
338     \label{demoMol}
339     \end{figure}
340    
341 skuang 3728 For both solvent molecules, straight chain {\it n}-hexane and aromatic
342     toluene, United-Atom (UA) and All-Atom (AA) models are used
343     respectively. The TraPPE-UA
344     parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
345     for our UA solvent molecules. In these models, pseudo-atoms are
346     located at the carbon centers for alkyl groups. By eliminating
347     explicit hydrogen atoms, these models are simple and computationally
348 skuang 3729 efficient, while maintains good accuracy. However, the TraPPE-UA for
349     alkanes is known to predict a lower boiling point than experimental
350     values. Considering that after an unphysical thermal flux is applied
351     to a system, the temperature of ``hot'' area in the liquid phase would be
352     significantly higher than the average, to prevent over heating and
353     boiling of the liquid phase, the average temperature in our
354 skuang 3730 simulations should be much lower than the liquid boiling point. [MORE DISCUSSION]
355 skuang 3729 For UA-toluene model, rigid body constraints are applied, so that the
356 skuang 3730 benzene ring and the methyl-CRar bond are kept rigid. This would save
357     computational time.[MORE DETAILS]
358 skuang 3721
359 skuang 3729 Besides the TraPPE-UA models, AA models for both organic solvents are
360 skuang 3730 included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
361     force field is used. [MORE DETAILS]
362 skuang 3729 For toluene, the United Force Field developed by Rapp\'{e} {\it et
363     al.}\cite{doi:10.1021/ja00051a040} is adopted.[MORE DETAILS]
364 skuang 3728
365 skuang 3729 The capping agent in our simulations, the butanethiol molecules can
366     either use UA or AA model. The TraPPE-UA force fields includes
367 skuang 3730 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
368     UA butanethiol model in our simulations. The OPLS-AA also provides
369     parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
370     surfaces do not have the hydrogen atom bonded to sulfur. To adapt this
371     change and derive suitable parameters for butanethiol adsorbed on
372 skuang 3736 Au(111) surfaces, we adopt the S parameters from Luedtke and
373     Landman\cite{landman:1998} and modify parameters for its neighbor C
374     atom for charge balance in the molecule. Note that the model choice
375     (UA or AA) of capping agent can be different from the
376     solvent. Regardless of model choice, the force field parameters for
377     interactions between capping agent and solvent can be derived using
378 skuang 3738 Lorentz-Berthelot Mixing Rule:
379     \begin{eqnarray}
380     \sigma_{IJ} & = & \frac{1}{2} \left(\sigma_{II} + \sigma_{JJ}\right) \\
381     \epsilon_{IJ} & = & \sqrt{\epsilon_{II}\epsilon_{JJ}}
382     \end{eqnarray}
383 skuang 3721
384     To describe the interactions between metal Au and non-metal capping
385 skuang 3730 agent and solvent particles, we refer to an adsorption study of alkyl
386     thiols on gold surfaces by Vlugt {\it et
387     al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
388     form of potential parameters for the interaction between Au and
389     pseudo-atoms CH$_x$ and S based on a well-established and widely-used
390 skuang 3736 effective potential of Hautman and Klein\cite{hautman:4994} for the
391     Au(111) surface. As our simulations require the gold lattice slab to
392     be non-rigid so that it could accommodate kinetic energy for thermal
393 skuang 3730 transport study purpose, the pair-wise form of potentials is
394     preferred.
395 skuang 3721
396 skuang 3730 Besides, the potentials developed from {\it ab initio} calculations by
397     Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
398     interactions between Au and aromatic C/H atoms in toluene.[MORE DETAILS]
399 skuang 3725
400 skuang 3730 However, the Lennard-Jones parameters between Au and other types of
401     particles in our simulations are not yet well-established. For these
402     interactions, we attempt to derive their parameters using the Mixing
403     Rule. To do this, the ``Metal-non-Metal'' (MnM) interaction parameters
404     for Au is first extracted from the Au-CH$_x$ parameters by applying
405     the Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
406     parameters in our simulations.
407 skuang 3729
408 skuang 3730 \begin{table*}
409     \begin{minipage}{\linewidth}
410     \begin{center}
411 skuang 3738 \caption{Non-bonded interaction paramters for non-metal
412     particles and metal-non-metal interactions in our
413     simulations.}
414 skuang 3730
415 skuang 3738 \begin{tabular}{cccccc}
416 skuang 3730 \hline\hline
417 skuang 3738 Non-metal atom $I$ & $\sigma_{II}$ & $\epsilon_{II}$ & $q_I$ &
418     $\sigma_{AuI}$ & $\epsilon_{AuI}$ \\
419     (or pseudo-atom) & \AA & kcal/mol & & \AA & kcal/mol \\
420 skuang 3730 \hline
421 skuang 3738 CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
422     CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
423     CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
424     CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
425     S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
426     CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
427     CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
428     CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
429     HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
430     CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
431     HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
432 skuang 3730 \hline\hline
433     \end{tabular}
434     \label{MnM}
435     \end{center}
436     \end{minipage}
437     \end{table*}
438 skuang 3729
439    
440 skuang 3730 \section{Results and Discussions}
441     [MAY HAVE A BRIEF SUMMARY]
442     \subsection{How Simulation Parameters Affects $G$}
443     [MAY NOT PUT AT FIRST]
444     We have varied our protocol or other parameters of the simulations in
445     order to investigate how these factors would affect the measurement of
446     $G$'s. It turned out that while some of these parameters would not
447     affect the results substantially, some other changes to the
448     simulations would have a significant impact on the measurement
449     results.
450 skuang 3725
451 skuang 3730 In some of our simulations, we allowed $L_x$ and $L_y$ to change
452     during equilibrating the liquid phase. Due to the stiffness of the Au
453     slab, $L_x$ and $L_y$ would not change noticeably after
454     equilibration. Although $L_z$ could fluctuate $\sim$1\% after a system
455     is fully equilibrated in the NPT ensemble, this fluctuation, as well
456     as those comparably smaller to $L_x$ and $L_y$, would not be magnified
457     on the calculated $G$'s, as shown in Table \ref{AuThiolHexaneUA}. This
458     insensivity to $L_i$ fluctuations allows reliable measurement of $G$'s
459     without the necessity of extremely cautious equilibration process.
460 skuang 3725
461 skuang 3730 As stated in our computational details, the spacing filled with
462     solvent molecules can be chosen within a range. This allows some
463     change of solvent molecule numbers for the same Au-butanethiol
464     surfaces. We did this study on our Au-butanethiol/hexane
465     simulations. Nevertheless, the results obtained from systems of
466     different $N_{hexane}$ did not indicate that the measurement of $G$ is
467     susceptible to this parameter. For computational efficiency concern,
468     smaller system size would be preferable, given that the liquid phase
469     structure is not affected.
470    
471     Our NIVS algorithm allows change of unphysical thermal flux both in
472     direction and in quantity. This feature extends our investigation of
473     interfacial thermal conductance. However, the magnitude of this
474     thermal flux is not arbitary if one aims to obtain a stable and
475     reliable thermal gradient. A temperature profile would be
476     substantially affected by noise when $|J_z|$ has a much too low
477     magnitude; while an excessively large $|J_z|$ that overwhelms the
478     conductance capacity of the interface would prevent a thermal gradient
479     to reach a stablized steady state. NIVS has the advantage of allowing
480     $J$ to vary in a wide range such that the optimal flux range for $G$
481     measurement can generally be simulated by the algorithm. Within the
482     optimal range, we were able to study how $G$ would change according to
483     the thermal flux across the interface. For our simulations, we denote
484     $J_z$ to be positive when the physical thermal flux is from the liquid
485     to metal, and negative vice versa. The $G$'s measured under different
486     $J_z$ is listed in Table \ref{AuThiolHexaneUA} and [REF]. These
487     results do not suggest that $G$ is dependent on $J_z$ within this flux
488     range. The linear response of flux to thermal gradient simplifies our
489     investigations in that we can rely on $G$ measurement with only a
490     couple $J_z$'s and do not need to test a large series of fluxes.
491    
492     %ADD MORE TO TABLE
493 skuang 3725 \begin{table*}
494     \begin{minipage}{\linewidth}
495     \begin{center}
496     \caption{Computed interfacial thermal conductivity ($G$ and
497 skuang 3731 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
498     interfaces with UA model and different hexane molecule numbers
499     at different temperatures using a range of energy fluxes.}
500 skuang 3730
501 skuang 3738 \begin{tabular}{ccccccc}
502 skuang 3730 \hline\hline
503 skuang 3738 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
504     $J_z$ & $G$ & $G^\prime$ \\
505     (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
506 skuang 3730 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
507     \hline
508 skuang 3738 200 & 266 & No & 0.672 & -0.96 & 102() & 80.0() \\
509     & 200 & Yes & 0.694 & 1.92 & 129() & 87.3() \\
510     & & Yes & 0.672 & 1.93 & 131() & 77.5() \\
511    
512     & 166 & Yes & 0.679 & 0.97 & 115() & 69.3() \\
513     & & Yes & 0.679 & 1.94 & 125() & 87.1() \\
514    
515     250 & 200 & No & 0.560 & 0.96 & 81.8() & 67.0() \\
516    
517     & 166 & Yes & 0.570 & 0.98 & 79.0() & 62.9() \\
518    
519     & & No & 0.569 & 1.44 & 76.2() & 64.8() \\
520    
521 skuang 3730 \hline\hline
522     \end{tabular}
523     \label{AuThiolHexaneUA}
524     \end{center}
525     \end{minipage}
526     \end{table*}
527    
528     Furthermore, we also attempted to increase system average temperatures
529     to above 200K. These simulations are first equilibrated in the NPT
530     ensemble under normal pressure. As stated above, the TraPPE-UA model
531     for hexane tends to predict a lower boiling point. In our simulations,
532     hexane had diffculty to remain in liquid phase when NPT equilibration
533     temperature is higher than 250K. Additionally, the equilibrated liquid
534     hexane density under 250K becomes lower than experimental value. This
535     expanded liquid phase leads to lower contact between hexane and
536     butanethiol as well.[MAY NEED FIGURE] And this reduced contact would
537     probably be accountable for a lower interfacial thermal conductance,
538     as shown in Table \ref{AuThiolHexaneUA}.
539    
540     A similar study for TraPPE-UA toluene agrees with the above result as
541     well. Having a higher boiling point, toluene tends to remain liquid in
542     our simulations even equilibrated under 300K in NPT
543     ensembles. Furthermore, the expansion of the toluene liquid phase is
544     not as significant as that of the hexane. This prevents severe
545     decrease of liquid-capping agent contact and the results (Table
546     \ref{AuThiolToluene}) show only a slightly decreased interface
547     conductance. Therefore, solvent-capping agent contact should play an
548     important role in the thermal transport process across the interface
549     in that higher degree of contact could yield increased conductance.
550    
551 skuang 3738 [ADD ERROR ESTIMATE TO TABLE]
552 skuang 3730 \begin{table*}
553     \begin{minipage}{\linewidth}
554     \begin{center}
555     \caption{Computed interfacial thermal conductivity ($G$ and
556 skuang 3731 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
557     interface at different temperatures using a range of energy
558     fluxes.}
559 skuang 3725
560 skuang 3738 \begin{tabular}{ccccc}
561 skuang 3725 \hline\hline
562 skuang 3738 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
563     (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
564 skuang 3725 \hline
565 skuang 3738 200 & 0.933 & -1.86 & 180() & 135() \\
566     & & 2.15 & 204() & 113() \\
567     & & -3.93 & 175() & 114() \\
568     \hline
569     300 & 0.855 & -1.91 & 143() & 125() \\
570     & & -4.19 & 134() & 113() \\
571 skuang 3725 \hline\hline
572     \end{tabular}
573     \label{AuThiolToluene}
574     \end{center}
575     \end{minipage}
576     \end{table*}
577    
578 skuang 3730 Besides lower interfacial thermal conductance, surfaces in relatively
579     high temperatures are susceptible to reconstructions, when
580     butanethiols have a full coverage on the Au(111) surface. These
581     reconstructions include surface Au atoms migrated outward to the S
582     atom layer, and butanethiol molecules embedded into the original
583     surface Au layer. The driving force for this behavior is the strong
584     Au-S interactions in our simulations. And these reconstructions lead
585     to higher ratio of Au-S attraction and thus is energetically
586     favorable. Furthermore, this phenomenon agrees with experimental
587     results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
588     {\it et al.} had kept their Au(111) slab rigid so that their
589     simulations can reach 300K without surface reconstructions. Without
590     this practice, simulating 100\% thiol covered interfaces under higher
591     temperatures could hardly avoid surface reconstructions. However, our
592     measurement is based on assuming homogeneity on $x$ and $y$ dimensions
593     so that measurement of $T$ at particular $z$ would be an effective
594     average of the particles of the same type. Since surface
595     reconstructions could eliminate the original $x$ and $y$ dimensional
596     homogeneity, measurement of $G$ is more difficult to conduct under
597     higher temperatures. Therefore, most of our measurements are
598 skuang 3732 undertaken at $\langle T\rangle\sim$200K.
599 skuang 3725
600 skuang 3730 However, when the surface is not completely covered by butanethiols,
601     the simulated system is more resistent to the reconstruction
602     above. Our Au-butanethiol/toluene system did not see this phenomena
603 skuang 3738 even at $\langle T\rangle\sim$300K. The Au(111) surfaces have a 90\%
604     coverage of butanethiols and have empty three-fold sites. These empty
605     sites could help prevent surface reconstruction in that they provide
606     other means of capping agent relaxation. It is observed that
607     butanethiols can migrate to their neighbor empty sites during a
608     simulation. Therefore, we were able to obtain $G$'s for these
609     interfaces even at a relatively high temperature without being
610     affected by surface reconstructions.
611 skuang 3725
612 skuang 3730 \subsection{Influence of Capping Agent Coverage on $G$}
613     To investigate the influence of butanethiol coverage on interfacial
614     thermal conductance, a series of different coverage Au-butanethiol
615     surfaces is prepared and solvated with various organic
616     molecules. These systems are then equilibrated and their interfacial
617     thermal conductivity are measured with our NIVS algorithm. Table
618     \ref{tlnUhxnUhxnD} lists these results for direct comparison between
619 skuang 3731 different coverages of butanethiol. To study the isotope effect in
620     interfacial thermal conductance, deuterated UA-hexane is included as
621     well.
622 skuang 3730
623 skuang 3731 It turned out that with partial covered butanethiol on the Au(111)
624     surface, the derivative definition for $G$ (Eq. \ref{derivativeG}) has
625     difficulty to apply, due to the difficulty in locating the maximum of
626     change of $\lambda$. Instead, the discrete definition
627     (Eq. \ref{discreteG}) is easier to apply, as max($\Delta T$) can still
628     be well-defined. Therefore, $G$'s (not $G^\prime$) are used for this
629     section.
630 skuang 3725
631 skuang 3731 From Table \ref{tlnUhxnUhxnD}, one can see the significance of the
632     presence of capping agents. Even when a fraction of the Au(111)
633     surface sites are covered with butanethiols, the conductivity would
634     see an enhancement by at least a factor of 3. This indicates the
635     important role cappping agent is playing for thermal transport
636     phenomena on metal/organic solvent surfaces.
637 skuang 3725
638 skuang 3731 Interestingly, as one could observe from our results, the maximum
639     conductance enhancement (largest $G$) happens while the surfaces are
640     about 75\% covered with butanethiols. This again indicates that
641     solvent-capping agent contact has an important role of the thermal
642     transport process. Slightly lower butanethiol coverage allows small
643     gaps between butanethiols to form. And these gaps could be filled with
644     solvent molecules, which acts like ``heat conductors'' on the
645     surface. The higher degree of interaction between these solvent
646     molecules and capping agents increases the enhancement effect and thus
647     produces a higher $G$ than densely packed butanethiol arrays. However,
648     once this maximum conductance enhancement is reached, $G$ decreases
649     when butanethiol coverage continues to decrease. Each capping agent
650     molecule reaches its maximum capacity for thermal
651     conductance. Therefore, even higher solvent-capping agent contact
652     would not offset this effect. Eventually, when butanethiol coverage
653     continues to decrease, solvent-capping agent contact actually
654     decreases with the disappearing of butanethiol molecules. In this
655     case, $G$ decrease could not be offset but instead accelerated.
656 skuang 3725
657 skuang 3731 A comparison of the results obtained from differenet organic solvents
658     can also provide useful information of the interfacial thermal
659     transport process. The deuterated hexane (UA) results do not appear to
660     be much different from those of normal hexane (UA), given that
661     butanethiol (UA) is non-deuterated for both solvents. These UA model
662     studies, even though eliminating C-H vibration samplings, still have
663     C-C vibrational frequencies different from each other. However, these
664 skuang 3732 differences in the infrared range do not seem to produce an observable
665 skuang 3731 difference for the results of $G$. [MAY NEED FIGURE]
666 skuang 3730
667 skuang 3731 Furthermore, results for rigid body toluene solvent, as well as other
668     UA-hexane solvents, are reasonable within the general experimental
669     ranges[CITATIONS]. This suggests that explicit hydrogen might not be a
670     required factor for modeling thermal transport phenomena of systems
671     such as Au-thiol/organic solvent.
672    
673     However, results for Au-butanethiol/toluene do not show an identical
674     trend with those for Au-butanethiol/hexane in that $G$'s remain at
675     approximately the same magnitue when butanethiol coverage differs from
676     25\% to 75\%. This might be rooted in the molecule shape difference
677     for plane-like toluene and chain-like {\it n}-hexane. Due to this
678     difference, toluene molecules have more difficulty in occupying
679     relatively small gaps among capping agents when their coverage is not
680     too low. Therefore, the solvent-capping agent contact may keep
681     increasing until the capping agent coverage reaches a relatively low
682     level. This becomes an offset for decreasing butanethiol molecules on
683     its effect to the process of interfacial thermal transport. Thus, one
684     can see a plateau of $G$ vs. butanethiol coverage in our results.
685    
686 skuang 3738 [NEED ERROR ESTIMATE, CONVERT TO FIGURE]
687 skuang 3725 \begin{table*}
688     \begin{minipage}{\linewidth}
689     \begin{center}
690 skuang 3732 \caption{Computed interfacial thermal conductivity ($G$) values
691     for the Au-butanethiol/solvent interface with various UA
692     models and different capping agent coverages at $\langle
693     T\rangle\sim$200K using certain energy flux respectively.}
694 skuang 3725
695 skuang 3731 \begin{tabular}{cccc}
696 skuang 3725 \hline\hline
697 skuang 3732 Thiol & \multicolumn{3}{c}{$G$(MW/m$^2$/K)} \\
698     coverage (\%) & hexane & hexane(D) & toluene \\
699 skuang 3725 \hline
700 skuang 3732 0.0 & 46.5() & 43.9() & 70.1() \\
701     25.0 & 151() & 153() & 249() \\
702     50.0 & 172() & 182() & 214() \\
703     75.0 & 242() & 229() & 244() \\
704     88.9 & 178() & - & - \\
705     100.0 & 137() & 153() & 187() \\
706 skuang 3725 \hline\hline
707     \end{tabular}
708 skuang 3730 \label{tlnUhxnUhxnD}
709 skuang 3725 \end{center}
710     \end{minipage}
711     \end{table*}
712    
713 skuang 3730 \subsection{Influence of Chosen Molecule Model on $G$}
714     [MAY COMBINE W MECHANISM STUDY]
715    
716 skuang 3732 In addition to UA solvent/capping agent models, AA models are included
717     in our simulations as well. Besides simulations of the same (UA or AA)
718     model for solvent and capping agent, different models can be applied
719     to different components. Furthermore, regardless of models chosen,
720     either the solvent or the capping agent can be deuterated, similar to
721     the previous section. Table \ref{modelTest} summarizes the results of
722     these studies.
723 skuang 3725
724 skuang 3732 [MORE DATA; ERROR ESTIMATE]
725 skuang 3725 \begin{table*}
726     \begin{minipage}{\linewidth}
727     \begin{center}
728    
729     \caption{Computed interfacial thermal conductivity ($G$ and
730 skuang 3732 $G^\prime$) values for interfaces using various models for
731     solvent and capping agent (or without capping agent) at
732     $\langle T\rangle\sim$200K.}
733 skuang 3725
734 skuang 3732 \begin{tabular}{ccccc}
735 skuang 3725 \hline\hline
736 skuang 3732 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
737     (or bare surface) & model & (GW/m$^2$) &
738     \multicolumn{2}{c}{(MW/m$^2$/K)} \\
739 skuang 3725 \hline
740 skuang 3732 UA & AA hexane & 1.94 & 135() & 129() \\
741     & & 2.86 & 126() & 115() \\
742     & AA toluene & 1.89 & 200() & 149() \\
743     AA & UA hexane & 1.94 & 116() & 129() \\
744     & AA hexane & 3.76 & 451() & 378() \\
745     & & 4.71 & 432() & 334() \\
746     & AA toluene & 3.79 & 487() & 290() \\
747     AA(D) & UA hexane & 1.94 & 158() & 172() \\
748     bare & AA hexane & 0.96 & 31.0() & 29.4() \\
749 skuang 3725 \hline\hline
750     \end{tabular}
751 skuang 3732 \label{modelTest}
752 skuang 3725 \end{center}
753     \end{minipage}
754     \end{table*}
755    
756 skuang 3732 To facilitate direct comparison, the same system with differnt models
757     for different components uses the same length scale for their
758     simulation cells. Without the presence of capping agent, using
759     different models for hexane yields similar results for both $G$ and
760     $G^\prime$, and these two definitions agree with eath other very
761     well. This indicates very weak interaction between the metal and the
762     solvent, and is a typical case for acoustic impedance mismatch between
763     these two phases.
764 skuang 3730
765 skuang 3732 As for Au(111) surfaces completely covered by butanethiols, the choice
766     of models for capping agent and solvent could impact the measurement
767     of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
768     interfaces, using AA model for both butanethiol and hexane yields
769     substantially higher conductivity values than using UA model for at
770     least one component of the solvent and capping agent, which exceeds
771     the upper bond of experimental value range. This is probably due to
772     the classically treated C-H vibrations in the AA model, which should
773     not be appreciably populated at normal temperatures. In comparison,
774     once either the hexanes or the butanethiols are deuterated, one can
775     see a significantly lower $G$ and $G^\prime$. In either of these
776     cases, the C-H(D) vibrational overlap between the solvent and the
777     capping agent is removed. [MAY NEED FIGURE] Conclusively, the
778     improperly treated C-H vibration in the AA model produced
779     over-predicted results accordingly. Compared to the AA model, the UA
780     model yields more reasonable results with higher computational
781     efficiency.
782 skuang 3731
783 skuang 3732 However, for Au-butanethiol/toluene interfaces, having the AA
784     butanethiol deuterated did not yield a significant change in the
785     measurement results.
786     . , so extra degrees of freedom
787     such as the C-H vibrations could enhance heat exchange between these
788     two phases and result in a much higher conductivity.
789 skuang 3731
790 skuang 3732
791     Although the QSC model for Au is known to predict an overly low value
792 skuang 3738 for bulk metal gold conductivity\cite{kuang:164101}, our computational
793 skuang 3732 results for $G$ and $G^\prime$ do not seem to be affected by this
794     drawback of the model for metal. Instead, the modeling of interfacial
795     thermal transport behavior relies mainly on an accurate description of
796     the interactions between components occupying the interfaces.
797    
798 skuang 3730 \subsection{Mechanism of Interfacial Thermal Conductance Enhancement
799     by Capping Agent}
800 skuang 3732 %OR\subsection{Vibrational spectrum study on conductance mechanism}
801 skuang 3730
802 skuang 3732 [MAY INTRODUCE PROTOCOL IN METHODOLOGY/COMPUTATIONAL DETAIL, EQN'S]
803 skuang 3730
804 skuang 3725 To investigate the mechanism of this interfacial thermal conductance,
805     the vibrational spectra of various gold systems were obtained and are
806     shown as in the upper panel of Fig. \ref{vibration}. To obtain these
807     spectra, one first runs a simulation in the NVE ensemble and collects
808     snapshots of configurations; these configurations are used to compute
809     the velocity auto-correlation functions, which is used to construct a
810 skuang 3732 power spectrum via a Fourier transform.
811 skuang 3725
812 skuang 3732 The gold surfaces covered by
813     butanethiol molecules, compared to bare gold surfaces, exhibit an
814     additional peak observed at a frequency of $\sim$170cm$^{-1}$, which
815     is attributed to the vibration of the S-Au bond. This vibration
816     enables efficient thermal transport from surface Au atoms to the
817     capping agents. Simultaneously, as shown in the lower panel of
818     Fig. \ref{vibration}, the large overlap of the vibration spectra of
819     butanethiol and hexane in the all-atom model, including the C-H
820     vibration, also suggests high thermal exchange efficiency. The
821     combination of these two effects produces the drastic interfacial
822     thermal conductance enhancement in the all-atom model.
823    
824     [MAY NEED TO CONVERT TO JPEG]
825 skuang 3725 \begin{figure}
826     \includegraphics[width=\linewidth]{vibration}
827     \caption{Vibrational spectra obtained for gold in different
828     environments (upper panel) and for Au/thiol/hexane simulation in
829     all-atom model (lower panel).}
830     \label{vibration}
831     \end{figure}
832    
833 skuang 3732 [COMPARISON OF TWO G'S; AU SLAB WIDTHS; ETC]
834     % The results show that the two definitions used for $G$ yield
835     % comparable values, though $G^\prime$ tends to be smaller.
836    
837 skuang 3730 \section{Conclusions}
838 skuang 3732 The NIVS algorithm we developed has been applied to simulations of
839     Au-butanethiol surfaces with organic solvents. This algorithm allows
840     effective unphysical thermal flux transferred between the metal and
841     the liquid phase. With the flux applied, we were able to measure the
842     corresponding thermal gradient and to obtain interfacial thermal
843     conductivities. Our simulations have seen significant conductance
844     enhancement with the presence of capping agent, compared to the bare
845     gold/liquid interfaces. The acoustic impedance mismatch between the
846     metal and the liquid phase is effectively eliminated by proper capping
847     agent. Furthermore, the coverage precentage of the capping agent plays
848     an important role in the interfacial thermal transport process.
849 skuang 3725
850 skuang 3732 Our measurement results, particularly of the UA models, agree with
851     available experimental data. This indicates that our force field
852     parameters have a nice description of the interactions between the
853     particles at the interfaces. AA models tend to overestimate the
854     interfacial thermal conductance in that the classically treated C-H
855     vibration would be overly sampled. Compared to the AA models, the UA
856     models have higher computational efficiency with satisfactory
857     accuracy, and thus are preferable in interfacial thermal transport
858     modelings.
859 skuang 3730
860 skuang 3732 Vlugt {\it et al.} has investigated the surface thiol structures for
861     nanocrystal gold and pointed out that they differs from those of the
862     Au(111) surface\cite{vlugt:cpc2007154}. This difference might lead to
863     change of interfacial thermal transport behavior as well. To
864     investigate this problem, an effective means to introduce thermal flux
865     and measure the corresponding thermal gradient is desirable for
866     simulating structures with spherical symmetry.
867 skuang 3730
868 skuang 3732
869 gezelter 3717 \section{Acknowledgments}
870     Support for this project was provided by the National Science
871     Foundation under grant CHE-0848243. Computational time was provided by
872     the Center for Research Computing (CRC) at the University of Notre
873 skuang 3730 Dame. \newpage
874 gezelter 3717
875     \bibliography{interfacial}
876    
877     \end{doublespace}
878     \end{document}
879