ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3729
Committed: Wed Jun 29 18:14:23 2011 UTC (13 years ago) by skuang
Content type: application/x-tex
File size: 25043 byte(s)
Log Message:
add more data, citations, some work on computational details.

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     We have developed a Non-Isotropic Velocity Scaling algorithm for
49     setting up and maintaining stable thermal gradients in non-equilibrium
50     molecular dynamics simulations. This approach effectively imposes
51     unphysical thermal flux even between particles of different
52     identities, conserves linear momentum and kinetic energy, and
53     minimally perturbs the velocity profile of a system when compared with
54     previous RNEMD methods. We have used this method to simulate thermal
55     conductance at metal / organic solvent interfaces both with and
56     without the presence of thiol-based capping agents. We obtained
57     values comparable with experimental values, and observed significant
58     conductance enhancement with the presence of capping agents. Computed
59     power spectra indicate the acoustic impedance mismatch between metal
60     and liquid phase is greatly reduced by the capping agents and thus
61     leads to higher interfacial thermal transfer efficiency.
62    
63 gezelter 3717 \end{abstract}
64    
65     \newpage
66    
67     %\narrowtext
68    
69     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70     % BODY OF TEXT
71     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72    
73     \section{Introduction}
74 skuang 3727 [BACKGROUND FOR INTERFACIAL THERMAL CONDUCTANCE PROBLEM]
75 skuang 3725 Interfacial thermal conductance is extensively studied both
76     experimentally and computationally, and systems with interfaces
77     present are generally heterogeneous. Although interfaces are commonly
78     barriers to heat transfer, it has been
79     reported\cite{doi:10.1021/la904855s} that under specific circustances,
80     e.g. with certain capping agents present on the surface, interfacial
81     conductance can be significantly enhanced. However, heat conductance
82     of molecular and nano-scale interfaces will be affected by the
83     chemical details of the surface and is challenging to
84     experimentalist. The lower thermal flux through interfaces is even
85     more difficult to measure with EMD and forward NEMD simulation
86     methods. Therefore, developing good simulation methods will be
87     desirable in order to investigate thermal transport across interfaces.
88 gezelter 3717
89 skuang 3725 Recently, we have developed the Non-Isotropic Velocity Scaling (NIVS)
90     algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
91     retains the desirable features of RNEMD (conservation of linear
92     momentum and total energy, compatibility with periodic boundary
93     conditions) while establishing true thermal distributions in each of
94     the two slabs. Furthermore, it allows more effective thermal exchange
95     between particles of different identities, and thus enables extensive
96     study of interfacial conductance.
97    
98 skuang 3721 \section{Methodology}
99     \subsection{Algorithm}
100 skuang 3727 [BACKGROUND FOR MD METHODS]
101 skuang 3721 There have been many algorithms for computing thermal conductivity
102     using molecular dynamics simulations. However, interfacial conductance
103     is at least an order of magnitude smaller. This would make the
104     calculation even more difficult for those slowly-converging
105     equilibrium methods. Imposed-flux non-equilibrium
106     methods\cite{MullerPlathe:1997xw} have the flux set {\it a priori} and
107     the response of temperature or momentum gradients are easier to
108     measure than the flux, if unknown, and thus, is a preferable way to
109     the forward NEMD methods. Although the momentum swapping approach for
110     flux-imposing can be used for exchanging energy between particles of
111     different identity, the kinetic energy transfer efficiency is affected
112     by the mass difference between the particles, which limits its
113     application on heterogeneous interfacial systems.
114    
115     The non-isotropic velocity scaling (NIVS)\cite{kuang:164101} approach in
116     non-equilibrium MD simulations is able to impose relatively large
117     kinetic energy flux without obvious perturbation to the velocity
118     distribution of the simulated systems. Furthermore, this approach has
119     the advantage in heterogeneous interfaces in that kinetic energy flux
120     can be applied between regions of particles of arbitary identity, and
121     the flux quantity is not restricted by particle mass difference.
122    
123     The NIVS algorithm scales the velocity vectors in two separate regions
124     of a simulation system with respective diagonal scaling matricies. To
125     determine these scaling factors in the matricies, a set of equations
126     including linear momentum conservation and kinetic energy conservation
127     constraints and target momentum/energy flux satisfaction is
128     solved. With the scaling operation applied to the system in a set
129     frequency, corresponding momentum/temperature gradients can be built,
130     which can be used for computing transportation properties and other
131     applications related to momentum/temperature gradients. The NIVS
132     algorithm conserves momenta and energy and does not depend on an
133     external thermostat.
134    
135 skuang 3727 \subsection{Defining Interfacial Thermal Conductivity $G$}
136     For interfaces with a relatively low interfacial conductance, the bulk
137     regions on either side of an interface rapidly come to a state in
138     which the two phases have relatively homogeneous (but distinct)
139     temperatures. The interfacial thermal conductivity $G$ can therefore
140     be approximated as:
141     \begin{equation}
142     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
143     \langle T_\mathrm{cold}\rangle \right)}
144     \label{lowG}
145     \end{equation}
146     where ${E_{total}}$ is the imposed non-physical kinetic energy
147     transfer and ${\langle T_\mathrm{hot}\rangle}$ and ${\langle
148     T_\mathrm{cold}\rangle}$ are the average observed temperature of the
149     two separated phases.
150 skuang 3721
151 skuang 3727 When the interfacial conductance is {\it not} small, two ways can be
152     used to define $G$.
153    
154     One way is to assume the temperature is discretely different on two
155     sides of the interface, $G$ can be calculated with the thermal flux
156     applied $J$ and the maximum temperature difference measured along the
157     thermal gradient max($\Delta T$), which occurs at the interface, as:
158     \begin{equation}
159     G=\frac{J}{\Delta T}
160     \label{discreteG}
161     \end{equation}
162    
163     The other approach is to assume a continuous temperature profile along
164     the thermal gradient axis (e.g. $z$) and define $G$ at the point where
165     the magnitude of thermal conductivity $\lambda$ change reach its
166     maximum, given that $\lambda$ is well-defined throughout the space:
167     \begin{equation}
168     G^\prime = \Big|\frac{\partial\lambda}{\partial z}\Big|
169     = \Big|\frac{\partial}{\partial z}\left(-J_z\Big/
170     \left(\frac{\partial T}{\partial z}\right)\right)\Big|
171     = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
172     \Big/\left(\frac{\partial T}{\partial z}\right)^2
173     \label{derivativeG}
174     \end{equation}
175    
176     With the temperature profile obtained from simulations, one is able to
177     approximate the first and second derivatives of $T$ with finite
178     difference method and thus calculate $G^\prime$.
179    
180     In what follows, both definitions are used for calculation and comparison.
181    
182     [IMPOSE G DEFINITION INTO OUR SYSTEMS]
183     To facilitate the use of the above definitions in calculating $G$ and
184     $G^\prime$, we have a metal slab with its (111) surfaces perpendicular
185     to the $z$-axis of our simulation cells. With or withour capping
186     agents on the surfaces, the metal slab is solvated with organic
187     solvents, as illustrated in Figure \ref{demoPic}.
188    
189     \begin{figure}
190     \includegraphics[width=\linewidth]{demoPic}
191     \caption{A sample showing how a metal slab has its (111) surface
192     covered by capping agent molecules and solvated by hexane.}
193     \label{demoPic}
194     \end{figure}
195    
196     With a simulation cell setup following the above manner, one is able
197     to equilibrate the system and impose an unphysical thermal flux
198     between the liquid and the metal phase with the NIVS algorithm. Under
199     a stablized thermal gradient induced by periodically applying the
200     unphysical flux, one is able to obtain a temperature profile and the
201     physical thermal flux corresponding to it, which equals to the
202     unphysical flux applied by NIVS. These data enables the evaluation of
203     the interfacial thermal conductance of a surface. Figure \ref{gradT}
204     is an example how those stablized thermal gradient can be used to
205     obtain the 1st and 2nd derivatives of the temperature profile.
206    
207     \begin{figure}
208     \includegraphics[width=\linewidth]{gradT}
209     \caption{The 1st and 2nd derivatives of temperature profile can be
210     obtained with finite difference approximation.}
211     \label{gradT}
212     \end{figure}
213    
214     \section{Computational Details}
215     \subsection{System Geometry}
216     In our simulations, Au is used to construct a metal slab with bare
217     (111) surface perpendicular to the $z$-axis. Different slab thickness
218     (layer numbers of Au) are simulated. This metal slab is first
219     equilibrated under normal pressure (1 atm) and a desired
220     temperature. After equilibration, butanethiol is used as the capping
221     agent molecule to cover the bare Au (111) surfaces evenly. The sulfur
222     atoms in the butanethiol molecules would occupy the three-fold sites
223     of the surfaces, and the maximal butanethiol capacity on Au surface is
224     $1/3$ of the total number of surface Au atoms[CITATION]. A series of
225     different coverage surfaces is investigated in order to study the
226     relation between coverage and conductance.
227    
228     [COVERAGE DISCRIPTION] However, since the interactions between surface
229     Au and butanethiol is non-bonded, the capping agent molecules are
230     allowed to migrate to an empty neighbor three-fold site during a
231     simulation. Therefore, the initial configuration would not severely
232     affect the sampling of a variety of configurations of the same
233     coverage, and the final conductance measurement would be an average
234     effect of these configurations explored in the simulations. [MAY NEED FIGURES]
235    
236     After the modified Au-butanethiol surface systems are equilibrated
237     under canonical ensemble, Packmol\cite{packmol} is used to pack
238     organic solvent molecules in the previously vacuum part of the
239     simulation cells, which guarantees that short range repulsive
240     interactions do not disrupt the simulations. Two solvents are
241     investigated, one which has little vibrational overlap with the
242     alkanethiol and plane-like shape (toluene), and one which has similar
243     vibrational frequencies and chain-like shape ({\it n}-hexane). The
244 skuang 3728 spacing filled by solvent molecules, i.e. the gap between periodically
245     repeated Au-butanethiol surfaces should be carefully chosen so that it
246     would not be too short to affect the liquid phase structure, nor too
247     long, leading to over cooling (freezing) or heating (boiling) when a
248     thermal flux is applied. In our simulations, this spacing is usually
249     $35 \sim 60$\AA.
250 skuang 3727
251 skuang 3728 The initial configurations generated by Packmol are further
252     equilibrated with the $x$ and $y$ dimensions fixed, only allowing
253     length scale change in $z$ dimension. This is to ensure that the
254     equilibration of liquid phase does not affect the metal crystal
255     structure in $x$ and $y$ dimensions. Further equilibration are run
256     under NVT and then NVE ensembles.
257    
258 skuang 3727 After the systems reach equilibrium, NIVS is implemented to impose a
259     periodic unphysical thermal flux between the metal and the liquid
260 skuang 3728 phase. Most of our simulations are under an average temperature of
261     $\sim$200K. Therefore, this flux usually comes from the metal to the
262 skuang 3727 liquid so that the liquid has a higher temperature and would not
263     freeze due to excessively low temperature. This induced temperature
264     gradient is stablized and the simulation cell is devided evenly into
265     N slabs along the $z$-axis and the temperatures of each slab are
266     recorded. When the slab width $d$ of each slab is the same, the
267     derivatives of $T$ with respect to slab number $n$ can be directly
268     used for $G^\prime$ calculations:
269     \begin{equation}
270     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
271     \Big/\left(\frac{\partial T}{\partial z}\right)^2
272     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
273     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
274     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
275     \Big/\left(\frac{\partial T}{\partial n}\right)^2
276     \label{derivativeG2}
277     \end{equation}
278    
279 skuang 3725 \subsection{Force Field Parameters}
280 skuang 3728 Our simulations include various components. Therefore, force field
281     parameter descriptions are needed for interactions both between the
282     same type of particles and between particles of different species.
283 skuang 3721
284     The Au-Au interactions in metal lattice slab is described by the
285     quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
286     potentials include zero-point quantum corrections and are
287     reparametrized for accurate surface energies compared to the
288     Sutton-Chen potentials\cite{Chen90}.
289    
290 skuang 3728 For both solvent molecules, straight chain {\it n}-hexane and aromatic
291     toluene, United-Atom (UA) and All-Atom (AA) models are used
292     respectively. The TraPPE-UA
293     parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
294     for our UA solvent molecules. In these models, pseudo-atoms are
295     located at the carbon centers for alkyl groups. By eliminating
296     explicit hydrogen atoms, these models are simple and computationally
297 skuang 3729 efficient, while maintains good accuracy. However, the TraPPE-UA for
298     alkanes is known to predict a lower boiling point than experimental
299     values. Considering that after an unphysical thermal flux is applied
300     to a system, the temperature of ``hot'' area in the liquid phase would be
301     significantly higher than the average, to prevent over heating and
302     boiling of the liquid phase, the average temperature in our
303     simulations should be much lower than the liquid boiling point. [NEED MORE DISCUSSION]
304     For UA-toluene model, rigid body constraints are applied, so that the
305     benzene ring and the methyl-C(aromatic) bond are kept rigid. This
306     would save computational time.[MORE DETAILS NEEDED]
307 skuang 3721
308 skuang 3729 Besides the TraPPE-UA models, AA models for both organic solvents are
309     included in our studies as well. For hexane, the OPLS
310     all-atom\cite{OPLSAA} force field is used. [MORE DETAILS]
311     For toluene, the United Force Field developed by Rapp\'{e} {\it et
312     al.}\cite{doi:10.1021/ja00051a040} is adopted.[MORE DETAILS]
313 skuang 3728
314 skuang 3729 The capping agent in our simulations, the butanethiol molecules can
315     either use UA or AA model. The TraPPE-UA force fields includes
316     parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used in
317     our simulations corresponding to our TraPPE-UA models for solvent.
318     and All-Atom models [NEED CITATIONS]
319     However, the model choice (UA or AA) of capping agent can be different
320     from the solvent. Regardless of model choice, the force field
321     parameters for interactions between capping agent and solvent can be
322     derived using Lorentz-Berthelot Mixing Rule.
323 skuang 3721
324     To describe the interactions between metal Au and non-metal capping
325     agent and solvent, we refer to Vlugt\cite{vlugt:cpc2007154} and derive
326 skuang 3729 other interactions which are not yet finely parametrized. [can add
327 skuang 3721 hautman and klein's paper here and more discussion; need to put
328 skuang 3729 aromatic-metal interaction approximation here]\cite{doi:10.1021/jp034405s}
329 skuang 3721
330 skuang 3725 [TABULATED FORCE FIELD PARAMETERS NEEDED]
331    
332 skuang 3729
333     [SURFACE RECONSTRUCTION PREVENTS SIMULATION TEMP TO GO HIGHER]
334    
335    
336 skuang 3725 \section{Results}
337 skuang 3729 [REARRANGEMENT NEEDED]
338 skuang 3725 \subsection{Toluene Solvent}
339    
340 skuang 3727 The results (Table \ref{AuThiolToluene}) show a
341 skuang 3725 significant conductance enhancement compared to the gold/water
342     interface without capping agent and agree with available experimental
343     data. This indicates that the metal-metal potential, though not
344     predicting an accurate bulk metal thermal conductivity, does not
345     greatly interfere with the simulation of the thermal conductance
346     behavior across a non-metal interface. The solvent model is not
347     particularly volatile, so the simulation cell does not expand
348     significantly under higher temperature. We did not observe a
349     significant conductance decrease when the temperature was increased to
350     300K. The results show that the two definitions used for $G$ yield
351     comparable values, though $G^\prime$ tends to be smaller.
352    
353     \begin{table*}
354     \begin{minipage}{\linewidth}
355     \begin{center}
356     \caption{Computed interfacial thermal conductivity ($G$ and
357     $G^\prime$) values for the Au/butanethiol/toluene interface at
358     different temperatures using a range of energy fluxes.}
359    
360     \begin{tabular}{cccc}
361     \hline\hline
362     $\langle T\rangle$ & $J_z$ & $G$ & $G^\prime$ \\
363     (K) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
364     \hline
365     200 & 1.86 & 180 & 135 \\
366     & 2.15 & 204 & 113 \\
367     & 3.93 & 175 & 114 \\
368     300 & 1.91 & 143 & 125 \\
369     & 4.19 & 134 & 113 \\
370     \hline\hline
371     \end{tabular}
372     \label{AuThiolToluene}
373     \end{center}
374     \end{minipage}
375     \end{table*}
376    
377     \subsection{Hexane Solvent}
378    
379     Using the united-atom model, different coverages of capping agent,
380     temperatures of simulations and numbers of solvent molecules were all
381     investigated and Table \ref{AuThiolHexaneUA} shows the results of
382     these computations. The number of hexane molecules in our simulations
383     does not affect the calculations significantly. However, a very long
384     length scale for the thermal gradient axis ($z$) may cause excessively
385     hot or cold temperatures in the middle of the solvent region and lead
386     to undesired phenomena such as solvent boiling or freezing, while too
387     few solvent molecules would change the normal behavior of the liquid
388     phase. Our $N_{hexane}$ values were chosen to ensure that these
389     extreme cases did not happen to our simulations.
390    
391     Table \ref{AuThiolHexaneUA} enables direct comparison between
392     different coverages of capping agent, when other system parameters are
393     held constant. With high coverage of butanethiol on the gold surface,
394     the interfacial thermal conductance is enhanced
395     significantly. Interestingly, a slightly lower butanethiol coverage
396     leads to a moderately higher conductivity. This is probably due to
397     more solvent/capping agent contact when butanethiol molecules are
398     not densely packed, which enhances the interactions between the two
399     phases and lowers the thermal transfer barrier of this interface.
400     % [COMPARE TO AU/WATER IN PAPER]
401    
402     It is also noted that the overall simulation temperature is another
403     factor that affects the interfacial thermal conductance. One
404     possibility of this effect may be rooted in the decrease in density of
405     the liquid phase. We observed that when the average temperature
406     increases from 200K to 250K, the bulk hexane density becomes lower
407     than experimental value, as the system is equilibrated under NPT
408     ensemble. This leads to lower contact between solvent and capping
409     agent, and thus lower conductivity.
410    
411     Conductivity values are more difficult to obtain under higher
412     temperatures. This is because the Au surface tends to undergo
413     reconstructions in relatively high temperatures. Surface Au atoms can
414     migrate outward to reach higher Au-S contact; and capping agent
415     molecules can be embedded into the surface Au layer due to the same
416     driving force. This phenomenon agrees with experimental
417     results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. A surface
418     fully covered in capping agent is more susceptible to reconstruction,
419     possibly because fully coverage prevents other means of capping agent
420     relaxation, such as migration to an empty neighbor three-fold site.
421    
422     %MAY ADD MORE DATA TO TABLE
423     \begin{table*}
424     \begin{minipage}{\linewidth}
425     \begin{center}
426     \caption{Computed interfacial thermal conductivity ($G$ and
427     $G^\prime$) values for the Au/butanethiol/hexane interface
428     with united-atom model and different capping agent coverage
429     and solvent molecule numbers at different temperatures using a
430     range of energy fluxes.}
431    
432     \begin{tabular}{cccccc}
433     \hline\hline
434     Thiol & $\langle T\rangle$ & & $J_z$ & $G$ & $G^\prime$ \\
435     coverage (\%) & (K) & $N_{hexane}$ & (GW/m$^2$) &
436     \multicolumn{2}{c}{(MW/m$^2$/K)} \\
437     \hline
438     0.0 & 200 & 200 & 0.96 & 43.3 & 42.7 \\
439     & & & 1.91 & 45.7 & 42.9 \\
440     & & 166 & 0.96 & 43.1 & 53.4 \\
441     88.9 & 200 & 166 & 1.94 & 172 & 108 \\
442     100.0 & 250 & 200 & 0.96 & 81.8 & 67.0 \\
443     & & 166 & 0.98 & 79.0 & 62.9 \\
444     & & & 1.44 & 76.2 & 64.8 \\
445     & 200 & 200 & 1.92 & 129 & 87.3 \\
446     & & & 1.93 & 131 & 77.5 \\
447     & & 166 & 0.97 & 115 & 69.3 \\
448     & & & 1.94 & 125 & 87.1 \\
449     \hline\hline
450     \end{tabular}
451     \label{AuThiolHexaneUA}
452     \end{center}
453     \end{minipage}
454     \end{table*}
455    
456     For the all-atom model, the liquid hexane phase was not stable under NPT
457     conditions. Therefore, the simulation length scale parameters are
458     adopted from previous equilibration results of the united-atom model
459     at 200K. Table \ref{AuThiolHexaneAA} shows the results of these
460     simulations. The conductivity values calculated with full capping
461     agent coverage are substantially larger than observed in the
462     united-atom model, and is even higher than predicted by
463     experiments. It is possible that our parameters for metal-non-metal
464     particle interactions lead to an overestimate of the interfacial
465     thermal conductivity, although the active C-H vibrations in the
466     all-atom model (which should not be appreciably populated at normal
467     temperatures) could also account for this high conductivity. The major
468     thermal transfer barrier of Au/butanethiol/hexane interface is between
469     the liquid phase and the capping agent, so extra degrees of freedom
470     such as the C-H vibrations could enhance heat exchange between these
471     two phases and result in a much higher conductivity.
472    
473     \begin{table*}
474     \begin{minipage}{\linewidth}
475     \begin{center}
476    
477     \caption{Computed interfacial thermal conductivity ($G$ and
478     $G^\prime$) values for the Au/butanethiol/hexane interface
479     with all-atom model and different capping agent coverage at
480     200K using a range of energy fluxes.}
481    
482     \begin{tabular}{cccc}
483     \hline\hline
484     Thiol & $J_z$ & $G$ & $G^\prime$ \\
485     coverage (\%) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
486     \hline
487     0.0 & 0.95 & 28.5 & 27.2 \\
488     & 1.88 & 30.3 & 28.9 \\
489     100.0 & 2.87 & 551 & 294 \\
490     & 3.81 & 494 & 193 \\
491     \hline\hline
492     \end{tabular}
493     \label{AuThiolHexaneAA}
494     \end{center}
495     \end{minipage}
496     \end{table*}
497    
498     %subsubsection{Vibrational spectrum study on conductance mechanism}
499     To investigate the mechanism of this interfacial thermal conductance,
500     the vibrational spectra of various gold systems were obtained and are
501     shown as in the upper panel of Fig. \ref{vibration}. To obtain these
502     spectra, one first runs a simulation in the NVE ensemble and collects
503     snapshots of configurations; these configurations are used to compute
504     the velocity auto-correlation functions, which is used to construct a
505     power spectrum via a Fourier transform. The gold surfaces covered by
506     butanethiol molecules exhibit an additional peak observed at a
507     frequency of $\sim$170cm$^{-1}$, which is attributed to the vibration
508     of the S-Au bond. This vibration enables efficient thermal transport
509     from surface Au atoms to the capping agents. Simultaneously, as shown
510     in the lower panel of Fig. \ref{vibration}, the large overlap of the
511     vibration spectra of butanethiol and hexane in the all-atom model,
512     including the C-H vibration, also suggests high thermal exchange
513     efficiency. The combination of these two effects produces the drastic
514     interfacial thermal conductance enhancement in the all-atom model.
515    
516     \begin{figure}
517     \includegraphics[width=\linewidth]{vibration}
518     \caption{Vibrational spectra obtained for gold in different
519     environments (upper panel) and for Au/thiol/hexane simulation in
520     all-atom model (lower panel).}
521     \label{vibration}
522     \end{figure}
523     % 600dpi, letter size. too large?
524    
525    
526 gezelter 3717 \section{Acknowledgments}
527     Support for this project was provided by the National Science
528     Foundation under grant CHE-0848243. Computational time was provided by
529     the Center for Research Computing (CRC) at the University of Notre
530     Dame. \newpage
531    
532     \bibliography{interfacial}
533    
534     \end{doublespace}
535     \end{document}
536