ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/interfacial/interfacial.tex
Revision: 3737
Committed: Tue Jul 12 22:11:11 2011 UTC (12 years, 11 months ago) by skuang
Content type: application/x-tex
Original Path: interfacial/interfacial.tex
File size: 42231 byte(s)
Log Message:
revisions. add missing citations.

File Contents

# Content
1 \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 %\renewcommand\citemid{\ } % no comma in optional reference note
26 \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
48 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
65 \end{abstract}
66
67 \newpage
68
69 %\narrowtext
70
71 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72 % BODY OF TEXT
73 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74
75 \section{Introduction}
76 Interfacial thermal conductance is extensively studied both
77 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 materials. Furthermore, these materials are generally heterogeneous,
84 which challenges traditional research methods for homogeneous
85 systems.
86
87 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 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 monolayer of alkylthiolate with relatively long chain (8-20 carbon
108 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 The relatively low thermal flux through interfaces is
112 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 methods being able to effectively apply an unphysical flux in
117 non-homogeneous systems.
118
119 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 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
128 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 underlying mechanism for this phenomena was investigated.
135
136 [MAY ADD WHY STUDY AU-THIOL SURFACE; CITE SHAOYI JIANG]
137
138 \section{Methodology}
139 \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 methods\cite{MullerPlathe:1997xw} have the flux set {\it a priori} and
149 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
156 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 the advantage in heterogeneous interfaces in that kinetic energy flux
161 can be applied between regions of particles of arbitary identity, and
162 the flux will not be restricted by difference in particle mass.
163
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 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
174 \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
190 When the interfacial conductance is {\it not} small, there are two
191 ways to define $G$.
192
193 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 \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 difference methods and thus calculate $G^\prime$.
219
220 In what follows, both definitions have been used for calculation and
221 are compared in the results.
222
223 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
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 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
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 \subsection{Simulation Protocol}
254 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
266 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
274 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
281 The space filled by solvent molecules, i.e. the gap between
282 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 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 After the systems reach equilibrium, NIVS is implemented to impose a
300 periodic unphysical thermal flux between the metal and the liquid
301 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 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 \subsection{Force Field Parameters}
321 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
325 The Au-Au interactions in metal lattice slab is described by the
326 quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59.3527}. The QSC
327 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 Figure \ref{demoMol} demonstrates how we name our pseudo-atoms of the
332 organic solvent molecules in our simulations.
333
334 \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 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 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 simulations should be much lower than the liquid boiling point. [MORE DISCUSSION]
355 For UA-toluene model, rigid body constraints are applied, so that the
356 benzene ring and the methyl-CRar bond are kept rigid. This would save
357 computational time.[MORE DETAILS]
358
359 Besides the TraPPE-UA models, AA models for both organic solvents are
360 included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
361 force field is used. [MORE DETAILS]
362 For toluene, the United Force Field developed by Rapp\'{e} {\it et
363 al.}\cite{doi:10.1021/ja00051a040} is adopted.[MORE DETAILS]
364
365 The capping agent in our simulations, the butanethiol molecules can
366 either use UA or AA model. The TraPPE-UA force fields includes
367 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 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 Lorentz-Berthelot Mixing Rule:[EQN'S]
379
380
381 To describe the interactions between metal Au and non-metal capping
382 agent and solvent particles, we refer to an adsorption study of alkyl
383 thiols on gold surfaces by Vlugt {\it et
384 al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
385 form of potential parameters for the interaction between Au and
386 pseudo-atoms CH$_x$ and S based on a well-established and widely-used
387 effective potential of Hautman and Klein\cite{hautman:4994} for the
388 Au(111) surface. As our simulations require the gold lattice slab to
389 be non-rigid so that it could accommodate kinetic energy for thermal
390 transport study purpose, the pair-wise form of potentials is
391 preferred.
392
393 Besides, the potentials developed from {\it ab initio} calculations by
394 Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
395 interactions between Au and aromatic C/H atoms in toluene.[MORE DETAILS]
396
397 However, the Lennard-Jones parameters between Au and other types of
398 particles in our simulations are not yet well-established. For these
399 interactions, we attempt to derive their parameters using the Mixing
400 Rule. To do this, the ``Metal-non-Metal'' (MnM) interaction parameters
401 for Au is first extracted from the Au-CH$_x$ parameters by applying
402 the Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
403 parameters in our simulations.
404
405 \begin{table*}
406 \begin{minipage}{\linewidth}
407 \begin{center}
408 \caption{Lennard-Jones parameters for Au-non-Metal
409 interactions in our simulations.}
410
411 \begin{tabular}{ccc}
412 \hline\hline
413 Non-metal atom & $\sigma$ & $\epsilon$ \\
414 (or pseudo-atom) & \AA & kcal/mol \\
415 \hline
416 S & 2.40 & 8.465 \\
417 CH3 & 3.54 & 0.2146 \\
418 CH2 & 3.54 & 0.1749 \\
419 CT3 & 3.365 & 0.1373 \\
420 CT2 & 3.365 & 0.1373 \\
421 CTT & 3.365 & 0.1373 \\
422 HC & 2.865 & 0.09256 \\
423 CHar & 3.4625 & 0.1680 \\
424 CRar & 3.555 & 0.1604 \\
425 CA & 3.173 & 0.0640 \\
426 HA & 2.746 & 0.0414 \\
427 \hline\hline
428 \end{tabular}
429 \label{MnM}
430 \end{center}
431 \end{minipage}
432 \end{table*}
433
434
435 \section{Results and Discussions}
436 [MAY HAVE A BRIEF SUMMARY]
437 \subsection{How Simulation Parameters Affects $G$}
438 [MAY NOT PUT AT FIRST]
439 We have varied our protocol or other parameters of the simulations in
440 order to investigate how these factors would affect the measurement of
441 $G$'s. It turned out that while some of these parameters would not
442 affect the results substantially, some other changes to the
443 simulations would have a significant impact on the measurement
444 results.
445
446 In some of our simulations, we allowed $L_x$ and $L_y$ to change
447 during equilibrating the liquid phase. Due to the stiffness of the Au
448 slab, $L_x$ and $L_y$ would not change noticeably after
449 equilibration. Although $L_z$ could fluctuate $\sim$1\% after a system
450 is fully equilibrated in the NPT ensemble, this fluctuation, as well
451 as those comparably smaller to $L_x$ and $L_y$, would not be magnified
452 on the calculated $G$'s, as shown in Table \ref{AuThiolHexaneUA}. This
453 insensivity to $L_i$ fluctuations allows reliable measurement of $G$'s
454 without the necessity of extremely cautious equilibration process.
455
456 As stated in our computational details, the spacing filled with
457 solvent molecules can be chosen within a range. This allows some
458 change of solvent molecule numbers for the same Au-butanethiol
459 surfaces. We did this study on our Au-butanethiol/hexane
460 simulations. Nevertheless, the results obtained from systems of
461 different $N_{hexane}$ did not indicate that the measurement of $G$ is
462 susceptible to this parameter. For computational efficiency concern,
463 smaller system size would be preferable, given that the liquid phase
464 structure is not affected.
465
466 Our NIVS algorithm allows change of unphysical thermal flux both in
467 direction and in quantity. This feature extends our investigation of
468 interfacial thermal conductance. However, the magnitude of this
469 thermal flux is not arbitary if one aims to obtain a stable and
470 reliable thermal gradient. A temperature profile would be
471 substantially affected by noise when $|J_z|$ has a much too low
472 magnitude; while an excessively large $|J_z|$ that overwhelms the
473 conductance capacity of the interface would prevent a thermal gradient
474 to reach a stablized steady state. NIVS has the advantage of allowing
475 $J$ to vary in a wide range such that the optimal flux range for $G$
476 measurement can generally be simulated by the algorithm. Within the
477 optimal range, we were able to study how $G$ would change according to
478 the thermal flux across the interface. For our simulations, we denote
479 $J_z$ to be positive when the physical thermal flux is from the liquid
480 to metal, and negative vice versa. The $G$'s measured under different
481 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and [REF]. These
482 results do not suggest that $G$ is dependent on $J_z$ within this flux
483 range. The linear response of flux to thermal gradient simplifies our
484 investigations in that we can rely on $G$ measurement with only a
485 couple $J_z$'s and do not need to test a large series of fluxes.
486
487 %ADD MORE TO TABLE
488 \begin{table*}
489 \begin{minipage}{\linewidth}
490 \begin{center}
491 \caption{Computed interfacial thermal conductivity ($G$ and
492 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
493 interfaces with UA model and different hexane molecule numbers
494 at different temperatures using a range of energy fluxes.}
495
496 \begin{tabular}{cccccccc}
497 \hline\hline
498 $\langle T\rangle$ & & $L_x$ & $L_y$ & $L_z$ & $J_z$ &
499 $G$ & $G^\prime$ \\
500 (K) & $N_{hexane}$ & \multicolumn{3}{c}{(\AA)} & (GW/m$^2$) &
501 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
502 \hline
503 200 & 266 & 29.86 & 25.80 & 113.1 & -0.96 &
504 102() & 80.0() \\
505 & 200 & 29.84 & 25.81 & 93.9 & 1.92 &
506 129() & 87.3() \\
507 & & 29.84 & 25.81 & 95.3 & 1.93 &
508 131() & 77.5() \\
509 & 166 & 29.84 & 25.81 & 85.7 & 0.97 &
510 115() & 69.3() \\
511 & & & & & 1.94 &
512 125() & 87.1() \\
513 250 & 200 & 29.84 & 25.87 & 106.8 & 0.96 &
514 81.8() & 67.0() \\
515 & 166 & 29.87 & 25.84 & 94.8 & 0.98 &
516 79.0() & 62.9() \\
517 & & 29.84 & 25.85 & 95.0 & 1.44 &
518 76.2() & 64.8() \\
519 \hline\hline
520 \end{tabular}
521 \label{AuThiolHexaneUA}
522 \end{center}
523 \end{minipage}
524 \end{table*}
525
526 Furthermore, we also attempted to increase system average temperatures
527 to above 200K. These simulations are first equilibrated in the NPT
528 ensemble under normal pressure. As stated above, the TraPPE-UA model
529 for hexane tends to predict a lower boiling point. In our simulations,
530 hexane had diffculty to remain in liquid phase when NPT equilibration
531 temperature is higher than 250K. Additionally, the equilibrated liquid
532 hexane density under 250K becomes lower than experimental value. This
533 expanded liquid phase leads to lower contact between hexane and
534 butanethiol as well.[MAY NEED FIGURE] And this reduced contact would
535 probably be accountable for a lower interfacial thermal conductance,
536 as shown in Table \ref{AuThiolHexaneUA}.
537
538 A similar study for TraPPE-UA toluene agrees with the above result as
539 well. Having a higher boiling point, toluene tends to remain liquid in
540 our simulations even equilibrated under 300K in NPT
541 ensembles. Furthermore, the expansion of the toluene liquid phase is
542 not as significant as that of the hexane. This prevents severe
543 decrease of liquid-capping agent contact and the results (Table
544 \ref{AuThiolToluene}) show only a slightly decreased interface
545 conductance. Therefore, solvent-capping agent contact should play an
546 important role in the thermal transport process across the interface
547 in that higher degree of contact could yield increased conductance.
548
549 [ADD Lxyz AND ERROR ESTIMATE TO TABLE]
550 \begin{table*}
551 \begin{minipage}{\linewidth}
552 \begin{center}
553 \caption{Computed interfacial thermal conductivity ($G$ and
554 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
555 interface at different temperatures using a range of energy
556 fluxes.}
557
558 \begin{tabular}{cccc}
559 \hline\hline
560 $\langle T\rangle$ & $J_z$ & $G$ & $G^\prime$ \\
561 (K) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
562 \hline
563 200 & -1.86 & 180() & 135() \\
564 & 2.15 & 204() & 113() \\
565 & -3.93 & 175() & 114() \\
566 300 & -1.91 & 143() & 125() \\
567 & -4.19 & 134() & 113() \\
568 \hline\hline
569 \end{tabular}
570 \label{AuThiolToluene}
571 \end{center}
572 \end{minipage}
573 \end{table*}
574
575 Besides lower interfacial thermal conductance, surfaces in relatively
576 high temperatures are susceptible to reconstructions, when
577 butanethiols have a full coverage on the Au(111) surface. These
578 reconstructions include surface Au atoms migrated outward to the S
579 atom layer, and butanethiol molecules embedded into the original
580 surface Au layer. The driving force for this behavior is the strong
581 Au-S interactions in our simulations. And these reconstructions lead
582 to higher ratio of Au-S attraction and thus is energetically
583 favorable. Furthermore, this phenomenon agrees with experimental
584 results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
585 {\it et al.} had kept their Au(111) slab rigid so that their
586 simulations can reach 300K without surface reconstructions. Without
587 this practice, simulating 100\% thiol covered interfaces under higher
588 temperatures could hardly avoid surface reconstructions. However, our
589 measurement is based on assuming homogeneity on $x$ and $y$ dimensions
590 so that measurement of $T$ at particular $z$ would be an effective
591 average of the particles of the same type. Since surface
592 reconstructions could eliminate the original $x$ and $y$ dimensional
593 homogeneity, measurement of $G$ is more difficult to conduct under
594 higher temperatures. Therefore, most of our measurements are
595 undertaken at $\langle T\rangle\sim$200K.
596
597 However, when the surface is not completely covered by butanethiols,
598 the simulated system is more resistent to the reconstruction
599 above. Our Au-butanethiol/toluene system did not see this phenomena
600 even at $\langle T\rangle\sim$300K. The Au(111) surfaces have a 90\% coverage of
601 butanethiols and have empty three-fold sites. These empty sites could
602 help prevent surface reconstruction in that they provide other means
603 of capping agent relaxation. It is observed that butanethiols can
604 migrate to their neighbor empty sites during a simulation. Therefore,
605 we were able to obtain $G$'s for these interfaces even at a relatively
606 high temperature without being affected by surface reconstructions.
607
608 \subsection{Influence of Capping Agent Coverage on $G$}
609 To investigate the influence of butanethiol coverage on interfacial
610 thermal conductance, a series of different coverage Au-butanethiol
611 surfaces is prepared and solvated with various organic
612 molecules. These systems are then equilibrated and their interfacial
613 thermal conductivity are measured with our NIVS algorithm. Table
614 \ref{tlnUhxnUhxnD} lists these results for direct comparison between
615 different coverages of butanethiol. To study the isotope effect in
616 interfacial thermal conductance, deuterated UA-hexane is included as
617 well.
618
619 It turned out that with partial covered butanethiol on the Au(111)
620 surface, the derivative definition for $G$ (Eq. \ref{derivativeG}) has
621 difficulty to apply, due to the difficulty in locating the maximum of
622 change of $\lambda$. Instead, the discrete definition
623 (Eq. \ref{discreteG}) is easier to apply, as max($\Delta T$) can still
624 be well-defined. Therefore, $G$'s (not $G^\prime$) are used for this
625 section.
626
627 From Table \ref{tlnUhxnUhxnD}, one can see the significance of the
628 presence of capping agents. Even when a fraction of the Au(111)
629 surface sites are covered with butanethiols, the conductivity would
630 see an enhancement by at least a factor of 3. This indicates the
631 important role cappping agent is playing for thermal transport
632 phenomena on metal/organic solvent surfaces.
633
634 Interestingly, as one could observe from our results, the maximum
635 conductance enhancement (largest $G$) happens while the surfaces are
636 about 75\% covered with butanethiols. This again indicates that
637 solvent-capping agent contact has an important role of the thermal
638 transport process. Slightly lower butanethiol coverage allows small
639 gaps between butanethiols to form. And these gaps could be filled with
640 solvent molecules, which acts like ``heat conductors'' on the
641 surface. The higher degree of interaction between these solvent
642 molecules and capping agents increases the enhancement effect and thus
643 produces a higher $G$ than densely packed butanethiol arrays. However,
644 once this maximum conductance enhancement is reached, $G$ decreases
645 when butanethiol coverage continues to decrease. Each capping agent
646 molecule reaches its maximum capacity for thermal
647 conductance. Therefore, even higher solvent-capping agent contact
648 would not offset this effect. Eventually, when butanethiol coverage
649 continues to decrease, solvent-capping agent contact actually
650 decreases with the disappearing of butanethiol molecules. In this
651 case, $G$ decrease could not be offset but instead accelerated.
652
653 A comparison of the results obtained from differenet organic solvents
654 can also provide useful information of the interfacial thermal
655 transport process. The deuterated hexane (UA) results do not appear to
656 be much different from those of normal hexane (UA), given that
657 butanethiol (UA) is non-deuterated for both solvents. These UA model
658 studies, even though eliminating C-H vibration samplings, still have
659 C-C vibrational frequencies different from each other. However, these
660 differences in the infrared range do not seem to produce an observable
661 difference for the results of $G$. [MAY NEED FIGURE]
662
663 Furthermore, results for rigid body toluene solvent, as well as other
664 UA-hexane solvents, are reasonable within the general experimental
665 ranges[CITATIONS]. This suggests that explicit hydrogen might not be a
666 required factor for modeling thermal transport phenomena of systems
667 such as Au-thiol/organic solvent.
668
669 However, results for Au-butanethiol/toluene do not show an identical
670 trend with those for Au-butanethiol/hexane in that $G$'s remain at
671 approximately the same magnitue when butanethiol coverage differs from
672 25\% to 75\%. This might be rooted in the molecule shape difference
673 for plane-like toluene and chain-like {\it n}-hexane. Due to this
674 difference, toluene molecules have more difficulty in occupying
675 relatively small gaps among capping agents when their coverage is not
676 too low. Therefore, the solvent-capping agent contact may keep
677 increasing until the capping agent coverage reaches a relatively low
678 level. This becomes an offset for decreasing butanethiol molecules on
679 its effect to the process of interfacial thermal transport. Thus, one
680 can see a plateau of $G$ vs. butanethiol coverage in our results.
681
682 [NEED ERROR ESTIMATE, MAY ALSO PUT J HERE]
683 \begin{table*}
684 \begin{minipage}{\linewidth}
685 \begin{center}
686 \caption{Computed interfacial thermal conductivity ($G$) values
687 for the Au-butanethiol/solvent interface with various UA
688 models and different capping agent coverages at $\langle
689 T\rangle\sim$200K using certain energy flux respectively.}
690
691 \begin{tabular}{cccc}
692 \hline\hline
693 Thiol & \multicolumn{3}{c}{$G$(MW/m$^2$/K)} \\
694 coverage (\%) & hexane & hexane(D) & toluene \\
695 \hline
696 0.0 & 46.5() & 43.9() & 70.1() \\
697 25.0 & 151() & 153() & 249() \\
698 50.0 & 172() & 182() & 214() \\
699 75.0 & 242() & 229() & 244() \\
700 88.9 & 178() & - & - \\
701 100.0 & 137() & 153() & 187() \\
702 \hline\hline
703 \end{tabular}
704 \label{tlnUhxnUhxnD}
705 \end{center}
706 \end{minipage}
707 \end{table*}
708
709 \subsection{Influence of Chosen Molecule Model on $G$}
710 [MAY COMBINE W MECHANISM STUDY]
711
712 In addition to UA solvent/capping agent models, AA models are included
713 in our simulations as well. Besides simulations of the same (UA or AA)
714 model for solvent and capping agent, different models can be applied
715 to different components. Furthermore, regardless of models chosen,
716 either the solvent or the capping agent can be deuterated, similar to
717 the previous section. Table \ref{modelTest} summarizes the results of
718 these studies.
719
720 [MORE DATA; ERROR ESTIMATE]
721 \begin{table*}
722 \begin{minipage}{\linewidth}
723 \begin{center}
724
725 \caption{Computed interfacial thermal conductivity ($G$ and
726 $G^\prime$) values for interfaces using various models for
727 solvent and capping agent (or without capping agent) at
728 $\langle T\rangle\sim$200K.}
729
730 \begin{tabular}{ccccc}
731 \hline\hline
732 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
733 (or bare surface) & model & (GW/m$^2$) &
734 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
735 \hline
736 UA & AA hexane & 1.94 & 135() & 129() \\
737 & & 2.86 & 126() & 115() \\
738 & AA toluene & 1.89 & 200() & 149() \\
739 AA & UA hexane & 1.94 & 116() & 129() \\
740 & AA hexane & 3.76 & 451() & 378() \\
741 & & 4.71 & 432() & 334() \\
742 & AA toluene & 3.79 & 487() & 290() \\
743 AA(D) & UA hexane & 1.94 & 158() & 172() \\
744 bare & AA hexane & 0.96 & 31.0() & 29.4() \\
745 \hline\hline
746 \end{tabular}
747 \label{modelTest}
748 \end{center}
749 \end{minipage}
750 \end{table*}
751
752 To facilitate direct comparison, the same system with differnt models
753 for different components uses the same length scale for their
754 simulation cells. Without the presence of capping agent, using
755 different models for hexane yields similar results for both $G$ and
756 $G^\prime$, and these two definitions agree with eath other very
757 well. This indicates very weak interaction between the metal and the
758 solvent, and is a typical case for acoustic impedance mismatch between
759 these two phases.
760
761 As for Au(111) surfaces completely covered by butanethiols, the choice
762 of models for capping agent and solvent could impact the measurement
763 of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
764 interfaces, using AA model for both butanethiol and hexane yields
765 substantially higher conductivity values than using UA model for at
766 least one component of the solvent and capping agent, which exceeds
767 the upper bond of experimental value range. This is probably due to
768 the classically treated C-H vibrations in the AA model, which should
769 not be appreciably populated at normal temperatures. In comparison,
770 once either the hexanes or the butanethiols are deuterated, one can
771 see a significantly lower $G$ and $G^\prime$. In either of these
772 cases, the C-H(D) vibrational overlap between the solvent and the
773 capping agent is removed. [MAY NEED FIGURE] Conclusively, the
774 improperly treated C-H vibration in the AA model produced
775 over-predicted results accordingly. Compared to the AA model, the UA
776 model yields more reasonable results with higher computational
777 efficiency.
778
779 However, for Au-butanethiol/toluene interfaces, having the AA
780 butanethiol deuterated did not yield a significant change in the
781 measurement results.
782 . , so extra degrees of freedom
783 such as the C-H vibrations could enhance heat exchange between these
784 two phases and result in a much higher conductivity.
785
786
787 Although the QSC model for Au is known to predict an overly low value
788 for bulk metal gold conductivity[CITE NIVSRNEMD], our computational
789 results for $G$ and $G^\prime$ do not seem to be affected by this
790 drawback of the model for metal. Instead, the modeling of interfacial
791 thermal transport behavior relies mainly on an accurate description of
792 the interactions between components occupying the interfaces.
793
794 \subsection{Mechanism of Interfacial Thermal Conductance Enhancement
795 by Capping Agent}
796 %OR\subsection{Vibrational spectrum study on conductance mechanism}
797
798 [MAY INTRODUCE PROTOCOL IN METHODOLOGY/COMPUTATIONAL DETAIL, EQN'S]
799
800 To investigate the mechanism of this interfacial thermal conductance,
801 the vibrational spectra of various gold systems were obtained and are
802 shown as in the upper panel of Fig. \ref{vibration}. To obtain these
803 spectra, one first runs a simulation in the NVE ensemble and collects
804 snapshots of configurations; these configurations are used to compute
805 the velocity auto-correlation functions, which is used to construct a
806 power spectrum via a Fourier transform.
807
808 The gold surfaces covered by
809 butanethiol molecules, compared to bare gold surfaces, exhibit an
810 additional peak observed at a frequency of $\sim$170cm$^{-1}$, which
811 is attributed to the vibration of the S-Au bond. This vibration
812 enables efficient thermal transport from surface Au atoms to the
813 capping agents. Simultaneously, as shown in the lower panel of
814 Fig. \ref{vibration}, the large overlap of the vibration spectra of
815 butanethiol and hexane in the all-atom model, including the C-H
816 vibration, also suggests high thermal exchange efficiency. The
817 combination of these two effects produces the drastic interfacial
818 thermal conductance enhancement in the all-atom model.
819
820 [MAY NEED TO CONVERT TO JPEG]
821 \begin{figure}
822 \includegraphics[width=\linewidth]{vibration}
823 \caption{Vibrational spectra obtained for gold in different
824 environments (upper panel) and for Au/thiol/hexane simulation in
825 all-atom model (lower panel).}
826 \label{vibration}
827 \end{figure}
828
829 [COMPARISON OF TWO G'S; AU SLAB WIDTHS; ETC]
830 % The results show that the two definitions used for $G$ yield
831 % comparable values, though $G^\prime$ tends to be smaller.
832
833 \section{Conclusions}
834 The NIVS algorithm we developed has been applied to simulations of
835 Au-butanethiol surfaces with organic solvents. This algorithm allows
836 effective unphysical thermal flux transferred between the metal and
837 the liquid phase. With the flux applied, we were able to measure the
838 corresponding thermal gradient and to obtain interfacial thermal
839 conductivities. Our simulations have seen significant conductance
840 enhancement with the presence of capping agent, compared to the bare
841 gold/liquid interfaces. The acoustic impedance mismatch between the
842 metal and the liquid phase is effectively eliminated by proper capping
843 agent. Furthermore, the coverage precentage of the capping agent plays
844 an important role in the interfacial thermal transport process.
845
846 Our measurement results, particularly of the UA models, agree with
847 available experimental data. This indicates that our force field
848 parameters have a nice description of the interactions between the
849 particles at the interfaces. AA models tend to overestimate the
850 interfacial thermal conductance in that the classically treated C-H
851 vibration would be overly sampled. Compared to the AA models, the UA
852 models have higher computational efficiency with satisfactory
853 accuracy, and thus are preferable in interfacial thermal transport
854 modelings.
855
856 Vlugt {\it et al.} has investigated the surface thiol structures for
857 nanocrystal gold and pointed out that they differs from those of the
858 Au(111) surface\cite{vlugt:cpc2007154}. This difference might lead to
859 change of interfacial thermal transport behavior as well. To
860 investigate this problem, an effective means to introduce thermal flux
861 and measure the corresponding thermal gradient is desirable for
862 simulating structures with spherical symmetry.
863
864
865 \section{Acknowledgments}
866 Support for this project was provided by the National Science
867 Foundation under grant CHE-0848243. Computational time was provided by
868 the Center for Research Computing (CRC) at the University of Notre
869 Dame. \newpage
870
871 \bibliography{interfacial}
872
873 \end{doublespace}
874 \end{document}
875