ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3740
Committed: Thu Jul 14 21:36:07 2011 UTC (13 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 44556 byte(s)
Log Message:
Added two figures

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{[}{]}{,}{n}{}{;}
27 \bibliographystyle{achemso}
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 alkylthiol 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 comparatively 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]{method}
231 \caption{Interfacial conductance can be calculated by applying an
232 (unphysical) kinetic energy flux between two slabs, one located
233 within the metal and another on the edge of the periodic box. The
234 system responds by forming a thermal response or a gradient. In
235 bulk liquids, this gradient typically has a single slope, but in
236 interfacial systems, there are distinct thermal conductivity
237 domains. The interfacial conductance, $G$ is found by measuring the
238 temperature gap at the Gibbs dividing surface, or by using second
239 derivatives of the thermal profile.}
240 \label{demoPic}
241 \end{figure}
242
243 With the simulation cell described above, we are able to equilibrate
244 the system and impose an unphysical thermal flux between the liquid
245 and the metal phase using the NIVS algorithm. By periodically applying
246 the unphysical flux, we are able to obtain a temperature profile and
247 its spatial derivatives. These quantities enable the evaluation of the
248 interfacial thermal conductance of a surface. Figure \ref{gradT} is an
249 example how those applied thermal fluxes can be used to obtain the 1st
250 and 2nd derivatives of the temperature profile.
251
252 \begin{figure}
253 \includegraphics[width=\linewidth]{gradT}
254 \caption{The 1st and 2nd derivatives of temperature profile can be
255 obtained with finite difference approximation.}
256 \label{gradT}
257 \end{figure}
258
259 \section{Computational Details}
260 \subsection{Simulation Protocol}
261 The NIVS algorithm has been implemented in our MD simulation code,
262 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our
263 simulations. Different slab thickness (layer numbers of Au) were
264 simulated. Metal slabs were first equilibrated under atmospheric
265 pressure (1 atm) and a desired temperature (e.g. 200K). After
266 equilibration, butanethiol capping agents were placed at three-fold
267 sites on the Au(111) surfaces. The maximum butanethiol capacity on Au
268 surface is $1/3$ of the total number of surface Au
269 atoms\cite{vlugt:cpc2007154}. A series of different coverages was
270 investigated in order to study the relation between coverage and
271 interfacial conductance.
272
273 The capping agent molecules were allowed to migrate during the
274 simulations. They distributed themselves uniformly and sampled a
275 number of three-fold sites throughout out study. Therefore, the
276 initial configuration would not noticeably affect the sampling of a
277 variety of configurations of the same coverage, and the final
278 conductance measurement would be an average effect of these
279 configurations explored in the simulations. [MAY NEED FIGURES]
280
281 After the modified Au-butanethiol surface systems were equilibrated
282 under canonical ensemble, organic solvent molecules were packed in the
283 previously empty part of the simulation cells\cite{packmol}. Two
284 solvents were investigated, one which has little vibrational overlap
285 with the alkanethiol and a planar shape (toluene), and one which has
286 similar vibrational frequencies and chain-like shape ({\it n}-hexane).
287
288 The space filled by solvent molecules, i.e. the gap between
289 periodically repeated Au-butanethiol surfaces should be carefully
290 chosen. A very long length scale for the thermal gradient axis ($z$)
291 may cause excessively hot or cold temperatures in the middle of the
292 solvent region and lead to undesired phenomena such as solvent boiling
293 or freezing when a thermal flux is applied. Conversely, too few
294 solvent molecules would change the normal behavior of the liquid
295 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
296 these extreme cases did not happen to our simulations. And the
297 corresponding spacing is usually $35 \sim 60$\AA.
298
299 The initial configurations generated by Packmol are further
300 equilibrated with the $x$ and $y$ dimensions fixed, only allowing
301 length scale change in $z$ dimension. This is to ensure that the
302 equilibration of liquid phase does not affect the metal crystal
303 structure in $x$ and $y$ dimensions. Further equilibration are run
304 under NVT and then NVE ensembles.
305
306 After the systems reach equilibrium, NIVS is implemented to impose a
307 periodic unphysical thermal flux between the metal and the liquid
308 phase. Most of our simulations are under an average temperature of
309 $\sim$200K. Therefore, this flux usually comes from the metal to the
310 liquid so that the liquid has a higher temperature and would not
311 freeze due to excessively low temperature. This induced temperature
312 gradient is stablized and the simulation cell is devided evenly into
313 N slabs along the $z$-axis and the temperatures of each slab are
314 recorded. When the slab width $d$ of each slab is the same, the
315 derivatives of $T$ with respect to slab number $n$ can be directly
316 used for $G^\prime$ calculations:
317 \begin{equation}
318 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
319 \Big/\left(\frac{\partial T}{\partial z}\right)^2
320 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
321 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
322 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
323 \Big/\left(\frac{\partial T}{\partial n}\right)^2
324 \label{derivativeG2}
325 \end{equation}
326
327 \subsection{Force Field Parameters}
328 Our simulations include various components. Therefore, force field
329 parameter descriptions are needed for interactions both between the
330 same type of particles and between particles of different species.
331
332 The Au-Au interactions in metal lattice slab is described by the
333 quantum Sutton-Chen (QSC) formulation\cite{PhysRevB.59.3527}. The QSC
334 potentials include zero-point quantum corrections and are
335 reparametrized for accurate surface energies compared to the
336 Sutton-Chen potentials\cite{Chen90}.
337
338 Figure \ref{demoMol} demonstrates how we name our pseudo-atoms of the
339 organic solvent molecules in our simulations.
340
341 \begin{figure}
342 \includegraphics[width=\linewidth]{structures}
343 \caption{Structures of the capping agent and solvents utilized in
344 these simulations. The chemically-distinct sites (a-e) are expanded
345 in terms of constituent atoms for both United Atom (UA) and All Atom
346 (AA) force fields. Most parameters are from
347 Refs. \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} (UA) and
348 \protect\cite{OPLSAA} (AA). Cross-interactions with the Au atoms are given
349 in Table \ref{MnM}.}
350 \label{demoMol}
351 \end{figure}
352
353 For both solvent molecules, straight chain {\it n}-hexane and aromatic
354 toluene, United-Atom (UA) and All-Atom (AA) models are used
355 respectively. The TraPPE-UA
356 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
357 for our UA solvent molecules. In these models, pseudo-atoms are
358 located at the carbon centers for alkyl groups. By eliminating
359 explicit hydrogen atoms, these models are simple and computationally
360 efficient, while maintains good accuracy. However, the TraPPE-UA for
361 alkanes is known to predict a lower boiling point than experimental
362 values. Considering that after an unphysical thermal flux is applied
363 to a system, the temperature of ``hot'' area in the liquid phase would be
364 significantly higher than the average, to prevent over heating and
365 boiling of the liquid phase, the average temperature in our
366 simulations should be much lower than the liquid boiling point. [MORE DISCUSSION]
367 For UA-toluene model, rigid body constraints are applied, so that the
368 benzene ring and the methyl-CRar bond are kept rigid. This would save
369 computational time.[MORE DETAILS]
370
371 Besides the TraPPE-UA models, AA models for both organic solvents are
372 included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
373 force field is used. [MORE DETAILS]
374 For toluene, the United Force Field developed by Rapp\'{e} {\it et
375 al.}\cite{doi:10.1021/ja00051a040} is adopted.[MORE DETAILS]
376
377 The capping agent in our simulations, the butanethiol molecules can
378 either use UA or AA model. The TraPPE-UA force fields includes
379 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
380 UA butanethiol model in our simulations. The OPLS-AA also provides
381 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
382 surfaces do not have the hydrogen atom bonded to sulfur. To adapt this
383 change and derive suitable parameters for butanethiol adsorbed on
384 Au(111) surfaces, we adopt the S parameters from Luedtke and
385 Landman\cite{landman:1998} and modify parameters for its neighbor C
386 atom for charge balance in the molecule. Note that the model choice
387 (UA or AA) of capping agent can be different from the
388 solvent. Regardless of model choice, the force field parameters for
389 interactions between capping agent and solvent can be derived using
390 Lorentz-Berthelot Mixing Rule:
391 \begin{eqnarray}
392 \sigma_{IJ} & = & \frac{1}{2} \left(\sigma_{II} + \sigma_{JJ}\right) \\
393 \epsilon_{IJ} & = & \sqrt{\epsilon_{II}\epsilon_{JJ}}
394 \end{eqnarray}
395
396 To describe the interactions between metal Au and non-metal capping
397 agent and solvent particles, we refer to an adsorption study of alkyl
398 thiols on gold surfaces by Vlugt {\it et
399 al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
400 form of potential parameters for the interaction between Au and
401 pseudo-atoms CH$_x$ and S based on a well-established and widely-used
402 effective potential of Hautman and Klein\cite{hautman:4994} for the
403 Au(111) surface. As our simulations require the gold lattice slab to
404 be non-rigid so that it could accommodate kinetic energy for thermal
405 transport study purpose, the pair-wise form of potentials is
406 preferred.
407
408 Besides, the potentials developed from {\it ab initio} calculations by
409 Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
410 interactions between Au and aromatic C/H atoms in toluene.[MORE DETAILS]
411
412 However, the Lennard-Jones parameters between Au and other types of
413 particles in our simulations are not yet well-established. For these
414 interactions, we attempt to derive their parameters using the Mixing
415 Rule. To do this, the ``Metal-non-Metal'' (MnM) interaction parameters
416 for Au is first extracted from the Au-CH$_x$ parameters by applying
417 the Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
418 parameters in our simulations.
419
420 \begin{table*}
421 \begin{minipage}{\linewidth}
422 \begin{center}
423 \caption{Non-bonded interaction paramters for non-metal
424 particles and metal-non-metal interactions in our
425 simulations.}
426
427 \begin{tabular}{cccccc}
428 \hline\hline
429 Non-metal atom $I$ & $\sigma_{II}$ & $\epsilon_{II}$ & $q_I$ &
430 $\sigma_{AuI}$ & $\epsilon_{AuI}$ \\
431 (or pseudo-atom) & \AA & kcal/mol & & \AA & kcal/mol \\
432 \hline
433 CH3 & 3.75 & 0.1947 & - & 3.54 & 0.2146 \\
434 CH2 & 3.95 & 0.0914 & - & 3.54 & 0.1749 \\
435 CHar & 3.695 & 0.1003 & - & 3.4625 & 0.1680 \\
436 CRar & 3.88 & 0.04173 & - & 3.555 & 0.1604 \\
437 S & 4.45 & 0.25 & - & 2.40 & 8.465 \\
438 CT3 & 3.50 & 0.066 & -0.18 & 3.365 & 0.1373 \\
439 CT2 & 3.50 & 0.066 & -0.12 & 3.365 & 0.1373 \\
440 CTT & 3.50 & 0.066 & -0.065 & 3.365 & 0.1373 \\
441 HC & 2.50 & 0.030 & 0.06 & 2.865 & 0.09256 \\
442 CA & 3.55 & 0.070 & -0.115 & 3.173 & 0.0640 \\
443 HA & 2.42 & 0.030 & 0.115 & 2.746 & 0.0414 \\
444 \hline\hline
445 \end{tabular}
446 \label{MnM}
447 \end{center}
448 \end{minipage}
449 \end{table*}
450
451
452 \section{Results and Discussions}
453 [MAY HAVE A BRIEF SUMMARY]
454 \subsection{How Simulation Parameters Affects $G$}
455 [MAY NOT PUT AT FIRST]
456 We have varied our protocol or other parameters of the simulations in
457 order to investigate how these factors would affect the measurement of
458 $G$'s. It turned out that while some of these parameters would not
459 affect the results substantially, some other changes to the
460 simulations would have a significant impact on the measurement
461 results.
462
463 In some of our simulations, we allowed $L_x$ and $L_y$ to change
464 during equilibrating the liquid phase. Due to the stiffness of the Au
465 slab, $L_x$ and $L_y$ would not change noticeably after
466 equilibration. Although $L_z$ could fluctuate $\sim$1\% after a system
467 is fully equilibrated in the NPT ensemble, this fluctuation, as well
468 as those comparably smaller to $L_x$ and $L_y$, would not be magnified
469 on the calculated $G$'s, as shown in Table \ref{AuThiolHexaneUA}. This
470 insensivity to $L_i$ fluctuations allows reliable measurement of $G$'s
471 without the necessity of extremely cautious equilibration process.
472
473 As stated in our computational details, the spacing filled with
474 solvent molecules can be chosen within a range. This allows some
475 change of solvent molecule numbers for the same Au-butanethiol
476 surfaces. We did this study on our Au-butanethiol/hexane
477 simulations. Nevertheless, the results obtained from systems of
478 different $N_{hexane}$ did not indicate that the measurement of $G$ is
479 susceptible to this parameter. For computational efficiency concern,
480 smaller system size would be preferable, given that the liquid phase
481 structure is not affected.
482
483 Our NIVS algorithm allows change of unphysical thermal flux both in
484 direction and in quantity. This feature extends our investigation of
485 interfacial thermal conductance. However, the magnitude of this
486 thermal flux is not arbitary if one aims to obtain a stable and
487 reliable thermal gradient. A temperature profile would be
488 substantially affected by noise when $|J_z|$ has a much too low
489 magnitude; while an excessively large $|J_z|$ that overwhelms the
490 conductance capacity of the interface would prevent a thermal gradient
491 to reach a stablized steady state. NIVS has the advantage of allowing
492 $J$ to vary in a wide range such that the optimal flux range for $G$
493 measurement can generally be simulated by the algorithm. Within the
494 optimal range, we were able to study how $G$ would change according to
495 the thermal flux across the interface. For our simulations, we denote
496 $J_z$ to be positive when the physical thermal flux is from the liquid
497 to metal, and negative vice versa. The $G$'s measured under different
498 $J_z$ is listed in Table \ref{AuThiolHexaneUA} and [REF]. These
499 results do not suggest that $G$ is dependent on $J_z$ within this flux
500 range. The linear response of flux to thermal gradient simplifies our
501 investigations in that we can rely on $G$ measurement with only a
502 couple $J_z$'s and do not need to test a large series of fluxes.
503
504 %ADD MORE TO TABLE
505 \begin{table*}
506 \begin{minipage}{\linewidth}
507 \begin{center}
508 \caption{Computed interfacial thermal conductivity ($G$ and
509 $G^\prime$) values for the 100\% covered Au-butanethiol/hexane
510 interfaces with UA model and different hexane molecule numbers
511 at different temperatures using a range of energy fluxes.}
512
513 \begin{tabular}{ccccccc}
514 \hline\hline
515 $\langle T\rangle$ & $N_{hexane}$ & Fixed & $\rho_{hexane}$ &
516 $J_z$ & $G$ & $G^\prime$ \\
517 (K) & & $L_x$ \& $L_y$? & (g/cm$^3$) & (GW/m$^2$) &
518 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
519 \hline
520 200 & 266 & No & 0.672 & -0.96 & 102() & 80.0() \\
521 & 200 & Yes & 0.694 & 1.92 & 129() & 87.3() \\
522 & & Yes & 0.672 & 1.93 & 131() & 77.5() \\
523 & & No & 0.688 & 0.96 & 125() & 90.2() \\
524 & & & & 1.91 & 139() & 101() \\
525 & & & & 2.83 & 141() & 89.9() \\
526 & 166 & Yes & 0.679 & 0.97 & 115() & 69.3() \\
527 & & & & 1.94 & 125() & 87.1() \\
528 & & No & 0.681 & 0.97 & 141() & 77.7() \\
529 & & & & 1.92 & 138() & 98.9() \\
530 \hline
531 250 & 200 & No & 0.560 & 0.96 & 74.8() & 61.8() \\
532 & & & & -0.95 & 49.4() & 45.7() \\
533 & 166 & Yes & 0.570 & 0.98 & 79.0() & 62.9() \\
534 & & No & 0.569 & 0.97 & 80.3() & 67.1() \\
535 & & & & 1.44 & 76.2() & 64.8() \\
536 & & & & -0.95 & 56.4() & 54.4() \\
537 & & & & -1.85 & 47.8() & 53.5() \\
538 \hline\hline
539 \end{tabular}
540 \label{AuThiolHexaneUA}
541 \end{center}
542 \end{minipage}
543 \end{table*}
544
545 Furthermore, we also attempted to increase system average temperatures
546 to above 200K. These simulations are first equilibrated in the NPT
547 ensemble under normal pressure. As stated above, the TraPPE-UA model
548 for hexane tends to predict a lower boiling point. In our simulations,
549 hexane had diffculty to remain in liquid phase when NPT equilibration
550 temperature is higher than 250K. Additionally, the equilibrated liquid
551 hexane density under 250K becomes lower than experimental value. This
552 expanded liquid phase leads to lower contact between hexane and
553 butanethiol as well.[MAY NEED FIGURE] And this reduced contact would
554 probably be accountable for a lower interfacial thermal conductance,
555 as shown in Table \ref{AuThiolHexaneUA}.
556
557 A similar study for TraPPE-UA toluene agrees with the above result as
558 well. Having a higher boiling point, toluene tends to remain liquid in
559 our simulations even equilibrated under 300K in NPT
560 ensembles. Furthermore, the expansion of the toluene liquid phase is
561 not as significant as that of the hexane. This prevents severe
562 decrease of liquid-capping agent contact and the results (Table
563 \ref{AuThiolToluene}) show only a slightly decreased interface
564 conductance. Therefore, solvent-capping agent contact should play an
565 important role in the thermal transport process across the interface
566 in that higher degree of contact could yield increased conductance.
567
568 [ADD ERROR ESTIMATE TO TABLE]
569 \begin{table*}
570 \begin{minipage}{\linewidth}
571 \begin{center}
572 \caption{Computed interfacial thermal conductivity ($G$ and
573 $G^\prime$) values for a 90\% coverage Au-butanethiol/toluene
574 interface at different temperatures using a range of energy
575 fluxes.}
576
577 \begin{tabular}{ccccc}
578 \hline\hline
579 $\langle T\rangle$ & $\rho_{toluene}$ & $J_z$ & $G$ & $G^\prime$ \\
580 (K) & (g/cm$^3$) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
581 \hline
582 200 & 0.933 & -1.86 & 180() & 135() \\
583 & & 2.15 & 204() & 113() \\
584 & & -3.93 & 175() & 114() \\
585 \hline
586 300 & 0.855 & -1.91 & 143() & 125() \\
587 & & -4.19 & 134() & 113() \\
588 \hline\hline
589 \end{tabular}
590 \label{AuThiolToluene}
591 \end{center}
592 \end{minipage}
593 \end{table*}
594
595 Besides lower interfacial thermal conductance, surfaces in relatively
596 high temperatures are susceptible to reconstructions, when
597 butanethiols have a full coverage on the Au(111) surface. These
598 reconstructions include surface Au atoms migrated outward to the S
599 atom layer, and butanethiol molecules embedded into the original
600 surface Au layer. The driving force for this behavior is the strong
601 Au-S interactions in our simulations. And these reconstructions lead
602 to higher ratio of Au-S attraction and thus is energetically
603 favorable. Furthermore, this phenomenon agrees with experimental
604 results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
605 {\it et al.} had kept their Au(111) slab rigid so that their
606 simulations can reach 300K without surface reconstructions. Without
607 this practice, simulating 100\% thiol covered interfaces under higher
608 temperatures could hardly avoid surface reconstructions. However, our
609 measurement is based on assuming homogeneity on $x$ and $y$ dimensions
610 so that measurement of $T$ at particular $z$ would be an effective
611 average of the particles of the same type. Since surface
612 reconstructions could eliminate the original $x$ and $y$ dimensional
613 homogeneity, measurement of $G$ is more difficult to conduct under
614 higher temperatures. Therefore, most of our measurements are
615 undertaken at $\langle T\rangle\sim$200K.
616
617 However, when the surface is not completely covered by butanethiols,
618 the simulated system is more resistent to the reconstruction
619 above. Our Au-butanethiol/toluene system did not see this phenomena
620 even at $\langle T\rangle\sim$300K. The Au(111) surfaces have a 90\%
621 coverage of butanethiols and have empty three-fold sites. These empty
622 sites could help prevent surface reconstruction in that they provide
623 other means of capping agent relaxation. It is observed that
624 butanethiols can migrate to their neighbor empty sites during a
625 simulation. Therefore, we were able to obtain $G$'s for these
626 interfaces even at a relatively high temperature without being
627 affected by surface reconstructions.
628
629 \subsection{Influence of Capping Agent Coverage on $G$}
630 To investigate the influence of butanethiol coverage on interfacial
631 thermal conductance, a series of different coverage Au-butanethiol
632 surfaces is prepared and solvated with various organic
633 molecules. These systems are then equilibrated and their interfacial
634 thermal conductivity are measured with our NIVS algorithm. Table
635 \ref{tlnUhxnUhxnD} lists these results for direct comparison between
636 different coverages of butanethiol. To study the isotope effect in
637 interfacial thermal conductance, deuterated UA-hexane is included as
638 well.
639
640 It turned out that with partial covered butanethiol on the Au(111)
641 surface, the derivative definition for $G$ (Eq. \ref{derivativeG}) has
642 difficulty to apply, due to the difficulty in locating the maximum of
643 change of $\lambda$. Instead, the discrete definition
644 (Eq. \ref{discreteG}) is easier to apply, as max($\Delta T$) can still
645 be well-defined. Therefore, $G$'s (not $G^\prime$) are used for this
646 section.
647
648 From Table \ref{tlnUhxnUhxnD}, one can see the significance of the
649 presence of capping agents. Even when a fraction of the Au(111)
650 surface sites are covered with butanethiols, the conductivity would
651 see an enhancement by at least a factor of 3. This indicates the
652 important role cappping agent is playing for thermal transport
653 phenomena on metal/organic solvent surfaces.
654
655 Interestingly, as one could observe from our results, the maximum
656 conductance enhancement (largest $G$) happens while the surfaces are
657 about 75\% covered with butanethiols. This again indicates that
658 solvent-capping agent contact has an important role of the thermal
659 transport process. Slightly lower butanethiol coverage allows small
660 gaps between butanethiols to form. And these gaps could be filled with
661 solvent molecules, which acts like ``heat conductors'' on the
662 surface. The higher degree of interaction between these solvent
663 molecules and capping agents increases the enhancement effect and thus
664 produces a higher $G$ than densely packed butanethiol arrays. However,
665 once this maximum conductance enhancement is reached, $G$ decreases
666 when butanethiol coverage continues to decrease. Each capping agent
667 molecule reaches its maximum capacity for thermal
668 conductance. Therefore, even higher solvent-capping agent contact
669 would not offset this effect. Eventually, when butanethiol coverage
670 continues to decrease, solvent-capping agent contact actually
671 decreases with the disappearing of butanethiol molecules. In this
672 case, $G$ decrease could not be offset but instead accelerated.
673
674 A comparison of the results obtained from differenet organic solvents
675 can also provide useful information of the interfacial thermal
676 transport process. The deuterated hexane (UA) results do not appear to
677 be much different from those of normal hexane (UA), given that
678 butanethiol (UA) is non-deuterated for both solvents. These UA model
679 studies, even though eliminating C-H vibration samplings, still have
680 C-C vibrational frequencies different from each other. However, these
681 differences in the infrared range do not seem to produce an observable
682 difference for the results of $G$. [MAY NEED FIGURE]
683
684 Furthermore, results for rigid body toluene solvent, as well as other
685 UA-hexane solvents, are reasonable within the general experimental
686 ranges[CITATIONS]. This suggests that explicit hydrogen might not be a
687 required factor for modeling thermal transport phenomena of systems
688 such as Au-thiol/organic solvent.
689
690 However, results for Au-butanethiol/toluene do not show an identical
691 trend with those for Au-butanethiol/hexane in that $G$'s remain at
692 approximately the same magnitue when butanethiol coverage differs from
693 25\% to 75\%. This might be rooted in the molecule shape difference
694 for plane-like toluene and chain-like {\it n}-hexane. Due to this
695 difference, toluene molecules have more difficulty in occupying
696 relatively small gaps among capping agents when their coverage is not
697 too low. Therefore, the solvent-capping agent contact may keep
698 increasing until the capping agent coverage reaches a relatively low
699 level. This becomes an offset for decreasing butanethiol molecules on
700 its effect to the process of interfacial thermal transport. Thus, one
701 can see a plateau of $G$ vs. butanethiol coverage in our results.
702
703 [NEED ERROR ESTIMATE]
704 \begin{figure}
705 \includegraphics[width=\linewidth]{coverage}
706 \caption{Comparison of interfacial thermal conductivity ($G$) values
707 for the Au-butanethiol/solvent interface with various UA models and
708 different capping agent coverages at $\langle T\rangle\sim$200K
709 using certain energy flux respectively.}
710 \label{coverage}
711 \end{figure}
712
713 \subsection{Influence of Chosen Molecule Model on $G$}
714 [MAY COMBINE W MECHANISM STUDY]
715
716 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
724 [MORE DATA; ERROR ESTIMATE]
725 \begin{table*}
726 \begin{minipage}{\linewidth}
727 \begin{center}
728
729 \caption{Computed interfacial thermal conductivity ($G$ and
730 $G^\prime$) values for interfaces using various models for
731 solvent and capping agent (or without capping agent) at
732 $\langle T\rangle\sim$200K. (D stands for deuterated solvent
733 or capping agent molecules; ``Avg.'' denotes results that are
734 averages of several simulations.)}
735
736 \begin{tabular}{ccccc}
737 \hline\hline
738 Butanethiol model & Solvent & $J_z$ & $G$ & $G^\prime$ \\
739 (or bare surface) & model & (GW/m$^2$) &
740 \multicolumn{2}{c}{(MW/m$^2$/K)} \\
741 \hline
742 UA & UA hexane & Avg. & 131() & 86.5() \\
743 & UA hexane(D) & 1.95 & 153() & 136() \\
744 & AA hexane & 1.94 & 135() & 129() \\
745 & & 2.86 & 126() & 115() \\
746 & UA toluene & 1.96 & 187() & 151() \\
747 & AA toluene & 1.89 & 200() & 149() \\
748 \hline
749 AA & UA hexane & 1.94 & 116() & 129() \\
750 & AA hexane & Avg. & 442() & 356() \\
751 & AA hexane(D) & 1.93 & 222() & 234() \\
752 & UA toluene & 1.98 & 125() & 96.5() \\
753 & AA toluene & 3.79 & 487() & 290() \\
754 \hline
755 AA(D) & UA hexane & 1.94 & 158() & 172() \\
756 & AA hexane & 1.92 & 243() & 191() \\
757 & AA toluene & 1.93 & 364() & 322() \\
758 \hline
759 bare & UA hexane & Avg. & 46.5() & 49.4() \\
760 & UA hexane(D) & 0.98 & 43.9() & 43.0() \\
761 & AA hexane & 0.96 & 31.0() & 29.4() \\
762 & UA toluene & 1.99 & 70.1() & 65.8() \\
763 \hline\hline
764 \end{tabular}
765 \label{modelTest}
766 \end{center}
767 \end{minipage}
768 \end{table*}
769
770 To facilitate direct comparison, the same system with differnt models
771 for different components uses the same length scale for their
772 simulation cells. Without the presence of capping agent, using
773 different models for hexane yields similar results for both $G$ and
774 $G^\prime$, and these two definitions agree with eath other very
775 well. This indicates very weak interaction between the metal and the
776 solvent, and is a typical case for acoustic impedance mismatch between
777 these two phases.
778
779 As for Au(111) surfaces completely covered by butanethiols, the choice
780 of models for capping agent and solvent could impact the measurement
781 of $G$ and $G^\prime$ quite significantly. For Au-butanethiol/hexane
782 interfaces, using AA model for both butanethiol and hexane yields
783 substantially higher conductivity values than using UA model for at
784 least one component of the solvent and capping agent, which exceeds
785 the upper bond of experimental value range. This is probably due to
786 the classically treated C-H vibrations in the AA model, which should
787 not be appreciably populated at normal temperatures. In comparison,
788 once either the hexanes or the butanethiols are deuterated, one can
789 see a significantly lower $G$ and $G^\prime$. In either of these
790 cases, the C-H(D) vibrational overlap between the solvent and the
791 capping agent is removed. [MAY NEED FIGURE] Conclusively, the
792 improperly treated C-H vibration in the AA model produced
793 over-predicted results accordingly. Compared to the AA model, the UA
794 model yields more reasonable results with higher computational
795 efficiency.
796
797 However, for Au-butanethiol/toluene interfaces, having the AA
798 butanethiol deuterated did not yield a significant change in the
799 measurement results. Compared to the C-H vibrational overlap between
800 hexane and butanethiol, both of which have alkyl chains, that overlap
801 between toluene and butanethiol is not so significant and thus does
802 not have as much contribution to the ``Intramolecular Vibration
803 Redistribution''[CITE HASE]. Conversely, extra degrees of freedom such
804 as the C-H vibrations could yield higher heat exchange rate between
805 these two phases and result in a much higher conductivity.
806
807 Although the QSC model for Au is known to predict an overly low value
808 for bulk metal gold conductivity\cite{kuang:164101}, our computational
809 results for $G$ and $G^\prime$ do not seem to be affected by this
810 drawback of the model for metal. Instead, our results suggest that the
811 modeling of interfacial thermal transport behavior relies mainly on
812 the accuracy of the interaction descriptions between components
813 occupying the interfaces.
814
815 \subsection{Mechanism of Interfacial Thermal Conductance Enhancement
816 by Capping Agent}
817 %OR\subsection{Vibrational spectrum study on conductance mechanism}
818
819 [MAY INTRODUCE PROTOCOL IN METHODOLOGY/COMPUTATIONAL DETAIL, EQN'S]
820
821 To investigate the mechanism of this interfacial thermal conductance,
822 the vibrational spectra of various gold systems were obtained and are
823 shown as in the upper panel of Fig. \ref{vibration}. To obtain these
824 spectra, one first runs a simulation in the NVE ensemble and collects
825 snapshots of configurations; these configurations are used to compute
826 the velocity auto-correlation functions, which is used to construct a
827 power spectrum via a Fourier transform.
828
829 [MAY RELATE TO HASE'S]
830 The gold surfaces covered by
831 butanethiol molecules, compared to bare gold surfaces, exhibit an
832 additional peak observed at a frequency of $\sim$170cm$^{-1}$, which
833 is attributed to the vibration of the S-Au bond. This vibration
834 enables efficient thermal transport from surface Au atoms to the
835 capping agents. Simultaneously, as shown in the lower panel of
836 Fig. \ref{vibration}, the large overlap of the vibration spectra of
837 butanethiol and hexane in the all-atom model, including the C-H
838 vibration, also suggests high thermal exchange efficiency. The
839 combination of these two effects produces the drastic interfacial
840 thermal conductance enhancement in the all-atom model.
841
842 [REDO. MAY NEED TO CONVERT TO JPEG]
843 \begin{figure}
844 \includegraphics[width=\linewidth]{vibration}
845 \caption{Vibrational spectra obtained for gold in different
846 environments (upper panel) and for Au/thiol/hexane simulation in
847 all-atom model (lower panel).}
848 \label{vibration}
849 \end{figure}
850
851 [COMPARISON OF TWO G'S; AU SLAB WIDTHS; ETC]
852 % The results show that the two definitions used for $G$ yield
853 % comparable values, though $G^\prime$ tends to be smaller.
854
855 \section{Conclusions}
856 The NIVS algorithm we developed has been applied to simulations of
857 Au-butanethiol surfaces with organic solvents. This algorithm allows
858 effective unphysical thermal flux transferred between the metal and
859 the liquid phase. With the flux applied, we were able to measure the
860 corresponding thermal gradient and to obtain interfacial thermal
861 conductivities. Our simulations have seen significant conductance
862 enhancement with the presence of capping agent, compared to the bare
863 gold/liquid interfaces. The acoustic impedance mismatch between the
864 metal and the liquid phase is effectively eliminated by proper capping
865 agent. Furthermore, the coverage precentage of the capping agent plays
866 an important role in the interfacial thermal transport process.
867
868 Our measurement results, particularly of the UA models, agree with
869 available experimental data. This indicates that our force field
870 parameters have a nice description of the interactions between the
871 particles at the interfaces. AA models tend to overestimate the
872 interfacial thermal conductance in that the classically treated C-H
873 vibration would be overly sampled. Compared to the AA models, the UA
874 models have higher computational efficiency with satisfactory
875 accuracy, and thus are preferable in interfacial thermal transport
876 modelings.
877
878 Vlugt {\it et al.} has investigated the surface thiol structures for
879 nanocrystal gold and pointed out that they differs from those of the
880 Au(111) surface\cite{vlugt:cpc2007154}. This difference might lead to
881 change of interfacial thermal transport behavior as well. To
882 investigate this problem, an effective means to introduce thermal flux
883 and measure the corresponding thermal gradient is desirable for
884 simulating structures with spherical symmetry.
885
886
887 \section{Acknowledgments}
888 Support for this project was provided by the National Science
889 Foundation under grant CHE-0848243. Computational time was provided by
890 the Center for Research Computing (CRC) at the University of Notre
891 Dame. \newpage
892
893 \bibliography{interfacial}
894
895 \end{doublespace}
896 \end{document}
897