ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
(Generate patch)

Comparing interfacial/interfacial.tex (file contents):
Revision 3721 by skuang, Sat Feb 5 00:02:11 2011 UTC vs.
Revision 3730 by skuang, Tue Jul 5 17:39:21 2011 UTC

# Line 22 | Line 22
22   \setlength{\abovecaptionskip}{20 pt}
23   \setlength{\belowcaptionskip}{30 pt}
24  
25 < %\renewcommand\citemid{\ } % no comma in optional referenc note
25 > %\renewcommand\citemid{\ } % no comma in optional reference note
26   \bibpunct{[}{]}{,}{s}{}{;}
27   \bibliographystyle{aip}
28  
# Line 44 | Line 44 | The abstract version 2
44   \begin{doublespace}
45  
46   \begin{abstract}
47 < The abstract version 2
47 >
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   \end{abstract}
64  
65   \newpage
# Line 56 | Line 71 | The abstract version 2
71   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72  
73   \section{Introduction}
74 + [BACKGROUND FOR INTERFACIAL THERMAL CONDUCTANCE PROBLEM]
75 + 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  
89 < The intro.
89 > 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   \section{Methodology}
99   \subsection{Algorithm}
100 + [BACKGROUND FOR MD METHODS]
101   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
# Line 95 | Line 132 | external thermostat.
132   algorithm conserves momenta and energy and does not depend on an
133   external thermostat.
134  
135 < (wondering how much detail of algorithm should be put here...)
135 > \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  
151 < \subsection{Simulation Parameters}
152 < Our simulation systems consists of metal gold lattice slab solvated by
102 < organic solvents. In order to study the role of capping agents in
103 < interfacial thermal conductance, butanethiol is chosen to cover gold
104 < surfaces in comparison to no capping agent present.
151 > When the interfacial conductance is {\it not} small, two ways can be
152 > used to define $G$.
153  
154 < The Au-Au interactions in metal lattice slab is described by the
155 < quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
156 < potentials include zero-point quantum corrections and are
157 < reparametrized for accurate surface energies compared to the
158 < Sutton-Chen potentials\cite{Chen90}.
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 < Straight chain {\it n}-hexane and aromatic toluene are respectively
164 < used as solvents. For hexane, both United-Atom\cite{TraPPE-UA.alkanes}
165 < and All-Atom\cite{OPLSAA} force fields are used for comparison; for
166 < toluene, United-Atom\cite{TraPPE-UA.alkylbenzenes} force fields are
167 < used with rigid body constraints applied. (maybe needs more details
168 < about rigid body)
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 < Buatnethiol molecules are used as capping agent for some of our
177 < simulations. United-Atom\cite{TraPPE-UA.thiols} and All-Atom models
178 < are respectively used corresponding to the force field type of
122 < solvent.
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 < To describe the interactions between metal Au and non-metal capping
125 < agent and solvent, we refer to Vlugt\cite{vlugt:cpc2007154} and derive
126 < other interactions which are not parametrized in their work. (can add
127 < hautman and klein's paper here and more discussion; need to put
128 < aromatic-metal interaction approximation here)
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 + [MAY INCLUDE POWER SPECTRUM PROTOCOL]
215 +
216   \section{Computational Details}
217 < Our simulation systems consists of a lattice Au slab with the (111)
218 < surface perpendicular to the $z$-axis, and a solvent layer between the
219 < periodic Au slabs along the $z$-axis. To set up the interfacial
220 < system, the Au slab is first equilibrated without solvent under room
221 < pressure and a desired temperature. After the metal slab is
222 < equilibrated, United-Atom or All-Atom butanethiols are replicated on
223 < the Au surface, each occupying the (??) among three Au atoms, and is
224 < equilibrated under NVT ensemble. According to (CITATION), the maximal
225 < thiol capacity on Au surface is $1/3$ of the total number of surface
226 < Au atoms.
217 > \subsection{Simulation Protocol}
218 > In our simulations, Au is used to construct a metal slab with bare
219 > (111) surface perpendicular to the $z$-axis. Different slab thickness
220 > (layer numbers of Au) are simulated. This metal slab is first
221 > equilibrated under normal pressure (1 atm) and a desired
222 > temperature. After equilibration, butanethiol is used as the capping
223 > agent molecule to cover the bare Au (111) surfaces evenly. The sulfur
224 > atoms in the butanethiol molecules would occupy the three-fold sites
225 > of the surfaces, and the maximal butanethiol capacity on Au surface is
226 > $1/3$ of the total number of surface Au atoms[CITATION]. A series of
227 > different coverage surfaces is investigated in order to study the
228 > relation between coverage and conductance.
229  
230 < \cite{packmol}
230 > [COVERAGE DISCRIPTION] However, since the interactions between surface
231 > Au and butanethiol is non-bonded, the capping agent molecules are
232 > allowed to migrate to an empty neighbor three-fold site during a
233 > simulation. Therefore, the initial configuration would not severely
234 > affect the sampling of a variety of configurations of the same
235 > coverage, and the final conductance measurement would be an average
236 > effect of these configurations explored in the simulations. [MAY NEED FIGURES]
237 >
238 > After the modified Au-butanethiol surface systems are equilibrated
239 > under canonical ensemble, Packmol\cite{packmol} is used to pack
240 > organic solvent molecules in the previously vacuum part of the
241 > simulation cells, which guarantees that short range repulsive
242 > interactions do not disrupt the simulations. Two solvents are
243 > investigated, one which has little vibrational overlap with the
244 > alkanethiol and plane-like shape (toluene), and one which has similar
245 > vibrational frequencies and chain-like shape ({\it n}-hexane). [MAY
246 > EXPLAIN WHY WE CHOOSE THEM]
247 >
248 > The spacing filled by solvent molecules, i.e. the gap between
249 > periodically repeated Au-butanethiol surfaces should be carefully
250 > chosen. A very long length scale for the thermal gradient axis ($z$)
251 > may cause excessively hot or cold temperatures in the middle of the
252 > solvent region and lead to undesired phenomena such as solvent boiling
253 > or freezing when a thermal flux is applied. Conversely, too few
254 > solvent molecules would change the normal behavior of the liquid
255 > phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
256 > these extreme cases did not happen to our simulations. And the
257 > corresponding spacing is usually $35 \sim 60$\AA.
258 >
259 > The initial configurations generated by Packmol are further
260 > equilibrated with the $x$ and $y$ dimensions fixed, only allowing
261 > length scale change in $z$ dimension. This is to ensure that the
262 > equilibration of liquid phase does not affect the metal crystal
263 > structure in $x$ and $y$ dimensions. Further equilibration are run
264 > under NVT and then NVE ensembles.
265 >
266 > After the systems reach equilibrium, NIVS is implemented to impose a
267 > periodic unphysical thermal flux between the metal and the liquid
268 > phase. Most of our simulations are under an average temperature of
269 > $\sim$200K. Therefore, this flux usually comes from the metal to the
270 > liquid so that the liquid has a higher temperature and would not
271 > freeze due to excessively low temperature. This induced temperature
272 > gradient is stablized and the simulation cell is devided evenly into
273 > N slabs along the $z$-axis and the temperatures of each slab are
274 > recorded. When the slab width $d$ of each slab is the same, the
275 > derivatives of $T$ with respect to slab number $n$ can be directly
276 > used for $G^\prime$ calculations:
277 > \begin{equation}
278 > G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
279 >         \Big/\left(\frac{\partial T}{\partial z}\right)^2
280 >         = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
281 >         \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
282 >         = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
283 >         \Big/\left(\frac{\partial T}{\partial n}\right)^2
284 > \label{derivativeG2}
285 > \end{equation}
286 >
287 > \subsection{Force Field Parameters}
288 > Our simulations include various components. Therefore, force field
289 > parameter descriptions are needed for interactions both between the
290 > same type of particles and between particles of different species.
291 >
292 > The Au-Au interactions in metal lattice slab is described by the
293 > quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
294 > potentials include zero-point quantum corrections and are
295 > reparametrized for accurate surface energies compared to the
296 > Sutton-Chen potentials\cite{Chen90}.
297 >
298 > Figure [REF] demonstrates how we name our pseudo-atoms of the
299 > molecules in our simulations.
300 > [FIGURE FOR MOLECULE NOMENCLATURE]
301 >
302 > For both solvent molecules, straight chain {\it n}-hexane and aromatic
303 > toluene, United-Atom (UA) and All-Atom (AA) models are used
304 > respectively. The TraPPE-UA
305 > parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
306 > for our UA solvent molecules. In these models, pseudo-atoms are
307 > located at the carbon centers for alkyl groups. By eliminating
308 > explicit hydrogen atoms, these models are simple and computationally
309 > efficient, while maintains good accuracy. However, the TraPPE-UA for
310 > alkanes is known to predict a lower boiling point than experimental
311 > values. Considering that after an unphysical thermal flux is applied
312 > to a system, the temperature of ``hot'' area in the liquid phase would be
313 > significantly higher than the average, to prevent over heating and
314 > boiling of the liquid phase, the average temperature in our
315 > simulations should be much lower than the liquid boiling point. [MORE DISCUSSION]
316 > For UA-toluene model, rigid body constraints are applied, so that the
317 > benzene ring and the methyl-CRar bond are kept rigid. This would save
318 > computational time.[MORE DETAILS]
319 >
320 > Besides the TraPPE-UA models, AA models for both organic solvents are
321 > included in our studies as well. For hexane, the OPLS-AA\cite{OPLSAA}
322 > force field is used. [MORE DETAILS]
323 > For toluene, the United Force Field developed by Rapp\'{e} {\it et
324 >  al.}\cite{doi:10.1021/ja00051a040} is adopted.[MORE DETAILS]
325 >
326 > The capping agent in our simulations, the butanethiol molecules can
327 > either use UA or AA model. The TraPPE-UA force fields includes
328 > parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
329 > UA butanethiol model in our simulations. The OPLS-AA also provides
330 > parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
331 > surfaces do not have the hydrogen atom bonded to sulfur. To adapt this
332 > change and derive suitable parameters for butanethiol adsorbed on
333 > Au(111) surfaces, we adopt the S parameters from [CITATION CF VLUGT]
334 > and modify parameters for its neighbor C atom for charge balance in
335 > the molecule. Note that the model choice (UA or AA) of capping agent
336 > can be different from the solvent. Regardless of model choice, the
337 > force field parameters for interactions between capping agent and
338 > solvent can be derived using Lorentz-Berthelot Mixing Rule:
339 >
340 >
341 > To describe the interactions between metal Au and non-metal capping
342 > agent and solvent particles, we refer to an adsorption study of alkyl
343 > thiols on gold surfaces by Vlugt {\it et
344 >  al.}\cite{vlugt:cpc2007154} They fitted an effective Lennard-Jones
345 > form of potential parameters for the interaction between Au and
346 > pseudo-atoms CH$_x$ and S based on a well-established and widely-used
347 > effective potential of Hautman and Klein[CITATION] for the Au(111)
348 > surface. As our simulations require the gold lattice slab to be
349 > non-rigid so that it could accommodate kinetic energy for thermal
350 > transport study purpose, the pair-wise form of potentials is
351 > preferred.
352 >
353 > Besides, the potentials developed from {\it ab initio} calculations by
354 > Leng {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
355 > interactions between Au and aromatic C/H atoms in toluene.[MORE DETAILS]
356 >
357 > However, the Lennard-Jones parameters between Au and other types of
358 > particles in our simulations are not yet well-established. For these
359 > interactions, we attempt to derive their parameters using the Mixing
360 > Rule. To do this, the ``Metal-non-Metal'' (MnM) interaction parameters
361 > for Au is first extracted from the Au-CH$_x$ parameters by applying
362 > the Mixing Rule reversely. Table \ref{MnM} summarizes these ``MnM''
363 > parameters in our simulations.
364 >
365 > \begin{table*}
366 >  \begin{minipage}{\linewidth}
367 >    \begin{center}
368 >      \caption{Lennard-Jones parameters for Au-non-Metal
369 >        interactions in our simulations.}
370 >      
371 >      \begin{tabular}{ccc}
372 >        \hline\hline
373 >        Non-metal & $\sigma$/\AA & $\epsilon$/kcal/mol \\
374 >        \hline
375 >        S    & 2.40   & 8.465   \\
376 >        CH3  & 3.54   & 0.2146  \\
377 >        CH2  & 3.54   & 0.1749  \\
378 >        CT3  & 3.365  & 0.1373  \\
379 >        CT2  & 3.365  & 0.1373  \\
380 >        CTT  & 3.365  & 0.1373  \\
381 >        HC   & 2.865  & 0.09256 \\
382 >        CHar & 3.4625 & 0.1680  \\
383 >        CRar & 3.555  & 0.1604  \\
384 >        CA   & 3.173  & 0.0640  \\
385 >        HA   & 2.746  & 0.0414  \\
386 >        \hline\hline
387 >      \end{tabular}
388 >      \label{MnM}
389 >    \end{center}
390 >  \end{minipage}
391 > \end{table*}
392  
393 +
394 + \section{Results and Discussions}
395 + [MAY HAVE A BRIEF SUMMARY]
396 + \subsection{How Simulation Parameters Affects $G$}
397 + [MAY NOT PUT AT FIRST]
398 + We have varied our protocol or other parameters of the simulations in
399 + order to investigate how these factors would affect the measurement of
400 + $G$'s. It turned out that while some of these parameters would not
401 + affect the results substantially, some other changes to the
402 + simulations would have a significant impact on the measurement
403 + results.
404 +
405 + In some of our simulations, we allowed $L_x$ and $L_y$ to change
406 + during equilibrating the liquid phase. Due to the stiffness of the Au
407 + slab, $L_x$ and $L_y$ would not change noticeably after
408 + equilibration. Although $L_z$ could fluctuate $\sim$1\% after a system
409 + is fully equilibrated in the NPT ensemble, this fluctuation, as well
410 + as those comparably smaller to $L_x$ and $L_y$, would not be magnified
411 + on the calculated $G$'s, as shown in Table \ref{AuThiolHexaneUA}. This
412 + insensivity to $L_i$ fluctuations allows reliable measurement of $G$'s
413 + without the necessity of extremely cautious equilibration process.
414 +
415 + As stated in our computational details, the spacing filled with
416 + solvent molecules can be chosen within a range. This allows some
417 + change of solvent molecule numbers for the same Au-butanethiol
418 + surfaces. We did this study on our Au-butanethiol/hexane
419 + simulations. Nevertheless, the results obtained from systems of
420 + different $N_{hexane}$ did not indicate that the measurement of $G$ is
421 + susceptible to this parameter. For computational efficiency concern,
422 + smaller system size would be preferable, given that the liquid phase
423 + structure is not affected.
424 +
425 + Our NIVS algorithm allows change of unphysical thermal flux both in
426 + direction and in quantity. This feature extends our investigation of
427 + interfacial thermal conductance. However, the magnitude of this
428 + thermal flux is not arbitary if one aims to obtain a stable and
429 + reliable thermal gradient. A temperature profile would be
430 + substantially affected by noise when $|J_z|$ has a much too low
431 + magnitude; while an excessively large $|J_z|$ that overwhelms the
432 + conductance capacity of the interface would prevent a thermal gradient
433 + to reach a stablized steady state. NIVS has the advantage of allowing
434 + $J$ to vary in a wide range such that the optimal flux range for $G$
435 + measurement can generally be simulated by the algorithm. Within the
436 + optimal range, we were able to study how $G$ would change according to
437 + the thermal flux across the interface. For our simulations, we denote
438 + $J_z$ to be positive when the physical thermal flux is from the liquid
439 + to metal, and negative vice versa. The $G$'s measured under different
440 + $J_z$ is listed in Table \ref{AuThiolHexaneUA} and [REF]. These
441 + results do not suggest that $G$ is dependent on $J_z$ within this flux
442 + range. The linear response of flux to thermal gradient simplifies our
443 + investigations in that we can rely on $G$ measurement with only a
444 + couple $J_z$'s and do not need to test a large series of fluxes.
445 +
446 + %ADD MORE TO TABLE
447 + \begin{table*}
448 +  \begin{minipage}{\linewidth}
449 +    \begin{center}
450 +      \caption{Computed interfacial thermal conductivity ($G$ and
451 +        $G^\prime$) values for the Au/butanethiol/hexane interface
452 +        with united-atom model and different capping agent coverage
453 +        and solvent molecule numbers at different temperatures using a
454 +        range of energy fluxes.}
455 +      
456 +      \begin{tabular}{cccccc}
457 +        \hline\hline
458 +        Thiol & $\langle T\rangle$ & & $J_z$ & $G$ & $G^\prime$ \\
459 +        coverage (\%) & (K) & $N_{hexane}$ & (GW/m$^2$) &
460 +        \multicolumn{2}{c}{(MW/m$^2$/K)} \\
461 +        \hline
462 +        0.0   & 200 & 200 & 0.96 & 43.3 & 42.7 \\
463 +              &     &     & 1.91 & 45.7 & 42.9 \\
464 +              &     & 166 & 0.96 & 43.1 & 53.4 \\
465 +        88.9  & 200 & 166 & 1.94 & 172  & 108  \\
466 +        100.0 & 250 & 200 & 0.96 & 81.8 & 67.0 \\
467 +              &     & 166 & 0.98 & 79.0 & 62.9 \\
468 +              &     &     & 1.44 & 76.2 & 64.8 \\
469 +              & 200 & 200 & 1.92 & 129  & 87.3 \\
470 +              &     &     & 1.93 & 131  & 77.5 \\
471 +              &     & 166 & 0.97 & 115  & 69.3 \\
472 +              &     &     & 1.94 & 125  & 87.1 \\
473 +        \hline\hline
474 +      \end{tabular}
475 +      \label{AuThiolHexaneUA}
476 +    \end{center}
477 +  \end{minipage}
478 + \end{table*}
479 +
480 + Furthermore, we also attempted to increase system average temperatures
481 + to above 200K. These simulations are first equilibrated in the NPT
482 + ensemble under normal pressure. As stated above, the TraPPE-UA model
483 + for hexane tends to predict a lower boiling point. In our simulations,
484 + hexane had diffculty to remain in liquid phase when NPT equilibration
485 + temperature is higher than 250K. Additionally, the equilibrated liquid
486 + hexane density under 250K becomes lower than experimental value. This
487 + expanded liquid phase leads to lower contact between hexane and
488 + butanethiol as well.[MAY NEED FIGURE] And this reduced contact would
489 + probably be accountable for a lower interfacial thermal conductance,
490 + as shown in Table \ref{AuThiolHexaneUA}.
491 +
492 + A similar study for TraPPE-UA toluene agrees with the above result as
493 + well. Having a higher boiling point, toluene tends to remain liquid in
494 + our simulations even equilibrated under 300K in NPT
495 + ensembles. Furthermore, the expansion of the toluene liquid phase is
496 + not as significant as that of the hexane. This prevents severe
497 + decrease of liquid-capping agent contact and the results (Table
498 + \ref{AuThiolToluene}) show only a slightly decreased interface
499 + conductance. Therefore, solvent-capping agent contact should play an
500 + important role in the thermal transport process across the interface
501 + in that higher degree of contact could yield increased conductance.
502 +
503 + [ADD SIGNS AND ERROR ESTIMATE TO TABLE]
504 + \begin{table*}
505 +  \begin{minipage}{\linewidth}
506 +    \begin{center}
507 +      \caption{Computed interfacial thermal conductivity ($G$ and
508 +        $G^\prime$) values for the Au/butanethiol/toluene interface at
509 +        different temperatures using a range of energy fluxes.}
510 +      
511 +      \begin{tabular}{cccc}
512 +        \hline\hline
513 +        $\langle T\rangle$ & $J_z$ & $G$ & $G^\prime$ \\
514 +        (K) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
515 +        \hline
516 +        200 & 1.86 & 180 & 135 \\
517 +            & 2.15 & 204 & 113 \\
518 +            & 3.93 & 175 & 114 \\
519 +        300 & 1.91 & 143 & 125 \\
520 +            & 4.19 & 134 & 113 \\
521 +        \hline\hline
522 +      \end{tabular}
523 +      \label{AuThiolToluene}
524 +    \end{center}
525 +  \end{minipage}
526 + \end{table*}
527 +
528 + Besides lower interfacial thermal conductance, surfaces in relatively
529 + high temperatures are susceptible to reconstructions, when
530 + butanethiols have a full coverage on the Au(111) surface. These
531 + reconstructions include surface Au atoms migrated outward to the S
532 + atom layer, and butanethiol molecules embedded into the original
533 + surface Au layer. The driving force for this behavior is the strong
534 + Au-S interactions in our simulations. And these reconstructions lead
535 + to higher ratio of Au-S attraction and thus is energetically
536 + favorable. Furthermore, this phenomenon agrees with experimental
537 + results\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
538 + {\it et al.} had kept their Au(111) slab rigid so that their
539 + simulations can reach 300K without surface reconstructions. Without
540 + this practice, simulating 100\% thiol covered interfaces under higher
541 + temperatures could hardly avoid surface reconstructions. However, our
542 + measurement is based on assuming homogeneity on $x$ and $y$ dimensions
543 + so that measurement of $T$ at particular $z$ would be an effective
544 + average of the particles of the same type. Since surface
545 + reconstructions could eliminate the original $x$ and $y$ dimensional
546 + homogeneity, measurement of $G$ is more difficult to conduct under
547 + higher temperatures. Therefore, most of our measurements are
548 + undertaken at $<T>\sim$200K.
549 +
550 + However, when the surface is not completely covered by butanethiols,
551 + the simulated system is more resistent to the reconstruction
552 + above. Our Au-butanethiol/toluene system did not see this phenomena
553 + even at $<T>\sim$300K. The Au(111) surfaces have a 90\% coverage of
554 + butanethiols and have empty three-fold sites. These empty sites could
555 + help prevent surface reconstruction in that they provide other means
556 + of capping agent relaxation. It is observed that butanethiols can
557 + migrate to their neighbor empty sites during a simulation. Therefore,
558 + we were able to obtain $G$'s for these interfaces even at a relatively
559 + high temperature without being affected by surface reconstructions.
560 +
561 + \subsection{Influence of Capping Agent Coverage on $G$}
562 + To investigate the influence of butanethiol coverage on interfacial
563 + thermal conductance, a series of different coverage Au-butanethiol
564 + surfaces is prepared and solvated with various organic
565 + molecules. These systems are then equilibrated and their interfacial
566 + thermal conductivity are measured with our NIVS algorithm. Table
567 + \ref{tlnUhxnUhxnD} lists these results for direct comparison between
568 + different coverages of butanethiol.
569 +
570 + With high coverage of butanethiol on the gold surface,
571 + the interfacial thermal conductance is enhanced
572 + significantly. Interestingly, a slightly lower butanethiol coverage
573 + leads to a moderately higher conductivity. This is probably due to
574 + more solvent/capping agent contact when butanethiol molecules are
575 + not densely packed, which enhances the interactions between the two
576 + phases and lowers the thermal transfer barrier of this interface.
577 + [COMPARE TO AU/WATER IN PAPER]
578 +
579 +
580 + significant conductance enhancement compared to the gold/water
581 + interface without capping agent and agree with available experimental
582 + data. This indicates that the metal-metal potential, though not
583 + predicting an accurate bulk metal thermal conductivity, does not
584 + greatly interfere with the simulation of the thermal conductance
585 + behavior across a non-metal interface.
586 + The results show that the two definitions used for $G$ yield
587 + comparable values, though $G^\prime$ tends to be smaller.
588 +
589 +
590 + \begin{table*}
591 +  \begin{minipage}{\linewidth}
592 +    \begin{center}
593 +      \caption{Computed interfacial thermal conductivity ($G$ and
594 +        $G^\prime$) values for the Au/butanethiol/hexane interface
595 +        with united-atom model and different capping agent coverage
596 +        and solvent molecule numbers at different temperatures using a
597 +        range of energy fluxes.}
598 +      
599 +      \begin{tabular}{cccccc}
600 +        \hline\hline
601 +        Thiol & $\langle T\rangle$ & & $J_z$ & $G$ & $G^\prime$ \\
602 +        coverage (\%) & (K) & $N_{hexane}$ & (GW/m$^2$) &
603 +        \multicolumn{2}{c}{(MW/m$^2$/K)} \\
604 +        \hline
605 +        0.0   & 200 & 200 & 0.96 & 43.3 & 42.7 \\
606 +              &     &     & 1.91 & 45.7 & 42.9 \\
607 +              &     & 166 & 0.96 & 43.1 & 53.4 \\
608 +        88.9  & 200 & 166 & 1.94 & 172  & 108  \\
609 +        100.0 & 250 & 200 & 0.96 & 81.8 & 67.0 \\
610 +              &     & 166 & 0.98 & 79.0 & 62.9 \\
611 +              &     &     & 1.44 & 76.2 & 64.8 \\
612 +              & 200 & 200 & 1.92 & 129  & 87.3 \\
613 +              &     &     & 1.93 & 131  & 77.5 \\
614 +              &     & 166 & 0.97 & 115  & 69.3 \\
615 +              &     &     & 1.94 & 125  & 87.1 \\
616 +        \hline\hline
617 +      \end{tabular}
618 +      \label{tlnUhxnUhxnD}
619 +    \end{center}
620 +  \end{minipage}
621 + \end{table*}
622 +
623 + \subsection{Influence of Chosen Molecule Model on $G$}
624 + [MAY COMBINE W MECHANISM STUDY]
625 +
626 + For the all-atom model, the liquid hexane phase was not stable under NPT
627 + conditions. Therefore, the simulation length scale parameters are
628 + adopted from previous equilibration results of the united-atom model
629 + at 200K. Table \ref{AuThiolHexaneAA} shows the results of these
630 + simulations. The conductivity values calculated with full capping
631 + agent coverage are substantially larger than observed in the
632 + united-atom model, and is even higher than predicted by
633 + experiments. It is possible that our parameters for metal-non-metal
634 + particle interactions lead to an overestimate of the interfacial
635 + thermal conductivity, although the active C-H vibrations in the
636 + all-atom model (which should not be appreciably populated at normal
637 + temperatures) could also account for this high conductivity. The major
638 + thermal transfer barrier of Au/butanethiol/hexane interface is between
639 + the liquid phase and the capping agent, so extra degrees of freedom
640 + such as the C-H vibrations could enhance heat exchange between these
641 + two phases and result in a much higher conductivity.
642 +
643 + \begin{table*}
644 +  \begin{minipage}{\linewidth}
645 +    \begin{center}
646 +      
647 +      \caption{Computed interfacial thermal conductivity ($G$ and
648 +        $G^\prime$) values for the Au/butanethiol/hexane interface
649 +        with all-atom model and different capping agent coverage at
650 +        200K using a range of energy fluxes.}
651 +      
652 +      \begin{tabular}{cccc}
653 +        \hline\hline
654 +        Thiol & $J_z$ & $G$ & $G^\prime$ \\
655 +        coverage (\%) & (GW/m$^2$) & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
656 +        \hline
657 +        0.0   & 0.95 & 28.5 & 27.2 \\
658 +              & 1.88 & 30.3 & 28.9 \\
659 +        100.0 & 2.87 & 551  & 294  \\
660 +              & 3.81 & 494  & 193  \\
661 +        \hline\hline
662 +      \end{tabular}
663 +      \label{AuThiolHexaneAA}
664 +    \end{center}
665 +  \end{minipage}
666 + \end{table*}
667 +
668 +
669 + \subsection{Mechanism of Interfacial Thermal Conductance Enhancement
670 +  by Capping Agent}
671 + [MAY INTRODUCE PROTOCOL IN METHODOLOGY/COMPUTATIONAL DETAIL]
672 +
673 +
674 + %subsubsection{Vibrational spectrum study on conductance mechanism}
675 + To investigate the mechanism of this interfacial thermal conductance,
676 + the vibrational spectra of various gold systems were obtained and are
677 + shown as in the upper panel of Fig. \ref{vibration}. To obtain these
678 + spectra, one first runs a simulation in the NVE ensemble and collects
679 + snapshots of configurations; these configurations are used to compute
680 + the velocity auto-correlation functions, which is used to construct a
681 + power spectrum via a Fourier transform. The gold surfaces covered by
682 + butanethiol molecules exhibit an additional peak observed at a
683 + frequency of $\sim$170cm$^{-1}$, which is attributed to the vibration
684 + of the S-Au bond. This vibration enables efficient thermal transport
685 + from surface Au atoms to the capping agents. Simultaneously, as shown
686 + in the lower panel of Fig. \ref{vibration}, the large overlap of the
687 + vibration spectra of butanethiol and hexane in the all-atom model,
688 + including the C-H vibration, also suggests high thermal exchange
689 + efficiency. The combination of these two effects produces the drastic
690 + interfacial thermal conductance enhancement in the all-atom model.
691 +
692 + \begin{figure}
693 + \includegraphics[width=\linewidth]{vibration}
694 + \caption{Vibrational spectra obtained for gold in different
695 +  environments (upper panel) and for Au/thiol/hexane simulation in
696 +  all-atom model (lower panel).}
697 + \label{vibration}
698 + \end{figure}
699 + % MAY NEED TO CONVERT TO JPEG
700 +
701 + \section{Conclusions}
702 +
703 +
704 + [NECESSITY TO STUDY THERMAL CONDUCTANCE IN NANOCRYSTAL STRUCTURE]\cite{vlugt:cpc2007154}
705 +
706   \section{Acknowledgments}
707   Support for this project was provided by the National Science
708   Foundation under grant CHE-0848243. Computational time was provided by
709   the Center for Research Computing (CRC) at the University of Notre
710 < Dame.  \newpage
710 > Dame. \newpage
711  
712   \bibliography{interfacial}
713  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines