ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/interfacial/interfacial.tex
Revision: 3727
Committed: Fri Jun 24 16:59:37 2011 UTC (13 years, 2 months ago) by skuang
Content type: application/x-tex
File size: 22881 byte(s)
Log Message:
add data. done much of methodology and simulation details.

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