ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3769
Committed: Tue Nov 15 17:07:03 2011 UTC (12 years, 10 months ago) by gezelter
Content type: application/x-tex
File size: 49707 byte(s)
Log Message:
Initial

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 With the Non-Isotropic Velocity Scaling (NIVS) approach to Reverse
48 Non-Equilibrium Molecular Dynamics (RNEMD) it is possible to impose
49 an unphysical thermal flux between different regions of
50 inhomogeneous systems such as solid / liquid interfaces. We have
51 applied NIVS to compute the interfacial thermal conductance at a
52 metal / organic solvent interface that has been chemically capped by
53 butanethiol molecules. Our calculations suggest that coupling
54 between the metal and liquid phases is enhanced by the capping
55 agents, leading to a greatly enhanced conductivity at the interface.
56 Specifically, the chemical bond between the metal and the capping
57 agent introduces a vibrational overlap that is not present without
58 the capping agent, and the overlap between the vibrational spectra
59 (metal to cap, cap to solvent) provides a mechanism for rapid
60 thermal transport across the interface. Our calculations also
61 suggest that this is a non-monotonic function of the fractional
62 coverage of the surface, as moderate coverages allow diffusive heat
63 transport of solvent molecules that have been in close contact with
64 the capping agent.
65
66 {\bf Keywords: non-equilibrium, molecular dynamics, vibrational
67 overlap, coverage dependent.}
68 \end{abstract}
69
70 \newpage
71
72 %\narrowtext
73
74 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
75 % BODY OF TEXT
76 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77
78 \section{Introduction}
79 Due to the importance of heat flow (and heat removal) in
80 nanotechnology, interfacial thermal conductance has been studied
81 extensively both experimentally and computationally.\cite{cahill:793}
82 Nanoscale materials have a significant fraction of their atoms at
83 interfaces, and the chemical details of these interfaces govern the
84 thermal transport properties. Furthermore, the interfaces are often
85 heterogeneous (e.g. solid - liquid), which provides a challenge to
86 computational methods which have been developed for homogeneous or
87 bulk systems.
88
89 Experimentally, the thermal properties of a number of interfaces have
90 been investigated. Cahill and coworkers studied nanoscale thermal
91 transport from metal nanoparticle/fluid interfaces, to epitaxial
92 TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
93 interfaces between water and solids with different self-assembled
94 monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
95 Wang {\it et al.} studied heat transport through long-chain
96 hydrocarbon monolayers on gold substrate at individual molecular
97 level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
98 cetyltrimethylammonium bromide (CTAB) on the thermal transport between
99 gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
100 et al.} studied the cooling dynamics, which is controlled by thermal
101 interface resistance of glass-embedded metal
102 nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
103 normally considered barriers for heat transport, Alper {\it et al.}
104 suggested that specific ligands (capping agents) could completely
105 eliminate this barrier
106 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
107
108 The acoustic mismatch model for interfacial conductance utilizes the
109 acoustic impedance ($Z_a = \rho_a v^s_a$) on both sides of the
110 interface.\cite{swartz1989} Here, $\rho_a$ and $v^s_a$ are the density
111 and speed of sound in material $a$. The phonon transmission
112 probability at the $a-b$ interface is
113 \begin{equation}
114 t_{ab} = \frac{4 Z_a Z_b}{\left(Z_a + Z_b \right)^2},
115 \end{equation}
116 and the interfacial conductance can then be approximated as
117 \begin{equation}
118 G_{ab} \approx \frac{1}{4} C_D v_D t_{ab}
119 \end{equation}
120 where $C_D$ is the Debye heat capacity of the hot side, and $v_D$ is
121 the Debye phonon velocity ($1/v_D^3 = 1/3v_L^3 + 2/3 v_T^3$) where
122 $v_L$ and $v_T$ are the longitudinal and transverse speeds of sound,
123 respectively. For the Au/hexane and Au/toluene interfaces, the
124 acoustic mismatch model predicts room-temperature $G \approx 87 \mbox{
125 and } 129$ MW m$^{-2}$ K$^{-1}$, respectively. However, it is not
126 clear how to apply the acoustic mismatch model to a
127 chemically-modified surface, particularly when the acoustic properties
128 of a monolayer film may not be well characterized.
129
130 More precise computational models have also been used to study the
131 interfacial thermal transport in order to gain an understanding of
132 this phenomena at the molecular level. Recently, Hase and coworkers
133 employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
134 study thermal transport from hot Au(111) substrate to a self-assembled
135 monolayer of alkylthiol with relatively long chain (8-20 carbon
136 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
137 measurements for heat conductance of interfaces between the capping
138 monolayer on Au and a solvent phase have yet to be studied with their
139 approach. The comparatively low thermal flux through interfaces is
140 difficult to measure with Equilibrium
141 MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
142 methods. Therefore, the Reverse NEMD (RNEMD)
143 methods\cite{MullerPlathe:1997xw,kuang:164101} would be advantageous
144 in that they {\it apply} the difficult to measure quantity (flux),
145 while {\it measuring} the easily-computed quantity (the thermal
146 gradient). This is particularly true for inhomogeneous interfaces
147 where it would not be clear how to apply a gradient {\it a priori}.
148 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
149 this approach to various liquid interfaces and studied how thermal
150 conductance (or resistance) is dependent on chemical details of a
151 number of hydrophobic and hydrophilic aqueous interfaces. And
152 recently, Luo {\it et al.} studied the thermal conductance of
153 Au-SAM-Au junctions using the same approach, comparing to a constant
154 temperature difference method.\cite{Luo20101} While this latter
155 approach establishes more ideal Maxwell-Boltzmann distributions than
156 previous RNEMD methods, it does not guarantee momentum or kinetic
157 energy conservation.
158
159 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
160 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
161 retains the desirable features of RNEMD (conservation of linear
162 momentum and total energy, compatibility with periodic boundary
163 conditions) while establishing true thermal distributions in each of
164 the two slabs. Furthermore, it allows effective thermal exchange
165 between particles of different identities, and thus makes the study of
166 interfacial conductance much simpler.
167
168 The work presented here deals with the Au(111) surface covered to
169 varying degrees by butanethiol, a capping agent with short carbon
170 chain, and solvated with organic solvents of different molecular
171 properties. Different models were used for both the capping agent and
172 the solvent force field parameters. Using the NIVS algorithm, the
173 thermal transport across these interfaces was studied and the
174 underlying mechanism for the phenomena was investigated.
175
176 \section{Methodology}
177 \subsection{Imposed-Flux Methods in MD Simulations}
178 Steady state MD simulations have an advantage in that not many
179 trajectories are needed to study the relationship between thermal flux
180 and thermal gradients. For systems with low interfacial conductance,
181 one must have a method capable of generating or measuring relatively
182 small fluxes, compared to those required for bulk conductivity. This
183 requirement makes the calculation even more difficult for
184 slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
185 NEMD methods impose a gradient (and measure a flux), but at interfaces
186 it is not clear what behavior should be imposed at the boundaries
187 between materials. Imposed-flux reverse non-equilibrium
188 methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
189 the thermal response becomes an easy-to-measure quantity. Although
190 M\"{u}ller-Plathe's original momentum swapping approach can be used
191 for exchanging energy between particles of different identity, the
192 kinetic energy transfer efficiency is affected by the mass difference
193 between the particles, which limits its application on heterogeneous
194 interfacial systems.
195
196 The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
197 to non-equilibrium MD simulations is able to impose a wide range of
198 kinetic energy fluxes without obvious perturbation to the velocity
199 distributions of the simulated systems. Furthermore, this approach has
200 the advantage in heterogeneous interfaces in that kinetic energy flux
201 can be applied between regions of particles of arbitrary identity, and
202 the flux will not be restricted by difference in particle mass.
203
204 The NIVS algorithm scales the velocity vectors in two separate regions
205 of a simulation system with respective diagonal scaling matrices. To
206 determine these scaling factors in the matrices, a set of equations
207 including linear momentum conservation and kinetic energy conservation
208 constraints and target energy flux satisfaction is solved. With the
209 scaling operation applied to the system in a set frequency, bulk
210 temperature gradients can be easily established, and these can be used
211 for computing thermal conductivities. The NIVS algorithm conserves
212 momenta and energy and does not depend on an external thermostat.
213
214 \subsection{Defining Interfacial Thermal Conductivity ($G$)}
215
216 For an interface with relatively low interfacial conductance, and a
217 thermal flux between two distinct bulk regions, the regions on either
218 side of the interface rapidly come to a state in which the two phases
219 have relatively homogeneous (but distinct) temperatures. The
220 interfacial thermal conductivity $G$ can therefore be approximated as:
221 \begin{equation}
222 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
223 \langle T_\mathrm{cold}\rangle \right)}
224 \label{lowG}
225 \end{equation}
226 where ${E_{total}}$ is the total imposed non-physical kinetic energy
227 transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
228 and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
229 temperature of the two separated phases. For an applied flux $J_z$
230 operating over a simulation time $t$ on a periodically-replicated slab
231 of dimensions $L_x \times L_y$, $E_{total} = 2 J_z t L_x L_y$.
232
233 When the interfacial conductance is {\it not} small, there are two
234 ways to define $G$. One common way is to assume the temperature is
235 discrete on the two sides of the interface. $G$ can be calculated
236 using the applied thermal flux $J$ and the maximum temperature
237 difference measured along the thermal gradient max($\Delta T$), which
238 occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
239 known as the Kapitza conductance, which is the inverse of the Kapitza
240 resistance.
241 \begin{equation}
242 G=\frac{J}{\Delta T}
243 \label{discreteG}
244 \end{equation}
245
246 \begin{figure}
247 \includegraphics[width=\linewidth]{method}
248 \caption{Interfacial conductance can be calculated by applying an
249 (unphysical) kinetic energy flux between two slabs, one located
250 within the metal and another on the edge of the periodic box. The
251 system responds by forming a thermal gradient. In bulk liquids,
252 this gradient typically has a single slope, but in interfacial
253 systems, there are distinct thermal conductivity domains. The
254 interfacial conductance, $G$ is found by measuring the temperature
255 gap at the Gibbs dividing surface, or by using second derivatives of
256 the thermal profile.}
257 \label{demoPic}
258 \end{figure}
259
260 Another approach is to assume that the temperature is continuous and
261 differentiable throughout the space. Given that $\lambda$ is also
262 differentiable, $G$ can be defined as its gradient ($\nabla\lambda$)
263 projected along a vector normal to the interface ($\mathbf{\hat{n}}$)
264 and evaluated at the interface location ($z_0$). This quantity,
265 \begin{align}
266 G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
267 &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
268 \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
269 &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
270 \left(\frac{\partial T}{\partial z}\right)_{z_0}^2 \label{derivativeG}
271 \end{align}
272 has the same units as the common definition for $G$, and the maximum
273 of its magnitude denotes where thermal conductivity has the largest
274 change, i.e. the interface. In the geometry used in this study, the
275 vector normal to the interface points along the $z$ axis, as do
276 $\vec{J}$ and the thermal gradient. This yields the simplified
277 expressions in Eq. \ref{derivativeG}.
278
279 With temperature profiles obtained from simulation, one is able to
280 approximate the first and second derivatives of $T$ with finite
281 difference methods and calculate $G^\prime$. In what follows, both
282 definitions have been used, and are compared in the results.
283
284 To investigate the interfacial conductivity at metal / solvent
285 interfaces, we have modeled a metal slab with its (111) surfaces
286 perpendicular to the $z$-axis of our simulation cells. The metal slab
287 has been prepared both with and without capping agents on the exposed
288 surface, and has been solvated with simple organic solvents, as
289 illustrated in Figure \ref{gradT}.
290
291 With the simulation cell described above, we are able to equilibrate
292 the system and impose an unphysical thermal flux between the liquid
293 and the metal phase using the NIVS algorithm. By periodically applying
294 the unphysical flux, we obtained a temperature profile and its spatial
295 derivatives. Figure \ref{gradT} shows how an applied thermal flux can
296 be used to obtain the 1st and 2nd derivatives of the temperature
297 profile.
298
299 \begin{figure}
300 \includegraphics[width=\linewidth]{gradT}
301 \caption{A sample of Au (111) / butanethiol / hexane interfacial
302 system with the temperature profile after a kinetic energy flux has
303 been imposed. Note that the largest temperature jump in the thermal
304 profile (corresponding to the lowest interfacial conductance) is at
305 the interface between the butanethiol molecules (blue) and the
306 solvent (grey). First and second derivatives of the temperature
307 profile are obtained using a finite difference approximation (lower
308 panel).}
309 \label{gradT}
310 \end{figure}
311
312 \section{Computational Details}
313 \subsection{Simulation Protocol}
314 The NIVS algorithm has been implemented in our MD simulation code,
315 OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
316 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
317 under atmospheric pressure (1 atm) and 200K. After equilibration,
318 butanethiol capping agents were placed at three-fold hollow sites on
319 the Au(111) surfaces. These sites are either {\it fcc} or {\it
320 hcp} sites, although Hase {\it et al.} found that they are
321 equivalent in a heat transfer process,\cite{hase:2010} so we did not
322 distinguish between these sites in our study. The maximum butanethiol
323 capacity on Au surface is $1/3$ of the total number of surface Au
324 atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
325 structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
326 series of lower coverages was also prepared by eliminating
327 butanethiols from the higher coverage surface in a regular manner. The
328 lower coverages were prepared in order to study the relation between
329 coverage and interfacial conductance.
330
331 The capping agent molecules were allowed to migrate during the
332 simulations. They distributed themselves uniformly and sampled a
333 number of three-fold sites throughout out study. Therefore, the
334 initial configuration does not noticeably affect the sampling of a
335 variety of configurations of the same coverage, and the final
336 conductance measurement would be an average effect of these
337 configurations explored in the simulations.
338
339 After the modified Au-butanethiol surface systems were equilibrated in
340 the canonical (NVT) ensemble, organic solvent molecules were packed in
341 the previously empty part of the simulation cells.\cite{packmol} Two
342 solvents were investigated, one which has little vibrational overlap
343 with the alkanethiol and which has a planar shape (toluene), and one
344 which has similar vibrational frequencies to the capping agent and
345 chain-like shape ({\it n}-hexane).
346
347 The simulation cells were not particularly extensive along the
348 $z$-axis, as a very long length scale for the thermal gradient may
349 cause excessively hot or cold temperatures in the middle of the
350 solvent region and lead to undesired phenomena such as solvent boiling
351 or freezing when a thermal flux is applied. Conversely, too few
352 solvent molecules would change the normal behavior of the liquid
353 phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
354 these extreme cases did not happen to our simulations. The spacing
355 between periodic images of the gold interfaces is $45 \sim 75$\AA in
356 our simulations.
357
358 The initial configurations generated are further equilibrated with the
359 $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
360 change. This is to ensure that the equilibration of liquid phase does
361 not affect the metal's crystalline structure. Comparisons were made
362 with simulations that allowed changes of $L_x$ and $L_y$ during NPT
363 equilibration. No substantial changes in the box geometry were noticed
364 in these simulations. After ensuring the liquid phase reaches
365 equilibrium at atmospheric pressure (1 atm), further equilibration was
366 carried out under canonical (NVT) and microcanonical (NVE) ensembles.
367
368 After the systems reach equilibrium, NIVS was used to impose an
369 unphysical thermal flux between the metal and the liquid phases. Most
370 of our simulations were done under an average temperature of
371 $\sim$200K. Therefore, thermal flux usually came from the metal to the
372 liquid so that the liquid has a higher temperature and would not
373 freeze due to lowered temperatures. After this induced temperature
374 gradient had stabilized, the temperature profile of the simulation cell
375 was recorded. To do this, the simulation cell is divided evenly into
376 $N$ slabs along the $z$-axis. The average temperatures of each slab
377 are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
378 the same, the derivatives of $T$ with respect to slab number $n$ can
379 be directly used for $G^\prime$ calculations: \begin{equation}
380 G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
381 \Big/\left(\frac{\partial T}{\partial z}\right)^2
382 = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
383 \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
384 = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
385 \Big/\left(\frac{\partial T}{\partial n}\right)^2
386 \label{derivativeG2}
387 \end{equation}
388 The absolute values in Eq. \ref{derivativeG2} appear because the
389 direction of the flux $\vec{J}$ is in an opposing direction on either
390 side of the metal slab.
391
392 All of the above simulation procedures use a time step of 1 fs. Each
393 equilibration stage took a minimum of 100 ps, although in some cases,
394 longer equilibration stages were utilized.
395
396 \subsection{Force Field Parameters}
397 Our simulations include a number of chemically distinct components.
398 Figure \ref{demoMol} demonstrates the sites defined for both
399 United-Atom and All-Atom models of the organic solvent and capping
400 agents in our simulations. Force field parameters are needed for
401 interactions both between the same type of particles and between
402 particles of different species.
403
404 \begin{figure}
405 \includegraphics[width=\linewidth]{structures}
406 \caption{Structures of the capping agent and solvents utilized in
407 these simulations. The chemically-distinct sites (a-e) are expanded
408 in terms of constituent atoms for both United Atom (UA) and All Atom
409 (AA) force fields. Most parameters are from References
410 \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
411 (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
412 atoms are given in Table 1 in the supporting information.}
413 \label{demoMol}
414 \end{figure}
415
416 The Au-Au interactions in metal lattice slab is described by the
417 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
418 potentials include zero-point quantum corrections and are
419 reparametrized for accurate surface energies compared to the
420 Sutton-Chen potentials.\cite{Chen90}
421
422 For the two solvent molecules, {\it n}-hexane and toluene, two
423 different atomistic models were utilized. Both solvents were modeled
424 using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
425 parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
426 for our UA solvent molecules. In these models, sites are located at
427 the carbon centers for alkyl groups. Bonding interactions, including
428 bond stretches and bends and torsions, were used for intra-molecular
429 sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
430 potentials are used.
431
432 By eliminating explicit hydrogen atoms, the TraPPE-UA models are
433 simple and computationally efficient, while maintaining good accuracy.
434 However, the TraPPE-UA model for alkanes is known to predict a slightly
435 lower boiling point than experimental values. This is one of the
436 reasons we used a lower average temperature (200K) for our
437 simulations. If heat is transferred to the liquid phase during the
438 NIVS simulation, the liquid in the hot slab can actually be
439 substantially warmer than the mean temperature in the simulation. The
440 lower mean temperatures therefore prevent solvent boiling.
441
442 For UA-toluene, the non-bonded potentials between intermolecular sites
443 have a similar Lennard-Jones formulation. The toluene molecules were
444 treated as a single rigid body, so there was no need for
445 intramolecular interactions (including bonds, bends, or torsions) in
446 this solvent model.
447
448 Besides the TraPPE-UA models, AA models for both organic solvents are
449 included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
450 were used. For hexane, additional explicit hydrogen sites were
451 included. Besides bonding and non-bonded site-site interactions,
452 partial charges and the electrostatic interactions were added to each
453 CT and HC site. For toluene, a flexible model for the toluene molecule
454 was utilized which included bond, bend, torsion, and inversion
455 potentials to enforce ring planarity.
456
457 The butanethiol capping agent in our simulations, were also modeled
458 with both UA and AA model. The TraPPE-UA force field includes
459 parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
460 UA butanethiol model in our simulations. The OPLS-AA also provides
461 parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
462 surfaces do not have the hydrogen atom bonded to sulfur. To derive
463 suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
464 adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
465 modify the parameters for the CTS atom to maintain charge neutrality
466 in the molecule. Note that the model choice (UA or AA) for the capping
467 agent can be different from the solvent. Regardless of model choice,
468 the force field parameters for interactions between capping agent and
469 solvent can be derived using Lorentz-Berthelot Mixing Rule:
470 \begin{eqnarray}
471 \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
472 \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
473 \end{eqnarray}
474
475 To describe the interactions between metal (Au) and non-metal atoms,
476 we refer to an adsorption study of alkyl thiols on gold surfaces by
477 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
478 Lennard-Jones form of potential parameters for the interaction between
479 Au and pseudo-atoms CH$_x$ and S based on a well-established and
480 widely-used effective potential of Hautman and Klein for the Au(111)
481 surface.\cite{hautman:4994} As our simulations require the gold slab
482 to be flexible to accommodate thermal excitation, the pair-wise form
483 of potentials they developed was used for our study.
484
485 The potentials developed from {\it ab initio} calculations by Leng
486 {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
487 interactions between Au and aromatic C/H atoms in toluene. However,
488 the Lennard-Jones parameters between Au and other types of particles,
489 (e.g. AA alkanes) have not yet been established. For these
490 interactions, the Lorentz-Berthelot mixing rule can be used to derive
491 effective single-atom LJ parameters for the metal using the fit values
492 for toluene. These are then used to construct reasonable mixing
493 parameters for the interactions between the gold and other atoms.
494 Table 1 in the supporting information summarizes the
495 ``metal/non-metal'' parameters utilized in our simulations.
496
497 \section{Results}
498 There are many factors contributing to the measured interfacial
499 conductance; some of these factors are physically motivated
500 (e.g. coverage of the surface by the capping agent coverage and
501 solvent identity), while some are governed by parameters of the
502 methodology (e.g. applied flux and the formulas used to obtain the
503 conductance). In this section we discuss the major physical and
504 calculational effects on the computed conductivity.
505
506 \subsection{Effects due to capping agent coverage}
507
508 A series of different initial conditions with a range of surface
509 coverages was prepared and solvated with various with both of the
510 solvent molecules. These systems were then equilibrated and their
511 interfacial thermal conductivity was measured with the NIVS
512 algorithm. Figure \ref{coverage} demonstrates the trend of conductance
513 with respect to surface coverage.
514
515 \begin{figure}
516 \includegraphics[width=\linewidth]{coverage}
517 \caption{The interfacial thermal conductivity ($G$) has a
518 non-monotonic dependence on the degree of surface capping. This
519 data is for the Au(111) / butanethiol / solvent interface with
520 various UA force fields at $\langle T\rangle \sim $200K.}
521 \label{coverage}
522 \end{figure}
523
524 In partially covered surfaces, the derivative definition for
525 $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
526 location of maximum change of $\lambda$ becomes washed out. The
527 discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
528 Gibbs dividing surface is still well-defined. Therefore, $G$ (not
529 $G^\prime$) was used in this section.
530
531 From Figure \ref{coverage}, one can see the significance of the
532 presence of capping agents. When even a small fraction of the Au(111)
533 surface sites are covered with butanethiols, the conductivity exhibits
534 an enhancement by at least a factor of 3. Capping agents are clearly
535 playing a major role in thermal transport at metal / organic solvent
536 surfaces.
537
538 We note a non-monotonic behavior in the interfacial conductance as a
539 function of surface coverage. The maximum conductance (largest $G$)
540 happens when the surfaces are about 75\% covered with butanethiol
541 caps. The reason for this behavior is not entirely clear. One
542 explanation is that incomplete butanethiol coverage allows small gaps
543 between butanethiols to form. These gaps can be filled by transient
544 solvent molecules. These solvent molecules couple very strongly with
545 the hot capping agent molecules near the surface, and can then carry
546 away (diffusively) the excess thermal energy from the surface.
547
548 There appears to be a competition between the conduction of the
549 thermal energy away from the surface by the capping agents (enhanced
550 by greater coverage) and the coupling of the capping agents with the
551 solvent (enhanced by interdigitation at lower coverages). This
552 competition would lead to the non-monotonic coverage behavior observed
553 here.
554
555 Results for rigid body toluene solvent, as well as the UA hexane, are
556 within the ranges expected from prior experimental
557 work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
558 that explicit hydrogen atoms might not be required for modeling
559 thermal transport in these systems. C-H vibrational modes do not see
560 significant excited state population at low temperatures, and are not
561 likely to carry lower frequency excitations from the solid layer into
562 the bulk liquid.
563
564 The toluene solvent does not exhibit the same behavior as hexane in
565 that $G$ remains at approximately the same magnitude when the capping
566 coverage increases from 25\% to 75\%. Toluene, as a rigid planar
567 molecule, cannot occupy the relatively small gaps between the capping
568 agents as easily as the chain-like {\it n}-hexane. The effect of
569 solvent coupling to the capping agent is therefore weaker in toluene
570 except at the very lowest coverage levels. This effect counters the
571 coverage-dependent conduction of heat away from the metal surface,
572 leading to a much flatter $G$ vs. coverage trend than is observed in
573 {\it n}-hexane.
574
575 \subsection{Effects due to Solvent \& Solvent Models}
576 In addition to UA solvent and capping agent models, AA models have
577 also been included in our simulations. In most of this work, the same
578 (UA or AA) model for solvent and capping agent was used, but it is
579 also possible to utilize different models for different components.
580 We have also included isotopic substitutions (Hydrogen to Deuterium)
581 to decrease the explicit vibrational overlap between solvent and
582 capping agent. Table \ref{modelTest} summarizes the results of these
583 studies.
584
585 \begin{table*}
586 \begin{minipage}{\linewidth}
587 \begin{center}
588
589 \caption{Computed interfacial thermal conductance ($G$ and
590 $G^\prime$) values for interfaces using various models for
591 solvent and capping agent (or without capping agent) at
592 $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
593 solvent or capping agent molecules. Error estimates are
594 indicated in parentheses.}
595
596 \begin{tabular}{llccc}
597 \hline\hline
598 Butanethiol model & Solvent & $G$ & $G^\prime$ \\
599 (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
600 \hline
601 UA & UA hexane & 131(9) & 87(10) \\
602 & UA hexane(D) & 153(5) & 136(13) \\
603 & AA hexane & 131(6) & 122(10) \\
604 & UA toluene & 187(16) & 151(11) \\
605 & AA toluene & 200(36) & 149(53) \\
606 \hline
607 AA & UA hexane & 116(9) & 129(8) \\
608 & AA hexane & 442(14) & 356(31) \\
609 & AA hexane(D) & 222(12) & 234(54) \\
610 & UA toluene & 125(25) & 97(60) \\
611 & AA toluene & 487(56) & 290(42) \\
612 \hline
613 AA(D) & UA hexane & 158(25) & 172(4) \\
614 & AA hexane & 243(29) & 191(11) \\
615 & AA toluene & 364(36) & 322(67) \\
616 \hline
617 bare & UA hexane & 46.5(3.2) & 49.4(4.5) \\
618 & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
619 & AA hexane & 31.0(1.4) & 29.4(1.3) \\
620 & UA toluene & 70.1(1.3) & 65.8(0.5) \\
621 \hline\hline
622 \end{tabular}
623 \label{modelTest}
624 \end{center}
625 \end{minipage}
626 \end{table*}
627
628 To facilitate direct comparison between force fields, systems with the
629 same capping agent and solvent were prepared with the same length
630 scales for the simulation cells.
631
632 On bare metal / solvent surfaces, different force field models for
633 hexane yield similar results for both $G$ and $G^\prime$, and these
634 two definitions agree with each other very well. This is primarily an
635 indicator of weak interactions between the metal and the solvent.
636
637 For the fully-covered surfaces, the choice of force field for the
638 capping agent and solvent has a large impact on the calculated values
639 of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
640 much larger than their UA to UA counterparts, and these values exceed
641 the experimental estimates by a large measure. The AA force field
642 allows significant energy to go into C-H (or C-D) stretching modes,
643 and since these modes are high frequency, this non-quantum behavior is
644 likely responsible for the overestimate of the conductivity. Compared
645 to the AA model, the UA model yields more reasonable conductivity
646 values with much higher computational efficiency.
647
648 \subsubsection{Are electronic excitations in the metal important?}
649 Because they lack electronic excitations, the QSC and related embedded
650 atom method (EAM) models for gold are known to predict unreasonably
651 low values for bulk conductivity
652 ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
653 conductance between the phases ($G$) is governed primarily by phonon
654 excitation (and not electronic degrees of freedom), one would expect a
655 classical model to capture most of the interfacial thermal
656 conductance. Our results for $G$ and $G^\prime$ indicate that this is
657 indeed the case, and suggest that the modeling of interfacial thermal
658 transport depends primarily on the description of the interactions
659 between the various components at the interface. When the metal is
660 chemically capped, the primary barrier to thermal conductivity appears
661 to be the interface between the capping agent and the surrounding
662 solvent, so the excitations in the metal have little impact on the
663 value of $G$.
664
665 \subsection{Effects due to methodology and simulation parameters}
666
667 We have varied the parameters of the simulations in order to
668 investigate how these factors would affect the computation of $G$. Of
669 particular interest are: 1) the length scale for the applied thermal
670 gradient (modified by increasing the amount of solvent in the system),
671 2) the sign and magnitude of the applied thermal flux, 3) the average
672 temperature of the simulation (which alters the solvent density during
673 equilibration), and 4) the definition of the interfacial conductance
674 (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
675 calculation.
676
677 Systems of different lengths were prepared by altering the number of
678 solvent molecules and extending the length of the box along the $z$
679 axis to accomodate the extra solvent. Equilibration at the same
680 temperature and pressure conditions led to nearly identical surface
681 areas ($L_x$ and $L_y$) available to the metal and capping agent,
682 while the extra solvent served mainly to lengthen the axis that was
683 used to apply the thermal flux. For a given value of the applied
684 flux, the different $z$ length scale has only a weak effect on the
685 computed conductivities.
686
687 \subsubsection{Effects of applied flux}
688 The NIVS algorithm allows changes in both the sign and magnitude of
689 the applied flux. It is possible to reverse the direction of heat
690 flow simply by changing the sign of the flux, and thermal gradients
691 which would be difficult to obtain experimentally ($5$ K/\AA) can be
692 easily simulated. However, the magnitude of the applied flux is not
693 arbitrary if one aims to obtain a stable and reliable thermal gradient.
694 A temperature gradient can be lost in the noise if $|J_z|$ is too
695 small, and excessive $|J_z|$ values can cause phase transitions if the
696 extremes of the simulation cell become widely separated in
697 temperature. Also, if $|J_z|$ is too large for the bulk conductivity
698 of the materials, the thermal gradient will never reach a stable
699 state.
700
701 Within a reasonable range of $J_z$ values, we were able to study how
702 $G$ changes as a function of this flux. In what follows, we use
703 positive $J_z$ values to denote the case where energy is being
704 transferred by the method from the metal phase and into the liquid.
705 The resulting gradient therefore has a higher temperature in the
706 liquid phase. Negative flux values reverse this transfer, and result
707 in higher temperature metal phases. The conductance measured under
708 different applied $J_z$ values is listed in Tables 2 and 3 in the
709 supporting information. These results do not indicate that $G$ depends
710 strongly on $J_z$ within this flux range. The linear response of flux
711 to thermal gradient simplifies our investigations in that we can rely
712 on $G$ measurement with only a small number $J_z$ values.
713
714 The sign of $J_z$ is a different matter, however, as this can alter
715 the temperature on the two sides of the interface. The average
716 temperature values reported are for the entire system, and not for the
717 liquid phase, so at a given $\langle T \rangle$, the system with
718 positive $J_z$ has a warmer liquid phase. This means that if the
719 liquid carries thermal energy via diffusive transport, {\it positive}
720 $J_z$ values will result in increased molecular motion on the liquid
721 side of the interface, and this will increase the measured
722 conductivity.
723
724 \subsubsection{Effects due to average temperature}
725
726 We also studied the effect of average system temperature on the
727 interfacial conductance. The simulations are first equilibrated in
728 the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
729 predict a lower boiling point (and liquid state density) than
730 experiments. This lower-density liquid phase leads to reduced contact
731 between the hexane and butanethiol, and this accounts for our
732 observation of lower conductance at higher temperatures. In raising
733 the average temperature from 200K to 250K, the density drop of
734 $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
735 conductance.
736
737 Similar behavior is observed in the TraPPE-UA model for toluene,
738 although this model has better agreement with the experimental
739 densities of toluene. The expansion of the toluene liquid phase is
740 not as significant as that of the hexane (8.3\% over 100K), and this
741 limits the effect to $\sim$20\% drop in thermal conductivity.
742
743 Although we have not mapped out the behavior at a large number of
744 temperatures, is clear that there will be a strong temperature
745 dependence in the interfacial conductance when the physical properties
746 of one side of the interface (notably the density) change rapidly as a
747 function of temperature.
748
749 Besides the lower interfacial thermal conductance, surfaces at
750 relatively high temperatures are susceptible to reconstructions,
751 particularly when butanethiols fully cover the Au(111) surface. These
752 reconstructions include surface Au atoms which migrate outward to the
753 S atom layer, and butanethiol molecules which embed into the surface
754 Au layer. The driving force for this behavior is the strong Au-S
755 interactions which are modeled here with a deep Lennard-Jones
756 potential. This phenomenon agrees with reconstructions that have been
757 experimentally
758 observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
759 {\it et al.} kept their Au(111) slab rigid so that their simulations
760 could reach 300K without surface
761 reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
762 blur the interface, the measurement of $G$ becomes more difficult to
763 conduct at higher temperatures. For this reason, most of our
764 measurements are undertaken at $\langle T\rangle\sim$200K where
765 reconstruction is minimized.
766
767 However, when the surface is not completely covered by butanethiols,
768 the simulated system appears to be more resistent to the
769 reconstruction. Our Au / butanethiol / toluene system had the Au(111)
770 surfaces 90\% covered by butanethiols, but did not see this above
771 phenomena even at $\langle T\rangle\sim$300K. That said, we did
772 observe butanethiols migrating to neighboring three-fold sites during
773 a simulation. Since the interface persisted in these simulations, we
774 were able to obtain $G$'s for these interfaces even at a relatively
775 high temperature without being affected by surface reconstructions.
776
777 \section{Discussion}
778
779 The primary result of this work is that the capping agent acts as an
780 efficient thermal coupler between solid and solvent phases. One of
781 the ways the capping agent can carry out this role is to down-shift
782 between the phonon vibrations in the solid (which carry the heat from
783 the gold) and the molecular vibrations in the liquid (which carry some
784 of the heat in the solvent).
785
786 To investigate the mechanism of interfacial thermal conductance, the
787 vibrational power spectrum was computed. Power spectra were taken for
788 individual components in different simulations. To obtain these
789 spectra, simulations were run after equilibration in the
790 microcanonical (NVE) ensemble and without a thermal
791 gradient. Snapshots of configurations were collected at a frequency
792 that is higher than that of the fastest vibrations occurring in the
793 simulations. With these configurations, the velocity auto-correlation
794 functions can be computed:
795 \begin{equation}
796 C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
797 \label{vCorr}
798 \end{equation}
799 The power spectrum is constructed via a Fourier transform of the
800 symmetrized velocity autocorrelation function,
801 \begin{equation}
802 \hat{f}(\omega) =
803 \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
804 \label{fourier}
805 \end{equation}
806
807 \subsection{The role of specific vibrations}
808 The vibrational spectra for gold slabs in different environments are
809 shown as in Figure \ref{specAu}. Regardless of the presence of
810 solvent, the gold surfaces which are covered by butanethiol molecules
811 exhibit an additional peak observed at a frequency of
812 $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
813 vibration. This vibration enables efficient thermal coupling of the
814 surface Au layer to the capping agents. Therefore, in our simulations,
815 the Au / S interfaces do not appear to be the primary barrier to
816 thermal transport when compared with the butanethiol / solvent
817 interfaces. This supports the results of Luo {\it et
818 al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
819 twice as large as what we have computed for the thiol-liquid
820 interfaces.
821
822 \begin{figure}
823 \includegraphics[width=\linewidth]{vibration}
824 \caption{The vibrational power spectrum for thiol-capped gold has an
825 additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
826 surfaces (both with and without a solvent over-layer) are missing
827 this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
828 the vibrational power spectrum for the butanethiol capping agents.}
829 \label{specAu}
830 \end{figure}
831
832 Also in this figure, we show the vibrational power spectrum for the
833 bound butanethiol molecules, which also exhibits the same
834 $\sim$165cm$^{-1}$ peak.
835
836 \subsection{Overlap of power spectra}
837 A comparison of the results obtained from the two different organic
838 solvents can also provide useful information of the interfacial
839 thermal transport process. In particular, the vibrational overlap
840 between the butanethiol and the organic solvents suggests a highly
841 efficient thermal exchange between these components. Very high
842 thermal conductivity was observed when AA models were used and C-H
843 vibrations were treated classically. The presence of extra degrees of
844 freedom in the AA force field yields higher heat exchange rates
845 between the two phases and results in a much higher conductivity than
846 in the UA force field. The all-atom classical models include high
847 frequency modes which should be unpopulated at our relatively low
848 temperatures. This artifact is likely the cause of the high thermal
849 conductance in all-atom MD simulations.
850
851 The similarity in the vibrational modes available to solvent and
852 capping agent can be reduced by deuterating one of the two components
853 (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
854 are deuterated, one can observe a significantly lower $G$ and
855 $G^\prime$ values (Table \ref{modelTest}).
856
857 \begin{figure}
858 \includegraphics[width=\linewidth]{aahxntln}
859 \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
860 systems. When butanethiol is deuterated (lower left), its
861 vibrational overlap with hexane decreases significantly. Since
862 aromatic molecules and the butanethiol are vibrationally dissimilar,
863 the change is not as dramatic when toluene is the solvent (right).}
864 \label{aahxntln}
865 \end{figure}
866
867 For the Au / butanethiol / toluene interfaces, having the AA
868 butanethiol deuterated did not yield a significant change in the
869 measured conductance. Compared to the C-H vibrational overlap between
870 hexane and butanethiol, both of which have alkyl chains, the overlap
871 between toluene and butanethiol is not as significant and thus does
872 not contribute as much to the heat exchange process.
873
874 Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
875 that the {\it intra}molecular heat transport due to alkylthiols is
876 highly efficient. Combining our observations with those of Zhang {\it
877 et al.}, it appears that butanethiol acts as a channel to expedite
878 heat flow from the gold surface and into the alkyl chain. The
879 vibrational coupling between the metal and the liquid phase can
880 therefore be enhanced with the presence of suitable capping agents.
881
882 Deuterated models in the UA force field did not decouple the thermal
883 transport as well as in the AA force field. The UA models, even
884 though they have eliminated the high frequency C-H vibrational
885 overlap, still have significant overlap in the lower-frequency
886 portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
887 the UA models did not decouple the low frequency region enough to
888 produce an observable difference for the results of $G$ (Table
889 \ref{modelTest}).
890
891 \begin{figure}
892 \includegraphics[width=\linewidth]{uahxnua}
893 \caption{Vibrational power spectra for UA models for the butanethiol
894 and hexane solvent (upper panel) show the high degree of overlap
895 between these two molecules, particularly at lower frequencies.
896 Deuterating a UA model for the solvent (lower panel) does not
897 decouple the two spectra to the same degree as in the AA force
898 field (see Fig \ref{aahxntln}).}
899 \label{uahxnua}
900 \end{figure}
901
902 \section{Conclusions}
903 The NIVS algorithm has been applied to simulations of
904 butanethiol-capped Au(111) surfaces in the presence of organic
905 solvents. This algorithm allows the application of unphysical thermal
906 flux to transfer heat between the metal and the liquid phase. With the
907 flux applied, we were able to measure the corresponding thermal
908 gradients and to obtain interfacial thermal conductivities. Under
909 steady states, 2-3 ns trajectory simulations are sufficient for
910 computation of this quantity.
911
912 Our simulations have seen significant conductance enhancement in the
913 presence of capping agent, compared with the bare gold / liquid
914 interfaces. The vibrational coupling between the metal and the liquid
915 phase is enhanced by a chemically-bonded capping agent. Furthermore,
916 the coverage percentage of the capping agent plays an important role
917 in the interfacial thermal transport process. Moderately low coverages
918 allow higher contact between capping agent and solvent, and thus could
919 further enhance the heat transfer process, giving a non-monotonic
920 behavior of conductance with increasing coverage.
921
922 Our results, particularly using the UA models, agree well with
923 available experimental data. The AA models tend to overestimate the
924 interfacial thermal conductance in that the classically treated C-H
925 vibrations become too easily populated. Compared to the AA models, the
926 UA models have higher computational efficiency with satisfactory
927 accuracy, and thus are preferable in modeling interfacial thermal
928 transport.
929
930 Of the two definitions for $G$, the discrete form
931 (Eq. \ref{discreteG}) was easier to use and gives out relatively
932 consistent results, while the derivative form (Eq. \ref{derivativeG})
933 is not as versatile. Although $G^\prime$ gives out comparable results
934 and follows similar trend with $G$ when measuring close to fully
935 covered or bare surfaces, the spatial resolution of $T$ profile
936 required for the use of a derivative form is limited by the number of
937 bins and the sampling required to obtain thermal gradient information.
938
939 Vlugt {\it et al.} have investigated the surface thiol structures for
940 nanocrystalline gold and pointed out that they differ from those of
941 the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
942 difference could also cause differences in the interfacial thermal
943 transport behavior. To investigate this problem, one would need an
944 effective method for applying thermal gradients in non-planar
945 (i.e. spherical) geometries.
946
947 \section{Acknowledgments}
948 Support for this project was provided by the National Science
949 Foundation under grant CHE-0848243. Computational time was provided by
950 the Center for Research Computing (CRC) at the University of Notre
951 Dame.
952
953 {\bf Supporting Information Available:} Force field parameters
954 utilized in the simulation are available in the supporting
955 information, as are thermal conductivity data for a range of applied
956 thermal fluxes, simulation temperatures, and solvent densities. This
957 information is available free of charge via the Internet at
958 http://pubs.acs.org.
959
960 \newpage
961
962 \bibliography{interfacial}
963
964 \end{doublespace}
965 \end{document}
966