ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3861
Committed: Mon Feb 18 18:09:53 2013 UTC (12 years, 5 months ago) by kstocke1
Content type: application/x-tex
File size: 41086 byte(s)
Log Message:

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3
4 \usepackage{caption}
5 \usepackage{float}
6 \usepackage{geometry}
7 \usepackage{natbib}
8 \usepackage{setspace}
9 \usepackage{xkeyval}
10 %%%%%%%%%%%%%%%%%%%%%%%
11 \usepackage{amsmath}
12 \usepackage{amssymb}
13 \usepackage{times}
14 \usepackage{mathptm}
15 \usepackage{setspace}
16 \usepackage{endfloat}
17 \usepackage{caption}
18 \usepackage{tabularx}
19 \usepackage{longtable}
20 \usepackage{graphicx}
21 \usepackage{multirow}
22 \usepackage{multicol}
23 \usepackage{achemso}
24 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25 % \usepackage[square, comma, sort&compress]{natbib}
26 \usepackage{url}
27 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29 9.0in \textwidth 6.5in \brokenpenalty=10000
30
31 % double space list of tables and figures
32 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33 \setlength{\abovecaptionskip}{20 pt}
34 \setlength{\belowcaptionskip}{30 pt}
35
36 % \bibpunct{}{}{,}{s}{}{;}
37
38 %\citestyle{nature}
39 % \bibliographystyle{achemso}
40
41 \title{Simulations of Heat Conduction at Thiolate-Capped Gold
42 Surfaces: The Role of Chain Length and Solvent Penetration}
43
44 \author{Kelsey M. Stocker}
45 \author{J. Daniel Gezelter}
46 \email{gezelter@nd.edu}
47 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\ Department of Chemistry and Biochemistry\\ University of Notre Dame\\ Notre Dame, Indiana 46556}
48
49 \begin{document}
50
51 \newcolumntype{A}{p{1.5in}}
52 \newcolumntype{B}{p{0.75in}}
53
54
55 % \author{Kelsey M. Stocker and J. Daniel
56 % Gezelter\footnote{Corresponding author. \ Electronic mail:
57 % gezelter@nd.edu} \\
58 % 251 Nieuwland Science Hall, \\
59 % Department of Chemistry and Biochemistry,\\
60 % University of Notre Dame\\
61 % Notre Dame, Indiana 46556}
62
63 \date{\today}
64
65 \maketitle
66
67 \begin{doublespace}
68
69 \begin{abstract}
70
71 We report on simulations of heat conduction through Au(111) / hexane
72 interfaces in which the surface has been protected by a mixture of
73 short and long chain alkanethiolate ligands. Reverse
74 non-equilibrium molecular dynamics (RNEMD) was used to create a
75 thermal flux between the metal and solvent, and thermal conductance
76 was computed using the resulting thermal profiles near the
77 interface. We find a non-monotonic dependence of the interfacial
78 thermal conductance on the fraction of long chains present at the
79 interface, and correlate this behavior to both solvent ordering and
80 the rate of solvent escape from the thiolate layer immediately in
81 contact with the metal surface. Our results suggest that a mixed
82 vibrational transfer / convection model is necessary to explain the
83 features of heat transfer at this interface. The alignment of the
84 solvent chains with the ordered ligand allows rapid transfer of
85 energy to the trapped solvent and is the dominant feature for
86 ordered ligand layers. Diffusion of the vibrationally excited
87 solvent into the bulk also plays a significant role when the ligands
88 are less tightly packed.
89
90 \end{abstract}
91
92 \newpage
93
94 %\narrowtext
95
96 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97 % **INTRODUCTION**
98 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
99 \section{Introduction}
100
101 The structural details of the interfaces of metal nanoparticles
102 determine how energy flows between these particles and their
103 surroundings. Understanding this energy flow is essential in designing
104 and functionalizing metallic nanoparticles for use in plasmonic photothermal
105 therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw}
106 which rely on the ability of metallic nanoparticles to absorb light in
107 the near-IR, a portion of the spectrum in which living tissue is very
108 nearly transparent. The relevant physical property controlling the
109 transfer of this energy as heat into the surrounding tissue is the
110 interfacial thermal conductance, $G$, which can be somewhat difficult
111 to determine experimentally.\cite{Wilson:2002uq,Plech:2005kx}
112
113 Metallic particles have also been proposed for use in efficient
114 thermal-transfer fluids, although careful experiments by Eapen {\it et
115 al.} have shown that metal-particle-based nanofluids have thermal
116 conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
117 likely cause of previously reported non-Maxwell
118 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
119 is percolation networks of nanoparticles exchanging energy via the
120 solvent,\cite{Eapen:2007mw} so it is important to get a detailed
121 molecular picture of particle-ligand and ligand-solvent interactions
122 in order to understand the thermal behavior of complex fluids. To
123 date, there have been few reported values (either from theory or
124 experiment) of $G$ for ligand-protected nanoparticles embedded in
125 liquids, and there is a significant gap in knowledge about how
126 chemically distinct ligands or protecting groups will affect heat
127 transport from the particles.
128
129 Experimentally, the thermal properties of various nanostructured
130 interfaces have been investigated by a number of groups. Cahill and
131 coworkers studied thermal transport from metal nanoparticle/fluid
132 interfaces, epitaxial TiN/single crystal oxide interfaces, and
133 hydrophilic and hydrophobic interfaces between water and solids with
134 different self-assembled
135 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101} Schmidt {\it et al.} studied the role of
136 cetyltrimethylammonium bromide (CTAB) on the thermal transport between
137 gold nanorods and solvent.\cite{doi:10.1021/jp8051888}
138 Wang {\it et al.} studied heat transport through long-chain
139 hydrocarbon monolayers on unsolvated gold substrate at the individual molecular
140 level.\cite{Wang10082007} The introduction of solvent adds yet another interface and potential barrier for heat transfer. Juv\'e {\it
141 et al.} studied the cooling dynamics of glass-embedded nanoparticles, which is controlled by thermal
142 interfacial resistance and heat diffusion in the matrix.\cite{PhysRevB.80.195406} Hartland also notes the importance of heat transfer through diffusive solvent behavior.\cite{hartland2011} Although interfaces are
143 normally considered barriers for heat transport, Alper {\it et al.}
144 have suggested that specific ligands (capping agents) could completely
145 eliminate this barrier
146 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
147
148 Recently, Hase and coworkers employed Non-Equilibrium Molecular
149 Dynamics (NEMD) simulations to study thermal transport from hot
150 Au(111) substrate to a self-assembled monolayer of alkylthiol with
151 relatively long chain (8-20 carbon atoms).\cite{hase:2010,hase:2011}
152 These simulations explained many of the features of the experiments of
153 Wang {\it et al.} However, ensemble averaged measurements for heat
154 conductance of interfaces between the capping monolayer on Au and a
155 solvent phase have yet to be studied with their approach. In previous
156 simulations, our group applied a variant of reverse non-equilibrium
157 molecular dynamics (RNEMD) to calculate the interfacial thermal
158 conductance at a metal / organic solvent interface that had been
159 chemically protected by butanethiolate groups.\cite{kuang:AuThl} Our
160 calculations suggested an explanation for the very large thermal
161 conductivity at alkanethiol-capped metal surfaces when compared with
162 bare metal/solvent interfaces. Specifically, the chemical bond
163 between the metal and the ligand introduces a vibrational overlap that
164 is not present without the protecting group, and the overlap between
165 the vibrational spectra (metal to ligand, ligand to solvent) provides
166 a mechanism for rapid thermal transport across the interface.
167
168 A notable result of the previous simulations was the non-monotonic
169 dependence of $G$ on the fractional coverage of the metal surface by
170 the chemical protecting group. Gaps in surface coverage allow the
171 solvent molecules to come into direct contact with ligands that have
172 been heated by the metal surface, absorb thermal energy from the
173 ligands, and then diffuse away. Quantifying the role of overall
174 surface coverage was difficult because the ligands have lateral
175 mobility on the surface and can aggregate to form domains on the
176 timescale of the simulation.
177
178 To isolate the effect of ligand/solvent coupling while avoiding
179 lateral mobility of the surface ligands, the current work utilizes
180 monolayers of mixed chain-lengths in which the length mismatch between
181 long and short chains is sufficient to accommodate solvent
182 molecules. These completely covered (but mixed-chain) surfaces are
183 significantly less prone to surface domain formation, and allow us to
184 further investigate the mechanism of heat transport to the solvent. A
185 thermal flux is created using velocity shearing and scaling reverse
186 non-equilibrium molecular dynamics (VSS-RNEMD), and the resulting
187 temperature profiles are analyzed to yield information about the
188 interfacial thermal conductance.
189
190
191 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192 % **METHODOLOGY**
193 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194 \section{Methodology}
195
196 There are many ways to compute bulk transport properties from
197 classical molecular dynamics simulations. Equilibrium Molecular
198 Dynamics (EMD) simulations can be used to compute the relevant time
199 correlation functions and transport coefficients can be calculated
200 assuming that linear response theory
201 holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
202 For some transport properties (notably thermal conductivity), EMD
203 approaches are subject to noise and poor convergence of the relevant
204 correlation functions. Traditional Non-equilibrium Molecular Dynamics
205 (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
206 simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
207 However, the resulting flux is often difficult to
208 measure. Furthermore, problems arise for NEMD simulations of
209 heterogeneous systems, such as phase-phase boundaries or interfaces,
210 where the type of gradient to enforce at the boundary between
211 materials is unclear.
212
213 {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
214 a different approach in that an unphysical {\it flux} is imposed
215 between different regions or ``slabs'' of the simulation
216 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
217 system responds by developing a temperature or momentum {\it gradient}
218 between the two regions. Since the amount of the applied flux is known
219 exactly, and the measurement of a gradient is generally less
220 complicated, imposed-flux methods typically take shorter simulation
221 times to obtain converged results for transport properties. The
222 corresponding temperature or velocity gradients which develop in
223 response to the applied flux are then related (via linear response
224 theory) to the transport coefficient of interest. These methods are
225 quite efficient, in that they do not need many trajectories to provide
226 information about transport properties. To date, they have been
227 utilized in computing thermal and mechanical transfer of both
228 homogeneous
229 liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
230 well as heterogeneous
231 systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
232
233 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
234 % VSS-RNEMD
235 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
236 \subsection{VSS-RNEMD}
237 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
238 al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
239 as a sequence of imaginary elastic collisions between particles in
240 regions separated by half of the simulation cell. In each collision,
241 the entire momentum vectors of both particles may be exchanged to
242 generate a thermal flux. Alternatively, a single component of the
243 momentum vectors may be exchanged to generate a shear flux. This
244 algorithm turns out to be quite useful in many simulations of bulk
245 liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
246 system away from ideal Maxwell-Boltzmann distributions, which has
247 undesirable side-effects when the applied flux becomes
248 large.\cite{Maginn:2010}
249
250 The most useful alternative RNEMD approach developed so far utilizes a
251 series of simultaneous velocity shearing and scaling (VSS) exchanges between
252 the two slabs.\cite{Kuang2012} This method provides a set of
253 conservation constraints while simultaneously creating a desired flux
254 between the two slabs. Satisfying the constraint equations ensures
255 that the new configurations are sampled from the same NVE ensemble.
256
257 The VSS moves are applied periodically to scale and shift the particle
258 velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
259 $C$) which are separated by half of the simulation box,
260 \begin{displaymath}
261 \begin{array}{rclcl}
262
263 & \underline{\mathrm{shearing}} & &
264 \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
265 \mathbf{v}_i \leftarrow &
266 \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
267 \rangle\right) + \langle\mathbf{v}_c\rangle \\
268 \mathbf{v}_j \leftarrow &
269 \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
270 \rangle\right) + \langle\mathbf{v}_h\rangle .
271
272 \end{array}
273 \end{displaymath}
274 Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
275 the center of mass velocities in the $C$ and $H$ slabs, respectively.
276 Within the two slabs, particles receive incremental changes or a
277 ``shear'' to their velocities. The amount of shear is governed by the
278 imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
279 \begin{eqnarray}
280 \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
281 \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
282 \end{eqnarray}
283 where $M_{\{c,h\}}$ is the total mass of particles within each of the
284 slabs and $\Delta t$ is the interval between two separate operations.
285
286 To simultaneously impose a thermal flux ($J_z$) between the slabs we
287 use energy conservation constraints,
288 \begin{eqnarray}
289 K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
290 \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
291 K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
292 \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
293 \mathbf{a}_h)^2 \label{vss4}.
294 \label{constraint}
295 \end{eqnarray}
296 Simultaneous solution of these quadratic formulae for the scaling
297 coefficients, $c$ and $h$, will ensure that the simulation samples from
298 the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
299 instantaneous translational kinetic energy of each slab. At each time
300 interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
301 and $\mathbf{a}_h$, subject to the imposed momentum flux,
302 $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
303 operations do not change the kinetic energy due to orientational
304 degrees of freedom or the potential energy of a system, configurations
305 after the VSS move have exactly the same energy (and linear
306 momentum) as before the move.
307
308 As the simulation progresses, the VSS moves are performed on a regular
309 basis, and the system develops a thermal or velocity gradient in
310 response to the applied flux. Using the slope of the temperature or
311 velocity gradient, it is quite simple to obtain the thermal
312 conductivity ($\lambda$),
313 \begin{equation}
314 J_z = -\lambda \frac{\partial T}{\partial z},
315 \end{equation}
316 and shear viscosity ($\eta$),
317 \begin{equation}
318 j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
319 \end{equation}
320 Here, the quantities on the left hand side are the imposed flux
321 values, while the slopes are obtained from linear fits to the
322 gradients that develop in the simulated system.
323
324 The VSS-RNEMD approach is versatile in that it may be used to
325 implement both thermal and shear transport either separately or
326 simultaneously. Perturbations of velocities away from the ideal
327 Maxwell-Boltzmann distributions are minimal, as is thermal anisotropy.
328 This ability to generate simultaneous thermal and shear fluxes has
329 been previously utilized to map out the shear viscosity of SPC/E water
330 over a wide range of temperatures (90~K) with a single 1 ns
331 simulation.\cite{Kuang2012}
332
333 \begin{figure}
334 \includegraphics[width=\linewidth]{figures/rnemd}
335 \caption{The VSS-RNEMD approach imposes unphysical transfer of
336 linear momentum or kinetic energy between a ``hot'' slab and a
337 ``cold'' slab in the simulation box. The system responds to this
338 imposed flux by generating velocity or temperature gradients. The
339 slope of the gradients can then be used to compute transport
340 properties (e.g. shear viscosity or thermal conductivity).}
341 \label{fig:rnemd}
342 \end{figure}
343
344 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
345 % INTERFACIAL CONDUCTANCE
346 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
347 \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
348 to interfacial transport}
349
350 Interfaces between dissimilar materials have transport properties
351 which can be defined as derivatives of the standard transport
352 coefficients in a direction normal to the interface. For example, the
353 {\it interfacial} thermal conductance ($G$) can be thought of as the
354 change in the thermal conductivity ($\lambda$) across the boundary
355 between materials:
356 \begin{align}
357 G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
358 &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
359 \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
360 \label{derivativeG}
361 \end{align}
362 where $z_0$ is the location of the interface between two materials and
363 $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
364 to be the $z$ direction from here on). RNEMD simulations, and
365 particularly the VSS-RNEMD approach, function by applying a momentum
366 or thermal flux and watching the gradient response of the
367 material. This means that the {\it interfacial} conductance is a
368 second derivative property which is subject to significant noise and
369 therefore requires extensive sampling. Previous work has demonstrated
370 the use of the second derivative approach to compute interfacial
371 conductance at chemically-modified metal / solvent interfaces.
372 However, a definition of the interfacial conductance in terms of a
373 temperature difference ($\Delta T$) measured at $z_0$,
374 \begin{displaymath}
375 G = \frac{J_z}{\Delta T_{z_0}},
376 \end{displaymath}
377 is useful once the RNEMD approach has generated a stable temperature
378 gap across the interface.
379
380 \begin{figure}
381 \includegraphics[width=\linewidth]{figures/resistor_series}
382 \caption{The inverse of the interfacial thermal conductance, $G$, is
383 the Kapitza resistance, $R_K$. Because the gold / thiolate /
384 solvent interface extends a significant distance from the metal
385 surface, the interfacial resistance $R_K$ can be computed by
386 summing a series of temperature drops between adjacent temperature
387 bins along the $z$ axis. The depicted temperature profile is from a RNEMD simulation of 100\% butanethiolate (C$_4$) coverage.}
388 \label{fig:resistor_series}
389 \end{figure}
390
391 In the particular case we are studying here, there are two interfaces
392 involved in the transfer of heat from the gold slab to the solvent:
393 the metal / thiolate interface and the thiolate / solvent interface. We
394 can treat the temperature on each side of an interface as discrete,
395 making the interfacial conductance the inverse of the Kaptiza
396 resistance, or $G = \frac{1}{R_k}$. To model the total conductance
397 across multiple interfaces, it is useful to think of the interfaces as
398 a set of resistors wired in series. The total resistance is then
399 additive, $R_{total} = \sum_i R_{i}$, and the interfacial conductance
400 is the inverse of the total resistance, or $G = \frac{1}{\sum_i
401 R_i}$). In the interfacial region, we treat each bin in the
402 VSS-RNEMD temperature profile as a resistor with resistance
403 $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
404 the set of resistors which spans the gold / thiolate interface, thiolate
405 chains, and thiolate / solvent interface simplifies to
406 \begin{equation}
407 R_{K} = \frac{T_{n}-T_{1}}{J_z},
408 \label{eq:finalG}
409 \end{equation}
410 or the temperature difference between the gold side of the
411 gold / thiolate interface and the solvent side of the thiolate / solvent
412 interface over the applied flux.
413
414
415 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
416 % **COMPUTATIONAL DETAILS**
417 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
418 \section{Computational Details}
419
420 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
421 % FORCE-FIELD PARAMETERS
422 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
423 \subsection{Force-Field Parameters}
424
425 Our simulations include a number of chemically distinct components.
426 Figure \ref{fig:structures} demonstrates the sites defined for both
427 the {\it n}-hexane and alkanethiolate ligands present in our
428 simulations.
429
430 \begin{figure}
431 \includegraphics[width=\linewidth]{figures/structures}
432 \caption{Topologies of the thiolate capping agents and solvent
433 utilized in the simulations. The chemically-distinct sites (S,
434 \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
435 parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
436 \protect\cite{TraPPE-UA.alkanes} and
437 \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
438 atoms were adapted from references
439 \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
440 \protect\cite{hautman:4994}.}
441 \label{fig:structures}
442 \end{figure}
443
444 The Au-Au interactions in the metal lattice slab were described by the
445 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
446 potentials include zero-point quantum corrections and are
447 reparametrized for accurate surface energies compared to the
448 Sutton-Chen potentials.\cite{Chen90}
449
450 For the {\it n}-hexane solvent molecules, the TraPPE-UA
451 parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
452 sites are located at the carbon centers for alkyl groups. Bonding
453 interactions, including bond stretches and bends and torsions, were
454 used for intra-molecular sites closer than 3 bonds. For non-bonded
455 interactions, Lennard-Jones potentials were used. We have previously
456 utilized both united atom (UA) and all-atom (AA) force fields for
457 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
458 atom force fields cannot populate the high-frequency modes that are
459 present in AA force fields, they appear to work better for modeling
460 thermal conductivity. The TraPPE-UA model for alkanes is known to
461 predict a slightly lower boiling point than experimental values. This
462 is one of the reasons we used a lower average temperature (220 K) for
463 our simulations.
464
465 The TraPPE-UA force field includes parameters for thiol
466 molecules\cite{TraPPE-UA.thiols} which were used for the
467 alkanethiolate molecules in our simulations. To derive suitable
468 parameters for butanethiolate adsorbed on Au(111) surfaces, we adopted
469 the S parameters from Luedtke and Landman\cite{landman:1998} and
470 modified the parameters for the CTS atom to maintain charge neutrality
471 in the molecule.
472
473 To describe the interactions between metal (Au) and non-metal atoms,
474 we refer to an adsorption study of alkyl thiols on gold surfaces by
475 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
476 Lennard-Jones form of potential parameters for the interaction between
477 Au and pseudo-atoms CH$_x$ and S based on a well-established and
478 widely-used effective potential of Hautman and Klein for the Au(111)
479 surface.\cite{hautman:4994} As our simulations require the gold slab
480 to be flexible to accommodate thermal excitation, the pair-wise form
481 of potentials they developed was used for our study.
482
483 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
484 % SIMULATION PROTOCOL
485 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486 \subsection{Simulation Protocol}
487
488 We have implemented the VSS-RNEMD algorithm in OpenMD, our group
489 molecular dynamics code.\cite{openmd} An 1188 atom gold slab was
490 prepared and equilibrated at 1 atm and 200 K. The periodic box was
491 then expanded in the $z$ direction to expose two Au(111) faces on
492 either side of the 11-layer slab.
493
494 A full monolayer of thiolates (1/3 the number of surface gold atoms)
495 was placed on three-fold hollow sites on the gold interfaces. The
496 effect of thiolate binding sites on the thermal conductance was tested
497 by placing thiolates at both fcc and hcp hollow sites. No appreciable
498 difference in the temperature profile due to the location of
499 thiolate binding was noted.
500
501 To test the role of thiolate chain length on interfacial thermal
502 conductance, full coverages of each of five chain lengths were tested:
503 butanethiolate (C$_4$), hexanethiolate (C$_6$), octanethiolate
504 (C$_8$), decanethiolate (C$_{10}$), and dodecanethiolate
505 (C$_{12}$). To test the effect of mixed chain lengths, full coverages
506 of C$_4$ / C$_{10}$ and C$_4$ / C$_{12}$ mixtures were created in
507 short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and
508 87.5/12.5. The short and long chains were placed on the surface hollow
509 sites in a random configuration.
510
511 The simulation box $z$-dimension was set to roughly double the length
512 of the gold / thiolate block. Hexane solvent molecules were placed in
513 the vacant portion of the box using the packmol
514 algorithm.\cite{packmol} Figure \ref{fig:timelapse} shows two of the
515 mixed chain length interfaces both before and after the RNEMD simulation.
516
517 \begin{figure}
518 \includegraphics[width=\linewidth]{figures/timelapse}
519 \caption{Images of 75\%~C$_4$~/~25\%~C$_{12}$ (top panel) and 25\%~C$_4$~/~75\%~C$_{12}$ (bottom panel) interfaces at the beginning and end of 3 ns simulations. Solvent molecules that were initially present in the thiolate layer are colored light blue. Diffusion of the initially-trapped solvent into the bulk is apparent in the interface with fewer long chains. Trapped solvent is orientationally locked to the ordered ligands (and is less able to diffuse into the bulk) when the fraction of long chains increases.}
520 \label{fig:timelapse}
521 \end{figure}
522
523 The system was equilibrated to 220 K in the canonical (NVT) ensemble,
524 allowing the thiolates and solvent to warm gradually. Pressure
525 correction to 1 atm was done using an isobaric-isothermal (NPT)
526 integrator that allowed expansion or contraction only in the $z$
527 direction, maintaining the crystalline structure of the gold as close
528 to the bulk result as possible. The diagonal elements of the pressure
529 tensor were monitored during the pressure equilibration stage. If the
530 $xx$ and/or $yy$ elements had a mean above zero throughout the
531 simulation -- indicating residual surface tension in the plane of the
532 gold slab -- an additional short NPT equilibration step was performed
533 allowing all box dimensions to change. Once the pressure was stable
534 at 1 atm, a final equilibration stage was performed at constant
535 temperature. All systems were equilibrated in the microcanonical (NVE)
536 ensemble before proceeding with the VSS-RNEMD and data collection
537 stages.
538
539 A kinetic energy flux was applied using VSS-RNEMD during a data
540 collection period of 3 nanoseconds, with velocity scaling moves
541 occurring every 10 femtoseconds. The ``hot'' slab was centered in the
542 gold and the ``cold'' slab was placed in the center of the solvent
543 region. The entire system had a (time-averaged) temperature of 220 K,
544 with a temperature difference between the hot and cold slabs of
545 approximately 30 K. The average temperature and kinetic energy flux
546 were selected to avoid solvent freezing (or glass formation) and to
547 prevent the thiolates from burying in the gold slab. The Au-S
548 interaction has a deep potential energy well, which allows sulfur
549 atoms to embed into the gold slab at temperatures above 250 K.
550 Simulation conditions were chosen to avoid both of these
551 situations.
552
553 Temperature profiles of the system were created by dividing the box
554 into $\sim$ 3 \AA \, bins along the $z$ axis and recording the average
555 temperature of each bin.
556
557
558
559 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560 % **RESULTS**
561 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
562 \section{Results}
563
564 The solvent, hexane, is a straight chain flexible alkane that is structurally
565 similar to the thiolate alkane tails. Previous work has shown that UA models
566 of hexane and butanethiolate have a high degree of vibrational
567 overlap.\cite{kuang:AuThl} This overlap provides a mechanism for thermal
568 energy conduction from the thiolates to the solvent. Indeed, we observe that
569 the interfacial conductance is twice as large with the thiolate monolayers (of
570 all chain lengths) than with the bare metal surface.
571
572 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
573 % CHAIN LENGTH
574 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
575 \subsection{Effect of Chain Length}
576
577 We examined full coverages of five alkyl chain lengths, C$_{4}$, C$_{6}$, C$_{8}$, C$_{10}$, and C$_{12}$. As shown in table \ref{table:chainlengthG}, the interfacial conductance is roughly constant as a function of chain length, indicating that the length of the thiolate alkyl chains does not play a significant role in the transport of heat across the gold/thiolate and thiolate/solvent interfaces. In all cases, the hexane solvent was unable to penetrate into the thiolate layer, leading to a persistent 2-4 \AA \, gap between the solvent region and the thiolates. However, while the identity of the alkyl thiolate capping agent has little effect on the interfacial thermal conductance, the presence of a full monolayer of capping agent provides a two-fold increase in the G value relative to a bare gold surface.
578 \begin{longtable}{p{4cm} p{3cm}}
579 \caption{Computed interfacial thermal conductance ($G$) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
580 \\
581 \centering {\bf Chain Length} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
582 \endhead
583 \hline
584 \endfoot
585 \centering bare metal & \centering\arraybackslash 30.2 \\
586 $~~~~~~~~~~~~~~~~~~$ C$_{4}$ & \centering\arraybackslash 59.4 \\
587 $~~~~~~~~~~~~~~~~~~$ C$_{6}$ & \centering\arraybackslash 60.2 \\
588 $~~~~~~~~~~~~~~~~~~$ C$_{8}$ & \centering\arraybackslash 61.0 \\
589 $~~~~~~~~~~~~~~~~~~$ C$_{10}$ & \centering\arraybackslash 58.2 \\
590 $~~~~~~~~~~~~~~~~~~$ C$_{12}$ & \centering\arraybackslash 58.8
591 \label{table:chainlengthG}
592 \end{longtable}
593
594 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595 % MIXED CHAINS
596 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
597 \subsection{Effect of Mixed Chain Lengths}
598
599 Previous simulations have demonstrated non-monotonic behavior for $G$ as a function of the surface coverage. One difficulty with the previous study was the ability of butanethiolate ligands to migrate on the Au(111) surface and to form segregated domains. To simulate the effect of low coverages while preventing thiolate domain formation, we maintain 100\% thiolate coverage while varying the proportions of short (butanethiolate, C$_4$) and long (decanethiolate, C$_{10}$, or dodecanethiolate, C$_{12}$) alkyl chains. Data on the conductance trend as the fraction of long chains was varied is shown in figure \ref{fig:Gstacks}. Note that as in the previous study, $G$ is dependent upon solvent accessibility to thermally excited ligands. Our simulations indicate a similar (but less dramatic) non-monotonic dependence on the fraction of long chains.
600 \begin{figure}
601 \includegraphics[width=\linewidth]{figures/Gstacks2}
602 \caption{Interfacial thermal conductivity of mixed-chains has a non-monotonic dependence on the fraction of long chains (lower panels). At low fractions of long chains, the solvent escape rate ($k_{escape}$) dominates the heat transfer process, while the solvent-thiolate orientational ordering ($<d>$) dominates in systems with higher fractions of long chains (upper panels).}
603 \label{fig:Gstacks}
604 \end{figure}
605
606 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
607 % **DISCUSSION**
608 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
609 \section{Discussion}
610
611 In the mixed chain-length simulations, solvent molecules
612 can become temporarily trapped or entangled with the thiolate chains. Their
613 residence in close proximity to the higher temperature environment close to
614 the surface allows them to carry heat away from the surface quite efficiently.
615 There are two aspects of this behavior that are relevant to thermal
616 conductance of the interface: the residence time of solvent molecules in the
617 thiolate layer, and the alignment of solvent molecules with the ligand alkyl
618 chains as a mechanism for transferring vibrational energy to these entrapped
619 solvent molecules. To quantify these competing effects, we have computed
620 solvent escape rates from the thiolate layer as well as a joint orientational
621 order parameter between the trapped solvent and the thiolate ligands.
622
623 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
624 % RESIDENCE TIME
625 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
626 \subsection{Mobility of solvent in the interfacial layer}
627
628 We use a simple survival correlation function, $C(t)$, to measure the
629 residence time of a solvent molecule in the thiolate
630 layer. This function correlates the identity of all hexane molecules
631 within the $z$-coordinate range of the thiolate layer at two separate
632 times. If the solvent molecule is present at both times, the
633 configuration contributes a $1$, while the absence of the molecule at
634 the later time indicates that the solvent molecule has migrated into
635 the bulk, and this configuration contributes a $0$. A steep decay in
636 $C(t)$ indicates a high turnover rate of solvent molecules from the
637 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
638 \begin{equation}
639 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
640 \label{eq:mobility}
641 \end{equation}
642 where T is the length of the simulation. This is a direct measure of
643 the rate at which solvent molecules initially entangled in the thiolate layer
644 can escape into the bulk. As $k_{escape} \rightarrow 0$, the
645 solvent becomes permanently trapped in the thiolate layer. In
646 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
647 decreases as the fraction of long thiolate chains increases.
648
649 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
650 % ORDER PARAMETER
651 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
652
653 \subsection{Vibrational coupling via orientational ordering}
654
655 As the fraction of long-chain thiolates increases, the entrapped
656 solvent molecules must find specific orientations relative to the mean
657 orientation of the thiolate chains. This alignment allows for
658 efficient thermal energy exchange between the thiolate alkyl chain and
659 the solvent molecules.
660
661 Once the interfacial solvent molecules have picked up thermal energy from the
662 thiolates, they carry heat away from the gold as they diffuse back
663 into the bulk solvent. When the percentage of long chains decreases,
664 the tails of the long chains are much more disordered and do not
665 provide structured channels for the solvent to fill.
666
667 To measure this cooperative ordering, we compute the orientational order
668 parameters and director axes for both the thiolate chains and for the
669 entrapped solvent. The director axis can be easily obtained by diagonalization
670 of the order parameter tensor,
671 \begin{equation}
672 \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
673 \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
674 \end{equation}
675 where $\mathbf{e}_{i \alpha}$ is the $\alpha = x,y,z$ component of
676 the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
677 For both the solvent and the ligand, the $\mathbf{e}$ vector is defined using
678 the terminal atoms of the molecule.
679
680 The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
681 traditionally used to obtain the orientational order parameter, while the
682 eigenvector corresponding to the order parameter yields the director
683 axis ($\mathbf{d}(t)$), which defines the average direction of
684 molecular alignment at any time $t$. The overlap between the director
685 axes of the thiolates and the entrapped solvent is time-averaged,
686 \begin{equation}
687 \langle d \rangle = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
688 \mathbf{d}_{solvent} \left( t \right) \rangle_t
689 \label{eq:orientation}
690 \end{equation}
691 and reported in figure \ref{fig:Gstacks}. Values of $\langle d \rangle$ range from $0$ (solvent molecules in the ligand layer are perpendicular to the thiolate chains) to $1$ (solvent and ligand chains are aligned parallel to each other).
692
693 C$_4$ / C$_{10}$ mixed monolayers have a peak interfacial conductance with 75\% long chains. At this fraction of long chains, the cooperative orientational ordering of the solvent molecules and chains becomes the dominant effect while the solvent escape rate is quite slow. C$_4$ / C$_{12}$ mixtures have a peak interfacial conductance for 87.5\% long chains. The solvent-thiolate orientational ordering reaches its maximum value at this long chain fraction. Long chain fractions of over $0.5$ for the C$_4$ / C$_{12}$ system are well ordered, but this effect is tempered by the exceptionally slow solvent escape rate ($\sim$ 1 molecule / 2 ns).
694
695 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
696 % **CONCLUSIONS**
697 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
698 \section{Conclusions}
699 Our results suggest that a mixed vibrational transfer / convection
700 model may be necessary to explain the features of heat transfer at
701 this interface. The alignment of the solvent chains with the ordered
702 ligand allows rapid transfer of energy to the trapped solvent and
703 becomes the dominant feature for ordered ligand layers. Diffusion of
704 the vibrationally excited solvent into the bulk also plays a
705 significant role when the ligands are less tightly packed.
706
707 In the language of earlier continuum approaches to interfacial
708 conductance,\cite{RevModPhys.61.605} the alignment of the chains is an
709 important factor in the transfer of phonons from the thiolate layer to the
710 trapped solvent. The aligned solvent and thiolate chains have nearly identical
711 acoustic impedances and the phonons can scatter directly into a solvent
712 molecule that has been forced into alignment. When the entrapped solvent has
713 more configurations available, the likelihood of an impedance mismatch is
714 higher, and the phonon scatters into the solvent with lower
715 efficiency. The fractional coverage of the long chains is therefore a simple
716 way of tuning the acoustic mismatch between the thiolate layer and the hexane
717 solvent.
718
719 Efficient heat transfer also can be accomplished via convective or diffusive motion of vibrationally excited solvent back into the bulk. Once the entrapped solvent becomes too tightly aligned with the ligands, however, the convective avenue of heat transfer is cut off.
720
721 Our simulations suggest a number of routes to make interfaces with high thermal conductance. If it is possible to create an interface which forces the solvent into alignment with a ligand that shares many of the solvent's vibrational modes, while simultaneously preserving the ability of the solvent to diffuse back into the bulk, we would expect a significant jump in the interfacial conductance. One possible way to accomplish this is to use polyene ligands with alternating unsaturated bonds. These are significantly more rigid than long alkanes, and could force solvent alignment (even at low relative coverages) while preserving mobility of solvent molecules within the ligand layer.
722
723 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
724 % **ACKNOWLEDGMENTS**
725 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
726 \section*{Acknowledgments}
727
728 We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
729 this project was provided by the National Science Foundation under grant
730 CHE-0848243. Computational time was provided by the Center for Research
731 Computing (CRC) at the University of Notre Dame.
732
733 \newpage
734
735 \bibliography{thiolsRNEMD}
736
737 \end{doublespace}
738 \end{document}
739
740
741
742
743
744
745
746

Properties

Name Value
svn:executable *