ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3822
Committed: Tue Dec 18 16:22:13 2012 UTC (12 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 34462 byte(s)
Log Message:
A few changes

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{times}
5 \usepackage{mathptm}
6 \usepackage{setspace}
7 \usepackage{endfloat}
8 \usepackage{caption}
9 \usepackage{graphicx}
10 \usepackage{multirow}
11 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
12 \usepackage[square, comma, sort&compress]{natbib}
13 \usepackage{url}
14 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
15 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
16 9.0in \textwidth 6.5in \brokenpenalty=10000
17
18 % double space list of tables and figures
19 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
20 \setlength{\abovecaptionskip}{20 pt}
21 \setlength{\belowcaptionskip}{30 pt}
22
23 \bibpunct{}{}{,}{s}{}{;}
24
25 \citestyle{nature}
26 \bibliographystyle{achemso}
27
28 \begin{document}
29
30 \title{Simulations of heat conduction at thiolate-capped gold
31 surfaces: The role of chain length and solvent penetration}
32
33 \author{Kelsey M. Stocker and J. Daniel
34 Gezelter\footnote{Corresponding author. \ Electronic mail:
35 gezelter@nd.edu} \\
36 251 Nieuwland Science Hall, \\
37 Department of Chemistry and Biochemistry,\\
38 University of Notre Dame\\
39 Notre Dame, Indiana 46556}
40
41 \date{\today}
42
43 \maketitle
44
45 \begin{doublespace}
46
47 \begin{abstract}
48
49 \end{abstract}
50
51 \newpage
52
53 %\narrowtext
54
55 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56 % **INTRODUCTION**
57 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58 \section{Introduction}
59
60 The structural and dynamical details of interfaces between metal
61 nanoparticles and solvents determines how energy flows between these
62 particles and their surroundings. Understanding this energy flow is
63 essential in designing and functionalizing metallic nanoparticles for
64 plasmonic photothermal
65 therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw} which rely on the ability of metallic
66 nanoparticles to absorb light in the near-IR, a portion of the
67 spectrum in which living tissue is very nearly transparent. The
68 principle of this therapy is to pump the particles at high power at
69 the plasmon resonance and to allow heat dissipation to kill targeted
70 (e.g. cancerous) cells. The relevant physical property controlling
71 this transfer of energy is the interfacial thermal conductance, $G$,
72 which can be somewhat difficult to determine
73 experimentally.\cite{Wilson:2002uq,Plech:2005kx}
74
75 Metallic particles have also been proposed for use in highly efficient
76 thermal-transfer fluids, although careful experiments by Eapen {\it et al.}
77 have shown that metal-particle-based ``nanofluids'' have thermal
78 conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
79 likely cause of previously reported non-Maxwell
80 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
81 is percolation networks of nanoparticles exchanging energy via the
82 solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
83 picture of particle-solvent interactions in order to understand the
84 thermal behavior of complex fluids. To date, there have been few
85 reported values (either from theory or experiment) for $G$ for
86 ligand-protected nanoparticles embedded in liquids, and there is a
87 significant gap in knowledge about how chemically distinct ligands or
88 protecting groups will affect heat transport from the particles.
89
90 The thermal properties of various nanostructured interfaces have been
91 investigated experimentally by a number of groups: Cahill and
92 coworkers studied nanoscale thermal transport from metal
93 nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxides
94 interfaces, and hydrophilic and hydrophobic interfaces between water
95 and solids with different self-assembled
96 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
97 Wang {\it et al.} studied heat transport through long-chain
98 hydrocarbon monolayers on gold substrate at the individual molecular
99 level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
100 cetyltrimethylammonium bromide (CTAB) on the thermal transport between
101 gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
102 et al.} studied the cooling dynamics, which is controlled by thermal
103 interface resistance of glass-embedded metal
104 nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
105 normally considered barriers for heat transport, Alper {\it et al.}
106 suggested that specific ligands (capping agents) could completely
107 eliminate this barrier
108 ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
109
110 In previous simulations, we applied a variant of reverse
111 non-equilibrium molecular dynamics (RNEMD) to calculate the
112 interfacial thermal conductance at a metal / organic solvent interface
113 that had been chemically protected by butanethiolate groups. Our
114 calculations suggest an explanation for the very large thermal
115 conductivity at alkanethiol-capped metal surfaces when compared with
116 bare metal/solvent interfaces. Specifically, the chemical bond
117 between the metal and the ligand introduces a vibrational overlap that
118 is not present without the protecting group, and the overlap between
119 the vibrational spectra (metal to ligand, ligand to solvent) provides
120 a mechanism for rapid thermal transport across the interface.
121
122 One interesting result of our previous work was the observation of
123 {\it non-monotonic dependence} of the thermal conductance on the
124 coverage of a metal surface by a chemical protecting group. Our
125 explanation for this behavior was that gaps in surface coverage
126 allowed solvent to penetrate close to the capping molecules that had
127 been heated by the metal surface, to absorb thermal energy from these
128 molecules, and then diffuse away. The effect of surface coverage is
129 relatively difficult to study as the individual protecting groups have
130 lateral mobility on the surface and can aggregate to form domains on
131 the timescale of the simulation.
132
133 The work reported here involves the use of velocity shearing and
134 scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD) to
135 study surfaces composed of mixed-length chains which collectively form
136 a complete monolayer on the surfaces. These complete (but
137 mixed-chain) surfaces are significantly less prone to surface domain
138 formation, and would allow us to further investigate the mechanism of
139 heat transport to the solvent.
140
141 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
142 % **METHODOLOGY**
143 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
144 \section{Methodology}
145
146 There are many ways to compute bulk transport properties from
147 classical molecular dynamics simulations. Equilibrium Molecular
148 Dynamics (EMD) simulations can be used by computing relevant time
149 correlation functions and assuming that linear response theory
150 holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
151 For some transport properties (notably thermal conductivity), EMD
152 approaches are subject to noise and poor convergence of the relevant
153 correlation functions. Traditional Non-equilibrium Molecular Dynamics
154 (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
155 simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
156 However, the resulting flux is often difficult to
157 measure. Furthermore, problems arise for NEMD simulations of
158 heterogeneous systems, such as phase-phase boundaries or interfaces,
159 where the type of gradient to enforce at the boundary between
160 materials is unclear.
161
162 {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
163 a different approach in that an unphysical {\it flux} is imposed
164 between different regions or ``slabs'' of the simulation
165 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
166 system responds by developing a temperature or momentum {\it gradient}
167 between the two regions. Since the amount of the applied flux is known
168 exactly, and the measurement of a gradient is generally less
169 complicated, imposed-flux methods typically take shorter simulation
170 times to obtain converged results for transport properties. The
171 corresponding temperature or velocity gradients which develop in
172 response to the applied flux are then related (via linear response
173 theory) to the transport coefficient of interest. These methods are
174 quite efficient, in that they do not need many trajectories to provide
175 information about transport properties. To date, they have been
176 utilized in computing thermal and mechanical transfer of both
177 homogeneous
178 liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
179 well as heterogeneous
180 systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
181
182 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
183 % VSS-RNEMD
184 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185 \subsection{VSS-RNEMD}
186 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
187 al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
188 as a sequence of imaginary elastic collisions between particles in
189 regions separated by half of the simulation cell. In each collision,
190 the entire momentum vectors of both particles may be exchanged to
191 generate a thermal flux. Alternatively, a single component of the
192 momentum vectors may be exchanged to generate a shear flux. This
193 algorithm turns out to be quite useful in many simulations of bulk
194 liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
195 system away from ideal Maxwell-Boltzmann distributions, and this has
196 undesirable side-effects when the applied flux becomes
197 large.\cite{Maginn:2010}
198
199 Instead of having momentum exchange between {\it individual particles}
200 in each slab, the NIVS algorithm applies velocity scaling to all of
201 the selected particles in both slabs.\cite{Kuang:2010uq} A combination
202 of linear momentum, kinetic energy, and flux constraint equations
203 governs the amount of velocity scaling performed at each step. NIVS
204 has been shown to be very effective at producing thermal gradients and
205 for computing thermal conductivities, particularly for heterogeneous
206 interfaces. Although the NIVS algorithm can also be applied to impose
207 a directional momentum flux, thermal anisotropy was observed in
208 relatively high flux simulations, and the method is not suitable for
209 imposing a shear flux or for computing shear viscosities.
210
211 The most useful RNEMD
212 approach developed so far utilizes a series of simultaneous velocity
213 shearing and scaling exchanges between the two
214 slabs.\cite{2012MolPh.110..691K} This method provides a set of
215 conservation constraints while simultaneously creating a desired flux
216 between the two slabs. Satisfying the constraint equations ensures
217 that the new configurations are sampled from the same NVE ensemble.
218
219 The VSS moves are applied periodically to scale and shift the particle
220 velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
221 $C$) which are separated by half of the simulation box,
222 \begin{displaymath}
223 \begin{array}{rclcl}
224
225 & \underline{\mathrm{shearing}} & &
226 \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
227 \mathbf{v}_i \leftarrow &
228 \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
229 \rangle\right) + \langle\mathbf{v}_c\rangle \\
230 \mathbf{v}_j \leftarrow &
231 \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
232 \rangle\right) + \langle\mathbf{v}_h\rangle .
233
234 \end{array}
235 \end{displaymath}
236 Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
237 the center of mass velocities in the $C$ and $H$ slabs, respectively.
238 Within the two slabs, particles receive incremental changes or a
239 ``shear'' to their velocities. The amount of shear is governed by the
240 imposed momentum flux, $j_z(\mathbf{p})$
241 \begin{eqnarray}
242 \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
243 \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
244 \end{eqnarray}
245 where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
246 is the interval between two separate operations.
247
248 To simultaneously impose a thermal flux ($J_z$) between the slabs we
249 use energy conservation constraints,
250 \begin{eqnarray}
251 K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
252 \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
253 K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
254 \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
255 \mathbf{a}_h)^2 \label{vss4}.
256 \label{constraint}
257 \end{eqnarray}
258 Simultaneous solution of these quadratic formulae for the scaling
259 coefficients, $c$ and $h$, will ensure that the simulation samples from
260 the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
261 instantaneous translational kinetic energy of each slab. At each time
262 interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
263 and $\mathbf{a}_h$, subject to the imposed momentum flux,
264 $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
265 operations do not change the kinetic energy due to orientational
266 degrees of freedom or the potential energy of a system, configurations
267 after the VSS move have exactly the same energy ({\it and} linear
268 momentum) as before the move.
269
270 As the simulation progresses, the VSS moves are performed on a regular
271 basis, and the system develops a thermal or velocity gradient in
272 response to the applied flux. Using the slope of the temperature or
273 velocity gradient, it is quite simple to obtain of thermal
274 conductivity ($\lambda$),
275 \begin{equation}
276 J_z = -\lambda \frac{\partial T}{\partial z},
277 \end{equation}
278 and shear viscosity ($\eta$),
279 \begin{equation}
280 j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
281 \end{equation}
282 Here, the quantities on the left hand side are the imposed flux
283 values, while the slopes are obtained from linear fits to the
284 gradients that develop in the simulated system.
285
286 The VSS-RNEMD approach is versatile in that it may be used to
287 implement both thermal and shear transport either separately or
288 simultaneously. Perturbations of velocities away from the ideal
289 Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
290 kept to a minimum. This ability to generate simultaneous thermal and
291 shear fluxes has been previously utilized to map out the shear
292 viscosity of SPC/E water over a wide range of temperatures (90~K) with
293 a {\it single 1 ns simulation}.\cite{2012MolPh.110..691K}
294
295 \begin{figure}
296 \includegraphics[width=\linewidth]{figures/rnemd}
297 \caption{The VSS-RNEMD approach imposes unphysical transfer of
298 linear momentum or kinetic energy between a ``hot'' slab and a
299 ``cold'' slab in the simulation box. The system responds to this
300 imposed flux by generating velocity or temperature gradients. The
301 slope of the gradients can then be used to compute transport
302 properties (e.g. shear viscosity or thermal conductivity).}
303 \label{fig:rnemd}
304 \end{figure}
305
306 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
307 % INTERFACIAL CONDUCTANCE
308 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
309 \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
310 to interfacial transport}
311
312 Interfaces between dissimilar materials have transport properties
313 which can be defined as derivatives of the standard transport
314 coefficients in a direction normal to the interface. For example, the
315 {\it interfacial} thermal conductance ($G$) can be thought of as the
316 change in the thermal conductivity ($\lambda$) across the boundary
317 between materials:
318 \begin{align}
319 G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
320 &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
321 \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
322 \label{derivativeG}
323 \end{align}
324 where $z_0$ is the location of the interface between two materials and
325 $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
326 to be the $z$ direction from here on). RNEMD simulations, and
327 particularly the VSS-RNEMD approach, function by applying a momentum
328 or thermal flux and watching the gradient response of the
329 material. This means that the {\it interfacial} conductance is a
330 second derivative property which is subject to significant noise and
331 therefore requires extensive sampling. We have been able to
332 demonstrate the use of the second derivative approach to compute
333 interfacial conductance at chemically-modified metal / solvent
334 interfaces. However, a definition of the interfacial conductance in
335 terms of a temperature difference ($\Delta T$) measured at $z_0$,
336 \begin{displaymath}
337 G = \frac{J_z}{\Delta T_{z_0}},
338 \end{displaymath}
339 is useful once the RNEMD approach has generated a stable temperature
340 gap across the interface.
341
342 \begin{figure}
343 \includegraphics[width=\linewidth]{figures/resistor_series}
344 \caption{The inverse of the interfacial thermal conductance, $G$, is
345 the Kapitza resistance, $R_K$. Because the gold / thiolate/
346 solvent interface extends a significant distance from the metal
347 surface, the interfacial resistance $R_K$ can be computed by
348 summing a series of temperature drops between adjacent temperature
349 bins along the $z$ axis.}
350 \label{fig:resistor_series}
351 \end{figure}
352
353 In the particular case we are studying here, there are two interfaces
354 involved in the transfer of heat from the gold slab to the solvent:
355 the gold/thiolate interface and the thiolate/solvent interface. We
356 could treat the temperature on each side of an interface as discrete,
357 making the interfacial conductance the inverse of the Kaptiza
358 resistance, or $G = \frac{1}{R_k}$. To model the total conductance
359 across multiple interfaces, it is useful to think of the interfaces as
360 a set of resistors wired in series. The total resistance is then
361 additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
362 is the inverse of the total resistance, or $G = \frac{1}{\sum_i
363 R_i}$). In the interfacial region, we treat each bin in the
364 VSS-RNEMD temperature profile as a resistor with resistance
365 $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
366 the set of resistors which spans the gold/thiolate interface, thiolate
367 chains, and thiolate/solvent interface simplifies to
368 \begin{equation}
369 \frac{T_{n}-T_{1}}{J_z},
370 \label{eq:finalG}
371 \end{equation}
372 or the temperature difference between the gold side of the
373 gold/thiolate interface and the solvent side of the thiolate/solvent
374 interface over the applied flux.
375
376
377 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378 % **COMPUTATIONAL DETAILS**
379 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
380 \section{Computational Details}
381
382 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
383 % FORCE-FIELD PARAMETERS
384 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385 \subsection{Force-Field Parameters}
386
387 Our simulations include a number of chemically distinct components.
388 Figure \ref{fig:structures} demonstrates the sites defined for both
389 the {\it n}-hexane and alkanethiolate ligands present in our
390 simulations. Force field parameters are needed for interactions both
391 between the same type of particles and between particles of different
392 species.
393
394 \begin{figure}
395 \includegraphics[width=\linewidth]{figures/structures}
396 \caption{Topologies of the thiolate capping agents and solvent
397 utilized in the simulations. The chemically-distinct sites (S,
398 \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
399 parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
400 \protect\cite{TraPPE-UA.alkanes} and
401 \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
402 atoms were adapted from references
403 \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
404 \protect\cite{hautman:4994}.}
405 \label{fig:structures}
406 \end{figure}
407
408 The Au-Au interactions in metal lattice slab is described by the
409 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
410 potentials include zero-point quantum corrections and are
411 reparametrized for accurate surface energies compared to the
412 Sutton-Chen potentials.\cite{Chen90}
413
414 For the {\it n}-hexane solvent molecules, the TraPPE-UA
415 parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
416 sites are located at the carbon centers for alkyl groups. Bonding
417 interactions, including bond stretches and bends and torsions, were
418 used for intra-molecular sites closer than 3 bonds. For non-bonded
419 interactions, Lennard-Jones potentials are used. We have previously
420 utilized both united atom (UA) and all-atom (AA) force fields for
421 thermal conductivity in previous work,\cite{} and since the united
422 atom force fields cannot populate the high-frequency modes that are
423 present in AA force fields, they appear to work better for modeling
424 thermal conductivity. The TraPPE-UA model for alkanes is known to
425 predict a slightly lower boiling point than experimental values. This
426 is one of the reasons we used a lower average temperature (200K) for
427 our simulations.
428
429 The TraPPE-UA force field includes parameters for thiol
430 molecules\cite{TraPPE-UA.thiols} which were used for the
431 alkanethiolate molecules in our simulations. To derive suitable
432 parameters for butanethiol adsorbed on Au(111) surfaces, we adopted
433 the S parameters from Luedtke and Landman\cite{landman:1998} and
434 modified the parameters for the CTS atom to maintain charge neutrality
435 in the molecule.
436
437 To describe the interactions between metal (Au) and non-metal atoms,
438 we refer to an adsorption study of alkyl thiols on gold surfaces by
439 Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
440 Lennard-Jones form of potential parameters for the interaction between
441 Au and pseudo-atoms CH$_x$ and S based on a well-established and
442 widely-used effective potential of Hautman and Klein for the Au(111)
443 surface.\cite{hautman:4994} As our simulations require the gold slab
444 to be flexible to accommodate thermal excitation, the pair-wise form
445 of potentials they developed was used for our study. Table
446 \ref{table:pars} in the supporting information summarizes the
447 ``metal/non-metal'' parameters utilized in our simulations.
448
449 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450 % SIMULATION PROTOCOL
451 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452 \subsection{Simulation Protocol}
453
454 We have implemented the VSS-RNEMD algorithm in OpenMD, our
455 group molecular dynamics code. A gold slab 11 atoms thick was
456 equilibrated at 1 atm and 200 K. The periodic box was expanded
457 in the z direction to expose two Au(111) faces.
458
459 A full monolayer of thiolates (1/3 the number of surface gold atoms) was placed on three-fold hollow sites on the gold interfaces. To efficiently test the effect of thiolate binding sites on the thermal conductance, all systems had one gold interface with thiolates placed only on fcc hollow sites and the other interface with thiolates only on hcp hollow sites. To test the effect of thiolate chain length on interfacial thermal conductance, full coverages of five chain lengths were tested: butanethiolate, hexanethiolate, octanethiolate, decanethiolate, and dodecanethiolate. To test the effect of mixed chain lengths, full coverages of butanethiolate/decanethiolate and butanethiolate/dodecanethiolate mixtures were created in short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and 87.5/12.5. The short and long chains were placed on the surface hollow sites in a random configuration.
460
461 The simulation box z dimension was set to roughly double the length of the gold/thiolate block. Hexane solvent molecules were placed in the vacant portion of the box using the packmol algorithm. Hexane, a straight chain flexible alkane, is very structurally similar to the thiolate alkane tails; previous work has shown that UA models of hexane and butanethiolate have a high degree of vibrational overlap.\cite{Kuang2011} This overlap should provide a mechanism for thermal energy transfer from the thiolates to the solvent.
462
463 The system was equilibrated to 220 K in the NVT ensemble, allowing the thiolates and solvent to warm gradually. Pressure correction to 1 atm was done in an NPT ensemble that allowed expansion or contraction only in the z direction, so as not to disrupt the crystalline structure of the gold. The diagonal elements of the pressure tensor were monitored during the pressure correction step. If the xx and/or yy elements had a mean above zero throughout the simulation -- indicating residual pressure in the plane of the gold slab -- an additional short NPT equilibration step was performed allowing all box dimensions to change. Once the pressure was stable at 1 atm, a final NVT simulation was performed. All systems were equilibrated in the microcanonical (NVE) ensemble before proceeding with the VSS-RNEMD step.
464
465 A kinetic energy flux was applied using VSS-RNEMD in the NVE ensemble. The total simulation time was 3 nanoseconds, with velocity scaling occurring every 10 femtoseconds. The hot slab was centered in the gold and the cold slab was placed in the center of the solvent region. The average temperature was 220 K, with a temperature difference between the hot and cold slabs of approximately 30 K. The average temperature and kinetic energy flux were carefully selected with two considerations in mind: 1) if the cold bin gets too cold (below ~180 K) the solvent may freeze or undergo a glassy transition, and 2) due to the deep sulfur-gold potential well, sulfur atoms routinely embed into the gold slab, particularly at temperatures above 250 K. Simulation conditions were chosen to avoid both of these undesirable situations. A reversed flux direction resulted in frozen long chain thiolates and solvent too near its boiling point.
466
467 A temperature profile of the system was created by dividing the box into $\sim$ 3 \AA \, bins along the z axis and recording the average temperature of each bin.
468
469
470
471 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
472 % **RESULTS**
473 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
474 \section{Results}
475
476
477 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478 % CHAIN LENGTH
479 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
480 \subsection{Effect of Chain Length}
481
482 We examined full coverages of five chain lengths, n = 4, 6, 8, 10, 12. 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. The trend of interfacial conductance is mostly flat 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. There is, however, a peak in conductance for a chain length of 6 (hexanethiolate). This may be due to the equivalent chain lengths of the hexane solvent and the alkyl chain of the capping agent, leading to an especially high degree of vibrational overlap between the thiolate and solvent. Strong vibrational overlap would allow for efficient thermal energy transfer across the thiolate/solvent interface.
483
484 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
485 % MIXED CHAINS
486 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487 \subsection{Effect of Mixed Chain Lengths}
488
489 Previous work demonstrated that for butanethiolate monolayers on a Au(111) surface, the interfacial conductance was a non-monotonic function of the percent coverage. This is believed to be due to enhanced solvent-thiolate coupling through greater penetration of solvent molecules into the thiolate layer. At lower coverages, hexane solvent can more easily line up lengthwise with the thiolate tails by fitting into gaps between the thiolates. However, a side effect of low coverages is surface aggregation of the thiolates. To simulate the effect of low coverages while preventing aggregation we maintain 100\% thiolate coverage while varying the proportions of short (butanethiolate, n = 4) and long (decanethiolate, n = 10 or dodecanethiolate, n = 12). In systems where there is a mix of short and long chain thiolates, interfacial conductance is a non-monotonic function of the percent of long chains. The depth of the gaps between the long chains is $n_{long} - n_{short}$, which has implications for the ability of the hexane solvent to fill in the gaps between the long chains.
490
491 \subsubsection{Butanethiolate/Decanethiolate}
492 Mixtures of butanethiolate/decanethiolate (n = 4, 10) have a peak interfacial condutance for 25\%/75\% short/long chains.
493
494 \subsubsection{Butanethiolate/Dodecanethiolate}
495 Mixtures of butanethiolate/dodecanethiolate (n = 4, 12) have a peak interfacial conductance for 12.5\%/87.5\% short/long chains.
496
497
498
499
500 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
501 % **DISCUSSION**
502 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
503 \section{Discussion}
504
505 In the mixed chain-length simulations, solvent molecules can become
506 temporarily trapped or entangled with the thiolate chains. Their
507 residence in close proximity to the higher temperature environment
508 close to the surface allows them to carry heat away from the surface
509 quite efficiently. There are two aspects of this behavior that are
510 relevant to thermal conductance of the interface: the residence time
511 of solvent molecules in the thiolate layer, and the orientational
512 ordering of the C-C chains as a mechanism for transferring vibrational
513 energy to these entrapped solvent molecules.
514
515 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
516 % RESIDENCE TIME
517 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
518 \subsection{Mobility for solvent in the interfacial layer}
519
520 We use a simple survival correlation function, $C(t)$, to quantify the
521 residence time of a solvent molecule in the long thiolate chain
522 layer. This function correlates the identity of all hexane molecules
523 within the $z$-coordinate range of the thiolate layer at two separate
524 times. If the solvent molecule is present at both times, the
525 configuration contributes a $1$, while the absence of the molecule at
526 the later time indicates that the solvent molecule has migrated into
527 the bulk, and this configuration contributes a $0$. A steep decay in
528 $C(t)$ indicates a high turnover rate of solvent molecules from the
529 chain region to the bulk. % The correlation function is easily fit
530 % using a biexponential,
531 % \begin{equation}
532 % C(t) = A \, e^{-t/\tau_{short}} + (1-A) e^{-t/\tau_{long}}
533 % \label{eq:biexponential}
534 % \end{equation}
535 % to determine short and long residence timescales and the relative populations of solvent molecules that can escape rapidly.
536 We define the mobility of solvent molecules at the interface as
537 \begin{equation}
538 M = \int_{0}^{T} 1 - C(t) dt,
539 \label{eq:mobility}
540 \end{equation}
541 where T is the length of the simulation.
542 In figure \ref{figure:res} we show that interfacial solvent mobility decreases as the percentage of long thiolate chains increases.
543
544
545 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
546 % ORDER PARAMETER
547 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
548
549 \subsection{Vibrational coupling via orientational ordering}
550
551 As the fraction of long-chain thiolates becomes large, the entrapped
552 solvent molecules must find specific orientations relative to the mean
553 orientation of the thiolate chains. This configuration allows for
554 efficient thermal energy exchange between the thiolate alkyl chain and
555 the solvent molecules.
556
557 To quantify this cooperative
558 ordering, we computed the orientational order parameters and director
559 axes for both the thiolate chains and for the entrapped solvent. The
560 director axis can be easily obtained by diagonalization of the order
561 parameter tensor,
562 \begin{equation}
563 \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
564 \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
565 \end{equation}
566 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
567 the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
568 For both kinds of molecules, the $\mathbf{e}$ vector is defined using
569 the terminal atoms of the chains.
570
571 The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
572 traditionally used to obtain orientational order parameter, while the
573 eigenvector corresponding to the order parameter yields the director
574 axis ($\mathbf{d}(t)$) which defines the average direction of
575 molecular alignment at any time $t$. The overlap between the director
576 axes of the thiolates and the entrapped solvent is time-averaged,
577 \begin{equation}
578 \left \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
579 \mathbf{d}_{solvent} \left( t \right) \right \rangle,
580 \label{eq:orientation}
581 \end{equation}
582 and reported in table \ref{table:ordering}.
583
584 Once the solvent molecules have picked up thermal energy from the
585 thiolates, they carry heat away from the gold as they diffuse back
586 into the bulk solvent. When the percentage of long chains decreases,
587 the tails of the long chains are much more disordered and do not
588 provide structured channels for the solvent to fill.
589
590 Although the alignment of the chains with the entrapped solvent is one
591 possible mechanism for the non-monotonic increase in the conductance
592 as a function
593
594
595 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
596 % **ACKNOWLEDGMENTS**
597 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
598 \section*{Acknowledgments}
599 Support for this project was provided by the
600 National Science Foundation under grant CHE-0848243. Computational
601 time was provided by the Center for Research Computing (CRC) at the
602 University of Notre Dame.
603
604 \newpage
605
606 \bibliography{thiolsRNEMD}
607
608 \end{doublespace}
609 \end{document}
610
611
612
613
614
615
616
617

Properties

Name Value
svn:executable *