ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3850
Committed: Fri Dec 21 19:29:36 2012 UTC (12 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 35671 byte(s)
Log Message:
edits

File Contents

# User Rev Content
1 kstocke1 3801 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4 gezelter 3819 \usepackage{times}
5     \usepackage{mathptm}
6 kstocke1 3801 \usepackage{setspace}
7     \usepackage{endfloat}
8     \usepackage{caption}
9 kstocke1 3830 \usepackage{tabularx}
10     \usepackage{longtable}
11 kstocke1 3801 \usepackage{graphicx}
12     \usepackage{multirow}
13 kstocke1 3830 \usepackage{multicol}
14 gezelter 3822 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
15 kstocke1 3801 \usepackage[square, comma, sort&compress]{natbib}
16     \usepackage{url}
17     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
18     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
19     9.0in \textwidth 6.5in \brokenpenalty=10000
20    
21     % double space list of tables and figures
22     %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23     \setlength{\abovecaptionskip}{20 pt}
24     \setlength{\belowcaptionskip}{30 pt}
25    
26     \bibpunct{}{}{,}{s}{}{;}
27 gezelter 3822
28     \citestyle{nature}
29 kstocke1 3801 \bibliographystyle{achemso}
30    
31     \begin{document}
32    
33 kstocke1 3830 \newcolumntype{A}{p{1.5in}}
34     \newcolumntype{B}{p{0.75in}}
35    
36 gezelter 3819 \title{Simulations of heat conduction at thiolate-capped gold
37     surfaces: The role of chain length and solvent penetration}
38 kstocke1 3801
39 gezelter 3803 \author{Kelsey M. Stocker and J. Daniel
40     Gezelter\footnote{Corresponding author. \ Electronic mail:
41     gezelter@nd.edu} \\
42     251 Nieuwland Science Hall, \\
43 kstocke1 3801 Department of Chemistry and Biochemistry,\\
44     University of Notre Dame\\
45     Notre Dame, Indiana 46556}
46    
47     \date{\today}
48    
49     \maketitle
50    
51     \begin{doublespace}
52    
53     \begin{abstract}
54 kstocke1 3815
55 kstocke1 3801 \end{abstract}
56    
57     \newpage
58    
59     %\narrowtext
60    
61     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62     % **INTRODUCTION**
63     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64     \section{Introduction}
65    
66 gezelter 3848 The structural details of the interfaces of metal nanoparticles
67     determines how energy flows between these particles and their
68     surroundings. Understanding this energy flow is essential in designing
69     and functionalizing metallic nanoparticles for plasmonic photothermal
70     therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw}
71     which rely on the ability of metallic nanoparticles to absorb light in
72     the near-IR, a portion of the spectrum in which living tissue is very
73     nearly transparent. The relevant physical property controlling the
74     transfer of this energy as heat into the surrounding tissue is the
75     interfacial thermal conductance, $G$, which can be somewhat difficult
76     to determine experimentally.\cite{Wilson:2002uq,Plech:2005kx}
77 gezelter 3803
78     Metallic particles have also been proposed for use in highly efficient
79 gezelter 3848 thermal-transfer fluids, although careful experiments by Eapen {\it et
80 gezelter 3850 al.} have shown that metal-particle-based nanofluids have thermal
81     conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
82     likely cause of previously reported non-Maxwell
83 gezelter 3803 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
84     is percolation networks of nanoparticles exchanging energy via the
85 gezelter 3850 solvent,\cite{Eapen:2007mw} so it is important to get a detailed
86     molecular picture of particle-ligand and ligand-solvent interactions
87     in order to understand the thermal behavior of complex fluids. To
88     date, there have been few reported values (either from theory or
89     experiment) for $G$ for ligand-protected nanoparticles embedded in
90     liquids, and there is a significant gap in knowledge about how
91     chemically distinct ligands or protecting groups will affect heat
92     transport from the particles.
93 gezelter 3803
94 gezelter 3848 Experimentally, the thermal properties of various nanostructured
95     interfaces have been investigated by a number of groups. Cahill and
96 gezelter 3850 coworkers studied thermal transport from metal nanoparticle/fluid
97     interfaces, epitaxial TiN/single crystal oxide interfaces, and
98     hydrophilic and hydrophobic interfaces between water and solids with
99     different self-assembled
100 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
101 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
102     hydrocarbon monolayers on gold substrate at the individual molecular
103     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
104     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
105     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
106     et al.} studied the cooling dynamics, which is controlled by thermal
107     interface resistance of glass-embedded metal
108     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
109     normally considered barriers for heat transport, Alper {\it et al.}
110 gezelter 3850 have suggested that specific ligands (capping agents) could completely
111 gezelter 3803 eliminate this barrier
112     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
113    
114     In previous simulations, we applied a variant of reverse
115     non-equilibrium molecular dynamics (RNEMD) to calculate the
116     interfacial thermal conductance at a metal / organic solvent interface
117     that had been chemically protected by butanethiolate groups. Our
118 gezelter 3850 calculations suggested an explanation for the very large thermal
119 gezelter 3803 conductivity at alkanethiol-capped metal surfaces when compared with
120     bare metal/solvent interfaces. Specifically, the chemical bond
121     between the metal and the ligand introduces a vibrational overlap that
122     is not present without the protecting group, and the overlap between
123     the vibrational spectra (metal to ligand, ligand to solvent) provides
124     a mechanism for rapid thermal transport across the interface.
125    
126 gezelter 3850 A notable result of the previous simulations was the non-monotonic
127     dependence of $G$ on the fractional coverage of the metal surface by
128     the chemical protecting group. Gaps in surface coverage allow the
129     solvent molecules come into direct contact with ligands that had been
130     heated by the metal surface, absorb thermal energy from the ligands,
131     and then diffuse away. Quantifying the role of surface coverage is
132     difficult as the ligands have lateral mobility on the surface and can
133     aggregate to form domains on the timescale of the simulation.
134 gezelter 3803
135 gezelter 3850 To isolate this effect without worrying about lateral mobility of the
136     surface ligands, the current work involves mixed-chain-length
137     monolayers in which the length mismatch between long and short chains
138     is sufficient to accomodate solvent molecules. These completely
139     covered (but mixed-chain) surfaces are significantly less prone to
140     surface domain formation, and allow us to further investigate the
141     mechanism of heat transport to the solvent. A thermal flux is created
142     using velocity shearing and scaling reverse non-equilibrium molecular
143     dynamics (VSS-RNEMD), and the resulting temperature profiles are
144     analyzed to yield information about the interfacial thermal
145     conductance.
146 gezelter 3803
147 gezelter 3848
148 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
149     % **METHODOLOGY**
150     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
151     \section{Methodology}
152    
153 gezelter 3803 There are many ways to compute bulk transport properties from
154     classical molecular dynamics simulations. Equilibrium Molecular
155     Dynamics (EMD) simulations can be used by computing relevant time
156     correlation functions and assuming that linear response theory
157     holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
158     For some transport properties (notably thermal conductivity), EMD
159     approaches are subject to noise and poor convergence of the relevant
160     correlation functions. Traditional Non-equilibrium Molecular Dynamics
161     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
162     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
163     However, the resulting flux is often difficult to
164     measure. Furthermore, problems arise for NEMD simulations of
165     heterogeneous systems, such as phase-phase boundaries or interfaces,
166     where the type of gradient to enforce at the boundary between
167     materials is unclear.
168    
169     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
170     a different approach in that an unphysical {\it flux} is imposed
171     between different regions or ``slabs'' of the simulation
172 gezelter 3850 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
173 gezelter 3803 system responds by developing a temperature or momentum {\it gradient}
174     between the two regions. Since the amount of the applied flux is known
175     exactly, and the measurement of a gradient is generally less
176     complicated, imposed-flux methods typically take shorter simulation
177     times to obtain converged results for transport properties. The
178     corresponding temperature or velocity gradients which develop in
179     response to the applied flux are then related (via linear response
180     theory) to the transport coefficient of interest. These methods are
181     quite efficient, in that they do not need many trajectories to provide
182     information about transport properties. To date, they have been
183     utilized in computing thermal and mechanical transfer of both
184     homogeneous
185     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
186     well as heterogeneous
187     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
188    
189 gezelter 3822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
190     % VSS-RNEMD
191     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192     \subsection{VSS-RNEMD}
193 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
194     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
195     as a sequence of imaginary elastic collisions between particles in
196     regions separated by half of the simulation cell. In each collision,
197     the entire momentum vectors of both particles may be exchanged to
198     generate a thermal flux. Alternatively, a single component of the
199     momentum vectors may be exchanged to generate a shear flux. This
200     algorithm turns out to be quite useful in many simulations of bulk
201     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
202     system away from ideal Maxwell-Boltzmann distributions, and this has
203     undesirable side-effects when the applied flux becomes
204     large.\cite{Maginn:2010}
205    
206 gezelter 3848 The most useful alternative RNEMD approach developed so far utilizes a
207     series of simultaneous velocity shearing and scaling exchanges between
208 gezelter 3850 the two slabs.\cite{Kuang2012} This method provides a set of
209 gezelter 3803 conservation constraints while simultaneously creating a desired flux
210     between the two slabs. Satisfying the constraint equations ensures
211     that the new configurations are sampled from the same NVE ensemble.
212    
213     The VSS moves are applied periodically to scale and shift the particle
214     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
215     $C$) which are separated by half of the simulation box,
216     \begin{displaymath}
217     \begin{array}{rclcl}
218    
219     & \underline{\mathrm{shearing}} & &
220     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
221     \mathbf{v}_i \leftarrow &
222     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
223     \rangle\right) + \langle\mathbf{v}_c\rangle \\
224     \mathbf{v}_j \leftarrow &
225     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
226     \rangle\right) + \langle\mathbf{v}_h\rangle .
227    
228     \end{array}
229     \end{displaymath}
230     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
231     the center of mass velocities in the $C$ and $H$ slabs, respectively.
232     Within the two slabs, particles receive incremental changes or a
233     ``shear'' to their velocities. The amount of shear is governed by the
234     imposed momentum flux, $j_z(\mathbf{p})$
235     \begin{eqnarray}
236     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
237     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
238     \end{eqnarray}
239     where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
240     is the interval between two separate operations.
241    
242     To simultaneously impose a thermal flux ($J_z$) between the slabs we
243     use energy conservation constraints,
244     \begin{eqnarray}
245     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
246     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
247     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
248     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
249     \mathbf{a}_h)^2 \label{vss4}.
250     \label{constraint}
251     \end{eqnarray}
252     Simultaneous solution of these quadratic formulae for the scaling
253     coefficients, $c$ and $h$, will ensure that the simulation samples from
254     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
255     instantaneous translational kinetic energy of each slab. At each time
256     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
257     and $\mathbf{a}_h$, subject to the imposed momentum flux,
258     $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
259     operations do not change the kinetic energy due to orientational
260     degrees of freedom or the potential energy of a system, configurations
261     after the VSS move have exactly the same energy ({\it and} linear
262     momentum) as before the move.
263    
264     As the simulation progresses, the VSS moves are performed on a regular
265     basis, and the system develops a thermal or velocity gradient in
266     response to the applied flux. Using the slope of the temperature or
267     velocity gradient, it is quite simple to obtain of thermal
268     conductivity ($\lambda$),
269     \begin{equation}
270     J_z = -\lambda \frac{\partial T}{\partial z},
271     \end{equation}
272     and shear viscosity ($\eta$),
273     \begin{equation}
274     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
275     \end{equation}
276     Here, the quantities on the left hand side are the imposed flux
277     values, while the slopes are obtained from linear fits to the
278     gradients that develop in the simulated system.
279    
280     The VSS-RNEMD approach is versatile in that it may be used to
281     implement both thermal and shear transport either separately or
282     simultaneously. Perturbations of velocities away from the ideal
283     Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
284     kept to a minimum. This ability to generate simultaneous thermal and
285     shear fluxes has been previously utilized to map out the shear
286     viscosity of SPC/E water over a wide range of temperatures (90~K) with
287 gezelter 3850 a single 1 ns simulation.\cite{Kuang2012}
288 gezelter 3803
289 gezelter 3822 \begin{figure}
290     \includegraphics[width=\linewidth]{figures/rnemd}
291     \caption{The VSS-RNEMD approach imposes unphysical transfer of
292     linear momentum or kinetic energy between a ``hot'' slab and a
293     ``cold'' slab in the simulation box. The system responds to this
294     imposed flux by generating velocity or temperature gradients. The
295     slope of the gradients can then be used to compute transport
296     properties (e.g. shear viscosity or thermal conductivity).}
297     \label{fig:rnemd}
298     \end{figure}
299    
300     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
301     % INTERFACIAL CONDUCTANCE
302     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303     \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
304 gezelter 3803 to interfacial transport}
305 kstocke1 3801
306 gezelter 3803 Interfaces between dissimilar materials have transport properties
307     which can be defined as derivatives of the standard transport
308     coefficients in a direction normal to the interface. For example, the
309     {\it interfacial} thermal conductance ($G$) can be thought of as the
310     change in the thermal conductivity ($\lambda$) across the boundary
311     between materials:
312     \begin{align}
313     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
314     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
315     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
316     \label{derivativeG}
317     \end{align}
318     where $z_0$ is the location of the interface between two materials and
319     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
320     to be the $z$ direction from here on). RNEMD simulations, and
321     particularly the VSS-RNEMD approach, function by applying a momentum
322     or thermal flux and watching the gradient response of the
323     material. This means that the {\it interfacial} conductance is a
324     second derivative property which is subject to significant noise and
325     therefore requires extensive sampling. We have been able to
326     demonstrate the use of the second derivative approach to compute
327     interfacial conductance at chemically-modified metal / solvent
328     interfaces. However, a definition of the interfacial conductance in
329     terms of a temperature difference ($\Delta T$) measured at $z_0$,
330     \begin{displaymath}
331     G = \frac{J_z}{\Delta T_{z_0}},
332     \end{displaymath}
333     is useful once the RNEMD approach has generated a stable temperature
334     gap across the interface.
335    
336     \begin{figure}
337     \includegraphics[width=\linewidth]{figures/resistor_series}
338 gezelter 3822 \caption{The inverse of the interfacial thermal conductance, $G$, is
339     the Kapitza resistance, $R_K$. Because the gold / thiolate/
340     solvent interface extends a significant distance from the metal
341     surface, the interfacial resistance $R_K$ can be computed by
342     summing a series of temperature drops between adjacent temperature
343     bins along the $z$ axis.}
344 gezelter 3803 \label{fig:resistor_series}
345     \end{figure}
346    
347     In the particular case we are studying here, there are two interfaces
348     involved in the transfer of heat from the gold slab to the solvent:
349 gezelter 3850 the metal/thiolate interface and the thiolate/solvent interface. We
350 gezelter 3803 could treat the temperature on each side of an interface as discrete,
351     making the interfacial conductance the inverse of the Kaptiza
352     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
353     across multiple interfaces, it is useful to think of the interfaces as
354     a set of resistors wired in series. The total resistance is then
355     additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
356     is the inverse of the total resistance, or $G = \frac{1}{\sum_i
357     R_i}$). In the interfacial region, we treat each bin in the
358     VSS-RNEMD temperature profile as a resistor with resistance
359     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
360     the set of resistors which spans the gold/thiolate interface, thiolate
361     chains, and thiolate/solvent interface simplifies to
362 kstocke1 3801 \begin{equation}
363 gezelter 3825 R_{K} = \frac{T_{n}-T_{1}}{J_z},
364 gezelter 3803 \label{eq:finalG}
365     \end{equation}
366     or the temperature difference between the gold side of the
367     gold/thiolate interface and the solvent side of the thiolate/solvent
368     interface over the applied flux.
369 kstocke1 3801
370    
371     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372     % **COMPUTATIONAL DETAILS**
373     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374     \section{Computational Details}
375    
376 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
377 gezelter 3819 % FORCE-FIELD PARAMETERS
378     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
379     \subsection{Force-Field Parameters}
380 kstocke1 3801
381 gezelter 3819 Our simulations include a number of chemically distinct components.
382     Figure \ref{fig:structures} demonstrates the sites defined for both
383     the {\it n}-hexane and alkanethiolate ligands present in our
384 gezelter 3850 simulations.
385 gezelter 3819
386     \begin{figure}
387     \includegraphics[width=\linewidth]{figures/structures}
388 gezelter 3822 \caption{Topologies of the thiolate capping agents and solvent
389     utilized in the simulations. The chemically-distinct sites (S,
390     \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
391     parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
392     \protect\cite{TraPPE-UA.alkanes} and
393     \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
394     atoms were adapted from references
395     \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
396     \protect\cite{hautman:4994}.}
397 gezelter 3819 \label{fig:structures}
398     \end{figure}
399    
400 kstocke1 3829 The Au-Au interactions in the metal lattice slab are described by the
401 gezelter 3819 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
402     potentials include zero-point quantum corrections and are
403     reparametrized for accurate surface energies compared to the
404     Sutton-Chen potentials.\cite{Chen90}
405    
406     For the {\it n}-hexane solvent molecules, the TraPPE-UA
407     parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
408     sites are located at the carbon centers for alkyl groups. Bonding
409     interactions, including bond stretches and bends and torsions, were
410     used for intra-molecular sites closer than 3 bonds. For non-bonded
411 gezelter 3850 interactions, Lennard-Jones potentials were used. We have previously
412 gezelter 3819 utilized both united atom (UA) and all-atom (AA) force fields for
413 gezelter 3850 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
414 gezelter 3819 atom force fields cannot populate the high-frequency modes that are
415     present in AA force fields, they appear to work better for modeling
416     thermal conductivity. The TraPPE-UA model for alkanes is known to
417     predict a slightly lower boiling point than experimental values. This
418     is one of the reasons we used a lower average temperature (200K) for
419 gezelter 3850 our simulations.
420 gezelter 3819
421     The TraPPE-UA force field includes parameters for thiol
422     molecules\cite{TraPPE-UA.thiols} which were used for the
423     alkanethiolate molecules in our simulations. To derive suitable
424     parameters for butanethiol adsorbed on Au(111) surfaces, we adopted
425     the S parameters from Luedtke and Landman\cite{landman:1998} and
426     modified the parameters for the CTS atom to maintain charge neutrality
427     in the molecule.
428    
429     To describe the interactions between metal (Au) and non-metal atoms,
430     we refer to an adsorption study of alkyl thiols on gold surfaces by
431     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
432     Lennard-Jones form of potential parameters for the interaction between
433     Au and pseudo-atoms CH$_x$ and S based on a well-established and
434     widely-used effective potential of Hautman and Klein for the Au(111)
435     surface.\cite{hautman:4994} As our simulations require the gold slab
436     to be flexible to accommodate thermal excitation, the pair-wise form
437 kstocke1 3832 of potentials they developed was used for our study.
438 kstocke1 3821
439     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
440     % SIMULATION PROTOCOL
441     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
442     \subsection{Simulation Protocol}
443    
444 gezelter 3850 We have implemented the VSS-RNEMD algorithm in OpenMD, our group
445     molecular dynamics code.\cite{openmd} An 1188 atom gold slab was
446     equilibrated at 1 atm and 200 K. The periodic box was then expanded
447     in the $z$-direction to expose two Au(111) faces on either side of the
448     11-atom thick slab.
449 kstocke1 3821
450 gezelter 3850 A full monolayer of thiolates (1/3 the number of surface gold atoms)
451     was placed on three-fold hollow sites on the gold interfaces. The
452     effect of thiolate binding sites on the thermal conductance was tested
453     by placing thiolates at both fcc and hcp hollow sites. No appreciable
454     difference in the temperature profile was noted due to the location of
455     thiolate binding.
456 kstocke1 3821
457 gezelter 3850 To test the role of thiolate chain length on interfacial thermal
458     conductance, full coverages of each of five chain lengths were tested:
459     butanethiolate (C$_4$), hexanethiolate (C$_6$), octanethiolate
460     (C$_8$), decanethiolate (C$_{10}$), and dodecanethiolate
461     (C$_{12}$). To test the effect of mixed chain lengths, full coverages
462     of C$_4$ / C$_{10}$ and C$_4$ / C$_{12}$ mixtures were created in
463     short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and
464     87.5/12.5. The short and long chains were placed on the surface hollow
465     sites in a random configuration.
466 kstocke1 3821
467 gezelter 3850 The simulation box $z$-dimension was set to roughly double the length
468     of the gold/thiolate block. Hexane solvent molecules were placed in
469     the vacant portion of the box using the packmol algorithm.
470 kstocke1 3821
471 gezelter 3850 The system was equilibrated to 220 K in the canonical (NVT) ensemble,
472     allowing the thiolates and solvent to warm gradually. Pressure
473     correction to 1 atm was done using an isobaric-isothermal (NPT)
474     integrator that allowed expansion or contraction only in the $z$
475     direction, maintaining the crystalline structure of the gold as close
476     to the bulk result as possible. The diagonal elements of the pressure
477     tensor were monitored during the pressure equilibration stage. If the
478     $xx$ and/or $yy$ elements had a mean above zero throughout the
479     simulation -- indicating residual surface tension in the plane of the
480     gold slab -- an additional short NPT equilibration step was performed
481     allowing all box dimensions to change. Once the pressure was stable
482     at 1 atm, a final equilibration stage was performed at constant
483     temperature. All systems were equilibrated in the microcanonical (NVE)
484     ensemble before proceeding with the VSS-RNEMD and data collection
485     stages.
486 kstocke1 3821
487 gezelter 3850 A kinetic energy flux was applied using VSS-RNEMD. The total
488     simulation time was 3 nanoseconds, with velocity scaling occurring
489     every 10 femtoseconds. The hot slab was centered in the gold and the
490     cold slab was placed in the center of the solvent region. The entire
491     system has a (time-averaged) temperature of 220 K, with a temperature
492     difference between the hot and cold slabs of approximately 30 K. The
493     average temperature and kinetic energy flux were selected to prevent
494     solvent freezing (or glass formation) and to not allow the thiolates
495     to bury in the gold slab. The Au-S interaction has a deep potential
496     energy well, which allows sulfur atoms to embed into the gold slab at
497     temperatures above 250 K. Simulation conditions were chosen to avoid
498     both of these undesirable situations.
499 kstocke1 3821
500 gezelter 3850 Temperature profiles of the system were created by dividing the box
501     into $\sim$ 3 \AA \, bins along the $z$ axis and recording the average
502     temperature of each bin.
503 kstocke1 3821
504 gezelter 3850
505 kstocke1 3801
506     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
507     % **RESULTS**
508     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
509     \section{Results}
510    
511 gezelter 3850 The solvent, hexane, a straight chain flexible alkane, is structurally
512     similar to the thiolate alkane tails, and previous work has shown that
513     UA models of hexane and butanethiolate have a high degree of
514     vibrational overlap.\cite{kuang:AuThl} This overlap provides a
515     mechanism for thermal energy transfer from the thiolates to the
516     solvent.
517 kstocke1 3801
518 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
519     % CHAIN LENGTH
520     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521     \subsection{Effect of Chain Length}
522    
523 kstocke1 3842 We examined full coverages of five chain lengths, n = 4, 6, 8, 10, 12. As shown in table \ref{table:chainlengthG}, the interfacial conductance is 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.
524 kstocke1 3830 \begin{longtable}{p{4cm} p{3cm}}
525     \caption{Computed interfacial thermal conductance (G) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
526     \\
527 kstocke1 3831 \centering {\bf Chain Length (n)} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
528 kstocke1 3830 \endhead
529     \hline
530     \endfoot
531 kstocke1 3832 \centering bare gold & \centering\arraybackslash 30.2 \\
532     \centering 4 & \centering\arraybackslash 59.4 \\
533     \centering 6 & \centering\arraybackslash 60.2 \\
534     \centering 8 & \centering\arraybackslash 61.0 \\
535     \centering 10 & \centering\arraybackslash 58.2 \\
536     \centering 12 & \centering\arraybackslash 58.8
537 kstocke1 3830 \label{table:chainlengthG}
538     \end{longtable}
539 kstocke1 3801
540 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541     % MIXED CHAINS
542     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
543     \subsection{Effect of Mixed Chain Lengths}
544    
545 kstocke1 3844 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 penetrate into the interfacial layer 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 thiolate domain formation, we maintain 100\% thiolate coverage while varying the proportions of short (butanethiolate, n = 4) and long (decanethiolate, n = 10 or dodecanethiolate, n = 12) thiolate alkyl chains. We show in figure \ref{fig:Gstacks} that the interfacial conductance of full coverage mixed-chain systems is a non-monotonic function of the fraction of long chains.
546 kstocke1 3829 \begin{figure}
547     \includegraphics[width=\linewidth]{figures/Gstacks}
548 kstocke1 3846 \caption{Interfacial thermal conductivity of mixed-chains has a non-monotonic dependence on the fraction of long chains (lower panels). At low long chain coverage, the solvent escape rate dominates the heat transfer process, while the solvent-thiolate orientational order dominates at high long chain coverage (upper panels).}
549 kstocke1 3829 \label{fig:Gstacks}
550     \end{figure}
551 kstocke1 3846 Mixtures of butanethiolate/decanethiolate (n = 4, 10) have a peak interfacial conductance for 25\%/75\% short/long chains. At this short/long chain fraction, the orientational ordering of the solvent molecules and chains becomes a dominant effect while the solvent escape rate sharply decreases. Mixtures of butanethiolate/dodecanethiolate (n = 4, 12) have a peak interfacial conductance for 12.5\%/87.5\% short/long chains. The solvent-thiolate orientational ordering reaches its maximum value at this short/long chain ratio. Long chain fractions of over $0.5$ for the n = 4, 12 system are very well ordered, but this effect is tempered by the exceptionally slow solvent escape rate.
552 kstocke1 3815
553 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
554     % **DISCUSSION**
555     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556     \section{Discussion}
557    
558 gezelter 3819 In the mixed chain-length simulations, solvent molecules can become
559     temporarily trapped or entangled with the thiolate chains. Their
560     residence in close proximity to the higher temperature environment
561     close to the surface allows them to carry heat away from the surface
562     quite efficiently. There are two aspects of this behavior that are
563     relevant to thermal conductance of the interface: the residence time
564     of solvent molecules in the thiolate layer, and the orientational
565     ordering of the C-C chains as a mechanism for transferring vibrational
566 kstocke1 3847 energy to these entrapped solvent molecules. These competing effects are responsible for the non-monotonic dependence of the interfacial thermal conductance on the fraction of long chain capping agents.
567 gezelter 3819
568     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
569     % RESIDENCE TIME
570     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
571 kstocke1 3841 \subsection{Mobility of solvent in the interfacial layer}
572 gezelter 3819
573     We use a simple survival correlation function, $C(t)$, to quantify the
574     residence time of a solvent molecule in the long thiolate chain
575     layer. This function correlates the identity of all hexane molecules
576     within the $z$-coordinate range of the thiolate layer at two separate
577     times. If the solvent molecule is present at both times, the
578     configuration contributes a $1$, while the absence of the molecule at
579     the later time indicates that the solvent molecule has migrated into
580     the bulk, and this configuration contributes a $0$. A steep decay in
581     $C(t)$ indicates a high turnover rate of solvent molecules from the
582 kstocke1 3837 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
583 kstocke1 3815 \begin{equation}
584 gezelter 3825 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
585 kstocke1 3821 \label{eq:mobility}
586 kstocke1 3815 \end{equation}
587 gezelter 3825 where T is the length of the simulation. This is a direct measure of
588     the rate at which solvent molecules entangled in the thiolate layer
589 kstocke1 3843 can escape into the bulk. As $k_{escape} \rightarrow 0$, the
590 gezelter 3825 solvent has become permanently trapped in the thiolate layer. In
591 kstocke1 3841 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
592 gezelter 3825 decreases as the percentage of long thiolate chains increases.
593 gezelter 3819
594 kstocke1 3832 \begin{figure}
595 kstocke1 3837 \includegraphics[width=\linewidth]{figures/timelapse}
596 kstocke1 3849 \caption{Images of 25\%/75\% C$_4$/C$_{12}$ (top panel) and 75\%/25\% C$_4$/C$_{12}$ (bottom panel) systems at the beginning and end of a 3 ns simulation. Initial interfacial solvent molecules are colored light blue. The solvent-thiolate disorder and interfacial solvent turnover of the 25\%/75\% short/long system stand in stark contrast to the highly ordered but entrapped interfacial solvent of the 75\%/25\% system.}
597 kstocke1 3832 \label{fig:timelapse}
598     \end{figure}
599 gezelter 3819
600     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
601     % ORDER PARAMETER
602     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
603    
604     \subsection{Vibrational coupling via orientational ordering}
605    
606     As the fraction of long-chain thiolates becomes large, the entrapped
607     solvent molecules must find specific orientations relative to the mean
608     orientation of the thiolate chains. This configuration allows for
609     efficient thermal energy exchange between the thiolate alkyl chain and
610     the solvent molecules.
611    
612 kstocke1 3821 To quantify this cooperative
613 gezelter 3819 ordering, we computed the orientational order parameters and director
614     axes for both the thiolate chains and for the entrapped solvent. The
615     director axis can be easily obtained by diagonalization of the order
616     parameter tensor,
617     \begin{equation}
618     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
619     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
620 kstocke1 3815 \end{equation}
621 gezelter 3819 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
622     the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
623     For both kinds of molecules, the $\mathbf{e}$ vector is defined using
624     the terminal atoms of the chains.
625    
626     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
627     traditionally used to obtain orientational order parameter, while the
628     eigenvector corresponding to the order parameter yields the director
629     axis ($\mathbf{d}(t)$) which defines the average direction of
630     molecular alignment at any time $t$. The overlap between the director
631     axes of the thiolates and the entrapped solvent is time-averaged,
632 kstocke1 3815 \begin{equation}
633 gezelter 3825 \bar{d} = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
634     \mathbf{d}_{solvent} \left( t \right) \rangle_t
635 kstocke1 3815 \label{eq:orientation}
636     \end{equation}
637 kstocke1 3840 and reported in figure \ref{fig:Gstacks}.
638 kstocke1 3815
639 gezelter 3819 Once the solvent molecules have picked up thermal energy from the
640     thiolates, they carry heat away from the gold as they diffuse back
641     into the bulk solvent. When the percentage of long chains decreases,
642     the tails of the long chains are much more disordered and do not
643     provide structured channels for the solvent to fill.
644 kstocke1 3815
645 gezelter 3819 Although the alignment of the chains with the entrapped solvent is one
646     possible mechanism for the non-monotonic increase in the conductance
647     as a function
648    
649    
650 kstocke1 3843 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
651     % **CONCLUSIONS**
652     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
653     \section{Conclusions}
654    
655    
656    
657 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
658     % **ACKNOWLEDGMENTS**
659     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
660     \section*{Acknowledgments}
661     Support for this project was provided by the
662     National Science Foundation under grant CHE-0848243. Computational
663     time was provided by the Center for Research Computing (CRC) at the
664     University of Notre Dame.
665    
666     \newpage
667    
668     \bibliography{thiolsRNEMD}
669    
670     \end{doublespace}
671     \end{document}
672    
673    
674    
675    
676    
677    
678    
679    

Properties

Name Value
svn:executable *