ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3851
Committed: Fri Dec 21 20:45:48 2012 UTC (11 years, 6 months ago) by kstocke1
Content type: application/x-tex
File size: 38970 byte(s)
Log Message:
Almost done

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 3851 We report on simulations of heat conduction from Au(111) / hexane interfaces in which the surface has been protected by a mix of short and long chain alkanethiolate ligands. A variant of reverse non-equilibrium molecular dynamics (RNEMD) was used to create a thermal flux between the metal and solvent, and thermal conductance was computed using the resulting thermal profiles of the interface. We find a non-monotonic dependence of the interfacial thermal conductance on the fraction of long chains present at the interface, and correlate this behavior to solvent ordering and escape from the thiolate layer immediately in contact with the metal surface.
56     Our results suggest that a mixed vibrational transfer / convection model is
57     necessary to explain the features of heat transfer at this interface. The
58     alignment of the solvent chains with the ordered ligand allows rapid transfer
59     of energy to the trapped solvent (and this is the dominant feature for ordered
60     ligand layers). Diffusion of the vibrationally excited solvent into the bulk
61     also plays a significant role when the ligands are less tightly packed.
62    
63 kstocke1 3801 \end{abstract}
64    
65     \newpage
66    
67     %\narrowtext
68    
69     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70     % **INTRODUCTION**
71     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72     \section{Introduction}
73    
74 gezelter 3848 The structural details of the interfaces of metal nanoparticles
75     determines how energy flows between these particles and their
76     surroundings. Understanding this energy flow is essential in designing
77     and functionalizing metallic nanoparticles for plasmonic photothermal
78     therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw}
79     which rely on the ability of metallic nanoparticles to absorb light in
80     the near-IR, a portion of the spectrum in which living tissue is very
81     nearly transparent. The relevant physical property controlling the
82     transfer of this energy as heat into the surrounding tissue is the
83     interfacial thermal conductance, $G$, which can be somewhat difficult
84     to determine experimentally.\cite{Wilson:2002uq,Plech:2005kx}
85 gezelter 3803
86     Metallic particles have also been proposed for use in highly efficient
87 gezelter 3848 thermal-transfer fluids, although careful experiments by Eapen {\it et
88 gezelter 3850 al.} have shown that metal-particle-based nanofluids have thermal
89     conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
90     likely cause of previously reported non-Maxwell
91 gezelter 3803 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
92     is percolation networks of nanoparticles exchanging energy via the
93 gezelter 3850 solvent,\cite{Eapen:2007mw} so it is important to get a detailed
94     molecular picture of particle-ligand and ligand-solvent interactions
95     in order to understand the thermal behavior of complex fluids. To
96     date, there have been few reported values (either from theory or
97     experiment) for $G$ for ligand-protected nanoparticles embedded in
98     liquids, and there is a significant gap in knowledge about how
99     chemically distinct ligands or protecting groups will affect heat
100     transport from the particles.
101 gezelter 3803
102 gezelter 3848 Experimentally, the thermal properties of various nanostructured
103     interfaces have been investigated by a number of groups. Cahill and
104 gezelter 3850 coworkers studied thermal transport from metal nanoparticle/fluid
105     interfaces, epitaxial TiN/single crystal oxide interfaces, and
106     hydrophilic and hydrophobic interfaces between water and solids with
107     different self-assembled
108 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
109 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
110     hydrocarbon monolayers on gold substrate at the individual molecular
111     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
112     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
113     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
114     et al.} studied the cooling dynamics, which is controlled by thermal
115     interface resistance of glass-embedded metal
116     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
117     normally considered barriers for heat transport, Alper {\it et al.}
118 gezelter 3850 have suggested that specific ligands (capping agents) could completely
119 gezelter 3803 eliminate this barrier
120     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
121    
122     In previous simulations, we applied a variant of reverse
123     non-equilibrium molecular dynamics (RNEMD) to calculate the
124     interfacial thermal conductance at a metal / organic solvent interface
125     that had been chemically protected by butanethiolate groups. Our
126 gezelter 3850 calculations suggested an explanation for the very large thermal
127 gezelter 3803 conductivity at alkanethiol-capped metal surfaces when compared with
128     bare metal/solvent interfaces. Specifically, the chemical bond
129     between the metal and the ligand introduces a vibrational overlap that
130     is not present without the protecting group, and the overlap between
131     the vibrational spectra (metal to ligand, ligand to solvent) provides
132     a mechanism for rapid thermal transport across the interface.
133    
134 gezelter 3850 A notable result of the previous simulations was the non-monotonic
135     dependence of $G$ on the fractional coverage of the metal surface by
136     the chemical protecting group. Gaps in surface coverage allow the
137     solvent molecules come into direct contact with ligands that had been
138     heated by the metal surface, absorb thermal energy from the ligands,
139     and then diffuse away. Quantifying the role of surface coverage is
140     difficult as the ligands have lateral mobility on the surface and can
141     aggregate to form domains on the timescale of the simulation.
142 gezelter 3803
143 gezelter 3850 To isolate this effect without worrying about lateral mobility of the
144     surface ligands, the current work involves mixed-chain-length
145     monolayers in which the length mismatch between long and short chains
146 kstocke1 3851 is sufficient to accommodate solvent molecules. These completely
147 gezelter 3850 covered (but mixed-chain) surfaces are significantly less prone to
148     surface domain formation, and allow us to further investigate the
149     mechanism of heat transport to the solvent. A thermal flux is created
150     using velocity shearing and scaling reverse non-equilibrium molecular
151     dynamics (VSS-RNEMD), and the resulting temperature profiles are
152     analyzed to yield information about the interfacial thermal
153     conductance.
154 gezelter 3803
155 gezelter 3848
156 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
157     % **METHODOLOGY**
158     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
159     \section{Methodology}
160    
161 gezelter 3803 There are many ways to compute bulk transport properties from
162     classical molecular dynamics simulations. Equilibrium Molecular
163     Dynamics (EMD) simulations can be used by computing relevant time
164     correlation functions and assuming that linear response theory
165     holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
166     For some transport properties (notably thermal conductivity), EMD
167     approaches are subject to noise and poor convergence of the relevant
168     correlation functions. Traditional Non-equilibrium Molecular Dynamics
169     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
170     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
171     However, the resulting flux is often difficult to
172     measure. Furthermore, problems arise for NEMD simulations of
173     heterogeneous systems, such as phase-phase boundaries or interfaces,
174     where the type of gradient to enforce at the boundary between
175     materials is unclear.
176    
177     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
178     a different approach in that an unphysical {\it flux} is imposed
179     between different regions or ``slabs'' of the simulation
180 gezelter 3850 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
181 gezelter 3803 system responds by developing a temperature or momentum {\it gradient}
182     between the two regions. Since the amount of the applied flux is known
183     exactly, and the measurement of a gradient is generally less
184     complicated, imposed-flux methods typically take shorter simulation
185     times to obtain converged results for transport properties. The
186     corresponding temperature or velocity gradients which develop in
187     response to the applied flux are then related (via linear response
188     theory) to the transport coefficient of interest. These methods are
189     quite efficient, in that they do not need many trajectories to provide
190     information about transport properties. To date, they have been
191     utilized in computing thermal and mechanical transfer of both
192     homogeneous
193     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
194     well as heterogeneous
195     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
196    
197 gezelter 3822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
198     % VSS-RNEMD
199     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
200     \subsection{VSS-RNEMD}
201 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
202     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
203     as a sequence of imaginary elastic collisions between particles in
204     regions separated by half of the simulation cell. In each collision,
205     the entire momentum vectors of both particles may be exchanged to
206     generate a thermal flux. Alternatively, a single component of the
207     momentum vectors may be exchanged to generate a shear flux. This
208     algorithm turns out to be quite useful in many simulations of bulk
209     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
210     system away from ideal Maxwell-Boltzmann distributions, and this has
211     undesirable side-effects when the applied flux becomes
212     large.\cite{Maginn:2010}
213    
214 gezelter 3848 The most useful alternative RNEMD approach developed so far utilizes a
215     series of simultaneous velocity shearing and scaling exchanges between
216 gezelter 3850 the two slabs.\cite{Kuang2012} This method provides a set of
217 gezelter 3803 conservation constraints while simultaneously creating a desired flux
218     between the two slabs. Satisfying the constraint equations ensures
219     that the new configurations are sampled from the same NVE ensemble.
220    
221     The VSS moves are applied periodically to scale and shift the particle
222     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
223     $C$) which are separated by half of the simulation box,
224     \begin{displaymath}
225     \begin{array}{rclcl}
226    
227     & \underline{\mathrm{shearing}} & &
228     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
229     \mathbf{v}_i \leftarrow &
230     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
231     \rangle\right) + \langle\mathbf{v}_c\rangle \\
232     \mathbf{v}_j \leftarrow &
233     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
234     \rangle\right) + \langle\mathbf{v}_h\rangle .
235    
236     \end{array}
237     \end{displaymath}
238     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
239     the center of mass velocities in the $C$ and $H$ slabs, respectively.
240     Within the two slabs, particles receive incremental changes or a
241     ``shear'' to their velocities. The amount of shear is governed by the
242     imposed momentum flux, $j_z(\mathbf{p})$
243     \begin{eqnarray}
244     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
245     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
246     \end{eqnarray}
247     where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
248     is the interval between two separate operations.
249    
250     To simultaneously impose a thermal flux ($J_z$) between the slabs we
251     use energy conservation constraints,
252     \begin{eqnarray}
253     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
254     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
255     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
256     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
257     \mathbf{a}_h)^2 \label{vss4}.
258     \label{constraint}
259     \end{eqnarray}
260     Simultaneous solution of these quadratic formulae for the scaling
261     coefficients, $c$ and $h$, will ensure that the simulation samples from
262     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
263     instantaneous translational kinetic energy of each slab. At each time
264     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
265     and $\mathbf{a}_h$, subject to the imposed momentum flux,
266     $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
267     operations do not change the kinetic energy due to orientational
268     degrees of freedom or the potential energy of a system, configurations
269     after the VSS move have exactly the same energy ({\it and} linear
270     momentum) as before the move.
271    
272     As the simulation progresses, the VSS moves are performed on a regular
273     basis, and the system develops a thermal or velocity gradient in
274     response to the applied flux. Using the slope of the temperature or
275     velocity gradient, it is quite simple to obtain of thermal
276     conductivity ($\lambda$),
277     \begin{equation}
278     J_z = -\lambda \frac{\partial T}{\partial z},
279     \end{equation}
280     and shear viscosity ($\eta$),
281     \begin{equation}
282     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
283     \end{equation}
284     Here, the quantities on the left hand side are the imposed flux
285     values, while the slopes are obtained from linear fits to the
286     gradients that develop in the simulated system.
287    
288     The VSS-RNEMD approach is versatile in that it may be used to
289     implement both thermal and shear transport either separately or
290     simultaneously. Perturbations of velocities away from the ideal
291     Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
292     kept to a minimum. This ability to generate simultaneous thermal and
293     shear fluxes has been previously utilized to map out the shear
294     viscosity of SPC/E water over a wide range of temperatures (90~K) with
295 gezelter 3850 a single 1 ns simulation.\cite{Kuang2012}
296 gezelter 3803
297 gezelter 3822 \begin{figure}
298     \includegraphics[width=\linewidth]{figures/rnemd}
299     \caption{The VSS-RNEMD approach imposes unphysical transfer of
300     linear momentum or kinetic energy between a ``hot'' slab and a
301     ``cold'' slab in the simulation box. The system responds to this
302     imposed flux by generating velocity or temperature gradients. The
303     slope of the gradients can then be used to compute transport
304     properties (e.g. shear viscosity or thermal conductivity).}
305     \label{fig:rnemd}
306     \end{figure}
307    
308     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
309     % INTERFACIAL CONDUCTANCE
310     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311     \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
312 gezelter 3803 to interfacial transport}
313 kstocke1 3801
314 gezelter 3803 Interfaces between dissimilar materials have transport properties
315     which can be defined as derivatives of the standard transport
316     coefficients in a direction normal to the interface. For example, the
317     {\it interfacial} thermal conductance ($G$) can be thought of as the
318     change in the thermal conductivity ($\lambda$) across the boundary
319     between materials:
320     \begin{align}
321     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
322     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
323     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
324     \label{derivativeG}
325     \end{align}
326     where $z_0$ is the location of the interface between two materials and
327     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
328     to be the $z$ direction from here on). RNEMD simulations, and
329     particularly the VSS-RNEMD approach, function by applying a momentum
330     or thermal flux and watching the gradient response of the
331     material. This means that the {\it interfacial} conductance is a
332     second derivative property which is subject to significant noise and
333     therefore requires extensive sampling. We have been able to
334     demonstrate the use of the second derivative approach to compute
335     interfacial conductance at chemically-modified metal / solvent
336     interfaces. However, a definition of the interfacial conductance in
337     terms of a temperature difference ($\Delta T$) measured at $z_0$,
338     \begin{displaymath}
339     G = \frac{J_z}{\Delta T_{z_0}},
340     \end{displaymath}
341     is useful once the RNEMD approach has generated a stable temperature
342     gap across the interface.
343    
344     \begin{figure}
345     \includegraphics[width=\linewidth]{figures/resistor_series}
346 gezelter 3822 \caption{The inverse of the interfacial thermal conductance, $G$, is
347     the Kapitza resistance, $R_K$. Because the gold / thiolate/
348     solvent interface extends a significant distance from the metal
349     surface, the interfacial resistance $R_K$ can be computed by
350     summing a series of temperature drops between adjacent temperature
351     bins along the $z$ axis.}
352 gezelter 3803 \label{fig:resistor_series}
353     \end{figure}
354    
355     In the particular case we are studying here, there are two interfaces
356     involved in the transfer of heat from the gold slab to the solvent:
357 gezelter 3850 the metal/thiolate interface and the thiolate/solvent interface. We
358 gezelter 3803 could treat the temperature on each side of an interface as discrete,
359     making the interfacial conductance the inverse of the Kaptiza
360     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
361     across multiple interfaces, it is useful to think of the interfaces as
362     a set of resistors wired in series. The total resistance is then
363     additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
364     is the inverse of the total resistance, or $G = \frac{1}{\sum_i
365     R_i}$). In the interfacial region, we treat each bin in the
366     VSS-RNEMD temperature profile as a resistor with resistance
367     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
368     the set of resistors which spans the gold/thiolate interface, thiolate
369     chains, and thiolate/solvent interface simplifies to
370 kstocke1 3801 \begin{equation}
371 gezelter 3825 R_{K} = \frac{T_{n}-T_{1}}{J_z},
372 gezelter 3803 \label{eq:finalG}
373     \end{equation}
374     or the temperature difference between the gold side of the
375     gold/thiolate interface and the solvent side of the thiolate/solvent
376     interface over the applied flux.
377 kstocke1 3801
378    
379     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
380     % **COMPUTATIONAL DETAILS**
381     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
382     \section{Computational Details}
383    
384 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385 gezelter 3819 % FORCE-FIELD PARAMETERS
386     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
387     \subsection{Force-Field Parameters}
388 kstocke1 3801
389 gezelter 3819 Our simulations include a number of chemically distinct components.
390     Figure \ref{fig:structures} demonstrates the sites defined for both
391     the {\it n}-hexane and alkanethiolate ligands present in our
392 gezelter 3850 simulations.
393 gezelter 3819
394     \begin{figure}
395     \includegraphics[width=\linewidth]{figures/structures}
396 gezelter 3822 \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 gezelter 3819 \label{fig:structures}
406     \end{figure}
407    
408 kstocke1 3829 The Au-Au interactions in the metal lattice slab are described by the
409 gezelter 3819 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 gezelter 3850 interactions, Lennard-Jones potentials were used. We have previously
420 gezelter 3819 utilized both united atom (UA) and all-atom (AA) force fields for
421 gezelter 3850 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
422 gezelter 3819 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 gezelter 3850 our simulations.
428 gezelter 3819
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 kstocke1 3832 of potentials they developed was used for our study.
446 kstocke1 3821
447     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
448     % SIMULATION PROTOCOL
449     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
450     \subsection{Simulation Protocol}
451    
452 gezelter 3850 We have implemented the VSS-RNEMD algorithm in OpenMD, our group
453     molecular dynamics code.\cite{openmd} An 1188 atom gold slab was
454     equilibrated at 1 atm and 200 K. The periodic box was then expanded
455     in the $z$-direction to expose two Au(111) faces on either side of the
456     11-atom thick slab.
457 kstocke1 3821
458 gezelter 3850 A full monolayer of thiolates (1/3 the number of surface gold atoms)
459     was placed on three-fold hollow sites on the gold interfaces. The
460     effect of thiolate binding sites on the thermal conductance was tested
461     by placing thiolates at both fcc and hcp hollow sites. No appreciable
462     difference in the temperature profile was noted due to the location of
463     thiolate binding.
464 kstocke1 3821
465 gezelter 3850 To test the role of thiolate chain length on interfacial thermal
466     conductance, full coverages of each of five chain lengths were tested:
467     butanethiolate (C$_4$), hexanethiolate (C$_6$), octanethiolate
468     (C$_8$), decanethiolate (C$_{10}$), and dodecanethiolate
469     (C$_{12}$). To test the effect of mixed chain lengths, full coverages
470     of C$_4$ / C$_{10}$ and C$_4$ / C$_{12}$ mixtures were created in
471     short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and
472     87.5/12.5. The short and long chains were placed on the surface hollow
473     sites in a random configuration.
474 kstocke1 3821
475 gezelter 3850 The simulation box $z$-dimension was set to roughly double the length
476     of the gold/thiolate block. Hexane solvent molecules were placed in
477 kstocke1 3851 the vacant portion of the box using the packmol algorithm. Figure \ref{fig:timelapse} shows two of the mixed chain length systems both before and after the RNEMD simulation.
478 kstocke1 3821
479 kstocke1 3851 \begin{figure}
480     \includegraphics[width=\linewidth]{figures/timelapse}
481     \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.}
482     \label{fig:timelapse}
483     \end{figure}
484    
485 gezelter 3850 The system was equilibrated to 220 K in the canonical (NVT) ensemble,
486     allowing the thiolates and solvent to warm gradually. Pressure
487     correction to 1 atm was done using an isobaric-isothermal (NPT)
488     integrator that allowed expansion or contraction only in the $z$
489     direction, maintaining the crystalline structure of the gold as close
490     to the bulk result as possible. The diagonal elements of the pressure
491     tensor were monitored during the pressure equilibration stage. If the
492     $xx$ and/or $yy$ elements had a mean above zero throughout the
493     simulation -- indicating residual surface tension in the plane of the
494     gold slab -- an additional short NPT equilibration step was performed
495     allowing all box dimensions to change. Once the pressure was stable
496     at 1 atm, a final equilibration stage was performed at constant
497     temperature. All systems were equilibrated in the microcanonical (NVE)
498     ensemble before proceeding with the VSS-RNEMD and data collection
499     stages.
500 kstocke1 3821
501 gezelter 3850 A kinetic energy flux was applied using VSS-RNEMD. The total
502     simulation time was 3 nanoseconds, with velocity scaling occurring
503     every 10 femtoseconds. The hot slab was centered in the gold and the
504     cold slab was placed in the center of the solvent region. The entire
505     system has a (time-averaged) temperature of 220 K, with a temperature
506     difference between the hot and cold slabs of approximately 30 K. The
507     average temperature and kinetic energy flux were selected to prevent
508     solvent freezing (or glass formation) and to not allow the thiolates
509     to bury in the gold slab. The Au-S interaction has a deep potential
510     energy well, which allows sulfur atoms to embed into the gold slab at
511     temperatures above 250 K. Simulation conditions were chosen to avoid
512     both of these undesirable situations.
513 kstocke1 3821
514 gezelter 3850 Temperature profiles of the system were created by dividing the box
515     into $\sim$ 3 \AA \, bins along the $z$ axis and recording the average
516     temperature of each bin.
517 kstocke1 3821
518 gezelter 3850
519 kstocke1 3801
520     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521     % **RESULTS**
522     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
523     \section{Results}
524    
525 kstocke1 3851 The solvent, hexane, is a straight chain flexible alkane that is structurally
526     similar to the thiolate alkane tails. Previous work has shown that UA models
527     of hexane and butanethiolate have a high degree of vibrational
528     overlap.\cite{kuang:AuThl} This overlap provides a mechanism for thermal
529     energy conduction from the thiolates to the solvent. Indeed, we observe that
530     the interfacial conductance is twice as large with the thiolate monolayers (of
531     all chain lengths) than with the bare metal surface.
532 kstocke1 3801
533 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
534     % CHAIN LENGTH
535     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
536     \subsection{Effect of Chain Length}
537    
538 kstocke1 3851 We examined full coverages of five chain lengths, n = 4, 6, 8, 10, and 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.
539 kstocke1 3830 \begin{longtable}{p{4cm} p{3cm}}
540 kstocke1 3851 \caption{Computed interfacial thermal conductance ($G$) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
541 kstocke1 3830 \\
542 kstocke1 3831 \centering {\bf Chain Length (n)} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
543 kstocke1 3830 \endhead
544     \hline
545     \endfoot
546 kstocke1 3851 \centering bare metal & \centering\arraybackslash 30.2 \\
547 kstocke1 3832 \centering 4 & \centering\arraybackslash 59.4 \\
548     \centering 6 & \centering\arraybackslash 60.2 \\
549     \centering 8 & \centering\arraybackslash 61.0 \\
550     \centering 10 & \centering\arraybackslash 58.2 \\
551     \centering 12 & \centering\arraybackslash 58.8
552 kstocke1 3830 \label{table:chainlengthG}
553     \end{longtable}
554 kstocke1 3801
555 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556     % MIXED CHAINS
557     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
558     \subsection{Effect of Mixed Chain Lengths}
559    
560 kstocke1 3851 Previous simulations have demonstrated a 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 as the fraction of long chains was varied is shown in figure \ref{fig:Gstacks}. Note that as in the previous study, $G$ is depends on solvent accessibility to thermally excited ligands. Our simulations indicate a similar (but not as dramatic) non-monotonic dependence on the fraction of long chains.
561 kstocke1 3829 \begin{figure}
562     \includegraphics[width=\linewidth]{figures/Gstacks}
563 kstocke1 3851 \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 dominates in systems with higher fractions of long chains (upper panels).}
564 kstocke1 3829 \label{fig:Gstacks}
565     \end{figure}
566 kstocke1 3815
567 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
568     % **DISCUSSION**
569     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
570     \section{Discussion}
571 gezelter 3819 In the mixed chain-length simulations, solvent molecules can become
572     temporarily trapped or entangled with the thiolate chains. Their
573     residence in close proximity to the higher temperature environment
574     close to the surface allows them to carry heat away from the surface
575     quite efficiently. There are two aspects of this behavior that are
576     relevant to thermal conductance of the interface: the residence time
577 kstocke1 3851 of solvent molecules in the thiolate layer, and the alignment
578     of the C-C chains as a mechanism for transferring vibrational
579     energy to these entrapped solvent molecules. To quantify these competing effects, we have computed solvent escape rates from the ligand layer as well as a joint orientational order parameter between the trapped solvent and the thiolate ligands.
580 gezelter 3819
581     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
582     % RESIDENCE TIME
583     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
584 kstocke1 3841 \subsection{Mobility of solvent in the interfacial layer}
585 gezelter 3819
586 kstocke1 3851 We use a simple survival correlation function, $C(t)$, to measure the
587     residence time of a solvent molecule in the thiolate
588 gezelter 3819 layer. This function correlates the identity of all hexane molecules
589     within the $z$-coordinate range of the thiolate layer at two separate
590     times. If the solvent molecule is present at both times, the
591     configuration contributes a $1$, while the absence of the molecule at
592     the later time indicates that the solvent molecule has migrated into
593     the bulk, and this configuration contributes a $0$. A steep decay in
594     $C(t)$ indicates a high turnover rate of solvent molecules from the
595 kstocke1 3837 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
596 kstocke1 3815 \begin{equation}
597 gezelter 3825 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
598 kstocke1 3821 \label{eq:mobility}
599 kstocke1 3815 \end{equation}
600 gezelter 3825 where T is the length of the simulation. This is a direct measure of
601 kstocke1 3851 the rate at which solvent molecules initially entangled in the thiolate layer
602 kstocke1 3843 can escape into the bulk. As $k_{escape} \rightarrow 0$, the
603 gezelter 3825 solvent has become permanently trapped in the thiolate layer. In
604 kstocke1 3841 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
605 gezelter 3825 decreases as the percentage of long thiolate chains increases.
606 gezelter 3819
607     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
608     % ORDER PARAMETER
609     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610    
611     \subsection{Vibrational coupling via orientational ordering}
612    
613 kstocke1 3851 As the fraction of long-chain thiolates increases, the entrapped
614 gezelter 3819 solvent molecules must find specific orientations relative to the mean
615 kstocke1 3851 orientation of the thiolate chains. This alignment allows for
616 gezelter 3819 efficient thermal energy exchange between the thiolate alkyl chain and
617     the solvent molecules.
618    
619 kstocke1 3851 To measure this cooperative ordering, we computed the orientational order
620     parameters and director axes for both the thiolate chains and for the
621     entrapped solvent. The director axis can be easily obtained by diagonalization
622     of the order parameter tensor,
623 gezelter 3819 \begin{equation}
624     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
625     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
626 kstocke1 3815 \end{equation}
627 gezelter 3819 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
628     the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
629     For both kinds of molecules, the $\mathbf{e}$ vector is defined using
630     the terminal atoms of the chains.
631    
632     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
633     traditionally used to obtain orientational order parameter, while the
634     eigenvector corresponding to the order parameter yields the director
635     axis ($\mathbf{d}(t)$) which defines the average direction of
636     molecular alignment at any time $t$. The overlap between the director
637     axes of the thiolates and the entrapped solvent is time-averaged,
638 kstocke1 3815 \begin{equation}
639 kstocke1 3851 \langle d \rangle = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
640 gezelter 3825 \mathbf{d}_{solvent} \left( t \right) \rangle_t
641 kstocke1 3815 \label{eq:orientation}
642     \end{equation}
643 kstocke1 3840 and reported in figure \ref{fig:Gstacks}.
644 kstocke1 3815
645 gezelter 3819 Once the solvent molecules have picked up thermal energy from the
646     thiolates, they carry heat away from the gold as they diffuse back
647     into the bulk solvent. When the percentage of long chains decreases,
648     the tails of the long chains are much more disordered and do not
649     provide structured channels for the solvent to fill.
650 kstocke1 3815
651 kstocke1 3851 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 sharply decreases. 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 n = 4, 12 system are well ordered, but this effect is tempered by the exceptionally slow solvent escape rate.
652    
653 kstocke1 3843 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
654     % **CONCLUSIONS**
655     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
656     \section{Conclusions}
657 kstocke1 3851 Our results suggest that a mixed vibrational transfer / convection model is necessary to explain the features of heat transfer at this interface. The alignment of the solvent chains with the ordered ligand allows rapid transfer of energy to the trapped solvent (and this is the dominant feature for ordered ligand layers). Diffusion of the vibrationally excited solvent into the bulk also plays a significant role when the ligands are less tightly packed.
658 kstocke1 3843
659 kstocke1 3851 In the language of earlier continuum approaches to interfacial
660     conductance,\cite{RevModPhys.61.605} the alignment of the chains is an
661     important factor in the transfer of phonons from the thiolate layer to the
662     trapped solvent. The aligned solvent and thiolate chains have nearly identical
663     acoustic impedances and the phonons can scatter directly into a solvent
664     molecule that has been forced into alignment. When the entrapped solvent has
665     more configurations available, the likelihood of an impedance mismatch is
666     higher, and the phonon scatters into the solvent with lower
667     efficiency. The fractional coverage of the long chains is therefore a simple
668     way of tuning the acoustic mismatch between the thiolate layer and the hexane
669     solvent.
670 kstocke1 3843
671 kstocke1 3851 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.
672 kstocke1 3843
673 kstocke1 3851 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 solvent mobility back to the bulk, we would expect a significant jump in the interfacial conductance. One possible way to do 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 into the bulk.
674    
675 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
676 kstocke1 3851 % **ACKNOWLEDGMENTS**
677 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
678 kstocke1 3851 \section*{Acknowledgments}
679 kstocke1 3801
680 kstocke1 3851 We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
681     this project was provided by the National Science Foundation under grant
682     CHE-0848243. Computational time was provided by the Center for Research
683     Computing (CRC) at the University of Notre Dame.
684    
685 kstocke1 3801 \newpage
686    
687     \bibliography{thiolsRNEMD}
688    
689     \end{doublespace}
690     \end{document}
691    
692    
693    
694    
695    
696    
697    
698    

Properties

Name Value
svn:executable *