ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3849
Committed: Fri Dec 21 19:07:54 2012 UTC (12 years, 7 months ago) by kstocke1
Content type: application/x-tex
File size: 35813 byte(s)
Log Message:

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     al.} have shown that metal-particle-based ``nanofluids'' have
81     thermal conductivities that match Maxwell
82     predictions.\cite{Eapen:2007th} The likely cause of previously
83     reported non-Maxwell
84 gezelter 3803 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
85     is percolation networks of nanoparticles exchanging energy via the
86     solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
87 gezelter 3848 picture of particle-ligand and particle-solvent interactions in order
88     to understand the thermal behavior of complex fluids. To date, there
89     have been few reported values (either from theory or experiment) for
90     $G$ for ligand-protected nanoparticles embedded in liquids, and there
91     is a significant gap in knowledge about how chemically distinct
92     ligands or protecting groups will affect heat transport from the
93     particles.
94 gezelter 3803
95 gezelter 3848 Experimentally, the thermal properties of various nanostructured
96     interfaces have been investigated by a number of groups. Cahill and
97 gezelter 3803 coworkers studied nanoscale thermal transport from metal
98 gezelter 3848 nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxide
99 gezelter 3803 interfaces, and hydrophilic and hydrophobic interfaces between water
100     and solids with different self-assembled
101 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
102 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
103     hydrocarbon monolayers on gold substrate at the individual molecular
104     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
105     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
106     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
107     et al.} studied the cooling dynamics, which is controlled by thermal
108     interface resistance of glass-embedded metal
109     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
110     normally considered barriers for heat transport, Alper {\it et al.}
111     suggested that specific ligands (capping agents) could completely
112     eliminate this barrier
113     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
114    
115     In previous simulations, we applied a variant of reverse
116     non-equilibrium molecular dynamics (RNEMD) to calculate the
117     interfacial thermal conductance at a metal / organic solvent interface
118     that had been chemically protected by butanethiolate groups. Our
119     calculations suggest an explanation for the very large thermal
120     conductivity at alkanethiol-capped metal surfaces when compared with
121     bare metal/solvent interfaces. Specifically, the chemical bond
122     between the metal and the ligand introduces a vibrational overlap that
123     is not present without the protecting group, and the overlap between
124     the vibrational spectra (metal to ligand, ligand to solvent) provides
125     a mechanism for rapid thermal transport across the interface.
126    
127 gezelter 3848 One notable result of our previous work was the observation of
128     non-monotonic dependence of the thermal conductance on the coverage of
129     a metal surface by a chemical protecting group. Our explanation for
130     this behavior was that gaps in surface coverage allowed solvent to
131     penetrate close to the capping molecules that had been heated by the
132     metal surface, to absorb thermal energy from these molecules, and then
133     diffuse away. The effect of surface coverage was relatively difficult
134     to study as the individual protecting groups have lateral mobility on
135     the surface and can aggregate to form domains on the timescale of the
136     simulation.
137 gezelter 3803
138 gezelter 3848 To prevent lateral mobility of the surface ligands, the current work
139     involves mixed-chain monolayers in which the length mismatch between
140     long and short chains is sufficient to accomodate solvent
141     molecules. These complete (but mixed-chain) surfaces are significantly
142     less prone to surface domain formation, and allow us to further
143     investigate the mechanism of heat transport to the solvent. A thermal
144     flux is created using velocity shearing and scaling reverse
145     non-equilibrium molecular dynamics (VSS-RNEMD), and the resulting
146     temperature profiles are analyzed to yield information about the
147     interfacial thermal conductance.
148 gezelter 3803
149 gezelter 3848
150 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
151     % **METHODOLOGY**
152     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153     \section{Methodology}
154    
155 gezelter 3803 There are many ways to compute bulk transport properties from
156     classical molecular dynamics simulations. Equilibrium Molecular
157     Dynamics (EMD) simulations can be used by computing relevant time
158     correlation functions and assuming that linear response theory
159     holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
160     For some transport properties (notably thermal conductivity), EMD
161     approaches are subject to noise and poor convergence of the relevant
162     correlation functions. Traditional Non-equilibrium Molecular Dynamics
163     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
164     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
165     However, the resulting flux is often difficult to
166     measure. Furthermore, problems arise for NEMD simulations of
167     heterogeneous systems, such as phase-phase boundaries or interfaces,
168     where the type of gradient to enforce at the boundary between
169     materials is unclear.
170    
171     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
172     a different approach in that an unphysical {\it flux} is imposed
173     between different regions or ``slabs'' of the simulation
174     box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
175     system responds by developing a temperature or momentum {\it gradient}
176     between the two regions. Since the amount of the applied flux is known
177     exactly, and the measurement of a gradient is generally less
178     complicated, imposed-flux methods typically take shorter simulation
179     times to obtain converged results for transport properties. The
180     corresponding temperature or velocity gradients which develop in
181     response to the applied flux are then related (via linear response
182     theory) to the transport coefficient of interest. These methods are
183     quite efficient, in that they do not need many trajectories to provide
184     information about transport properties. To date, they have been
185     utilized in computing thermal and mechanical transfer of both
186     homogeneous
187     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
188     well as heterogeneous
189     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
190    
191 gezelter 3822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192     % VSS-RNEMD
193     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194     \subsection{VSS-RNEMD}
195 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
196     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
197     as a sequence of imaginary elastic collisions between particles in
198     regions separated by half of the simulation cell. In each collision,
199     the entire momentum vectors of both particles may be exchanged to
200     generate a thermal flux. Alternatively, a single component of the
201     momentum vectors may be exchanged to generate a shear flux. This
202     algorithm turns out to be quite useful in many simulations of bulk
203     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
204     system away from ideal Maxwell-Boltzmann distributions, and this has
205     undesirable side-effects when the applied flux becomes
206     large.\cite{Maginn:2010}
207    
208 gezelter 3848 The most useful alternative RNEMD approach developed so far utilizes a
209     series of simultaneous velocity shearing and scaling exchanges between
210     the two slabs.\cite{2012MolPh.110..691K} This method provides a set of
211 gezelter 3803 conservation constraints while simultaneously creating a desired flux
212     between the two slabs. Satisfying the constraint equations ensures
213     that the new configurations are sampled from the same NVE ensemble.
214    
215     The VSS moves are applied periodically to scale and shift the particle
216     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
217     $C$) which are separated by half of the simulation box,
218     \begin{displaymath}
219     \begin{array}{rclcl}
220    
221     & \underline{\mathrm{shearing}} & &
222     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
223     \mathbf{v}_i \leftarrow &
224     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
225     \rangle\right) + \langle\mathbf{v}_c\rangle \\
226     \mathbf{v}_j \leftarrow &
227     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
228     \rangle\right) + \langle\mathbf{v}_h\rangle .
229    
230     \end{array}
231     \end{displaymath}
232     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
233     the center of mass velocities in the $C$ and $H$ slabs, respectively.
234     Within the two slabs, particles receive incremental changes or a
235     ``shear'' to their velocities. The amount of shear is governed by the
236     imposed momentum flux, $j_z(\mathbf{p})$
237     \begin{eqnarray}
238     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
239     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
240     \end{eqnarray}
241     where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
242     is the interval between two separate operations.
243    
244     To simultaneously impose a thermal flux ($J_z$) between the slabs we
245     use energy conservation constraints,
246     \begin{eqnarray}
247     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
248     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
249     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
250     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
251     \mathbf{a}_h)^2 \label{vss4}.
252     \label{constraint}
253     \end{eqnarray}
254     Simultaneous solution of these quadratic formulae for the scaling
255     coefficients, $c$ and $h$, will ensure that the simulation samples from
256     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
257     instantaneous translational kinetic energy of each slab. At each time
258     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
259     and $\mathbf{a}_h$, subject to the imposed momentum flux,
260     $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
261     operations do not change the kinetic energy due to orientational
262     degrees of freedom or the potential energy of a system, configurations
263     after the VSS move have exactly the same energy ({\it and} linear
264     momentum) as before the move.
265    
266     As the simulation progresses, the VSS moves are performed on a regular
267     basis, and the system develops a thermal or velocity gradient in
268     response to the applied flux. Using the slope of the temperature or
269     velocity gradient, it is quite simple to obtain of thermal
270     conductivity ($\lambda$),
271     \begin{equation}
272     J_z = -\lambda \frac{\partial T}{\partial z},
273     \end{equation}
274     and shear viscosity ($\eta$),
275     \begin{equation}
276     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
277     \end{equation}
278     Here, the quantities on the left hand side are the imposed flux
279     values, while the slopes are obtained from linear fits to the
280     gradients that develop in the simulated system.
281    
282     The VSS-RNEMD approach is versatile in that it may be used to
283     implement both thermal and shear transport either separately or
284     simultaneously. Perturbations of velocities away from the ideal
285     Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
286     kept to a minimum. This ability to generate simultaneous thermal and
287     shear fluxes has been previously utilized to map out the shear
288     viscosity of SPC/E water over a wide range of temperatures (90~K) with
289 gezelter 3848 a single 1 ns simulation.\cite{2012MolPh.110..691K}
290 gezelter 3803
291 gezelter 3822 \begin{figure}
292     \includegraphics[width=\linewidth]{figures/rnemd}
293     \caption{The VSS-RNEMD approach imposes unphysical transfer of
294     linear momentum or kinetic energy between a ``hot'' slab and a
295     ``cold'' slab in the simulation box. The system responds to this
296     imposed flux by generating velocity or temperature gradients. The
297     slope of the gradients can then be used to compute transport
298     properties (e.g. shear viscosity or thermal conductivity).}
299     \label{fig:rnemd}
300     \end{figure}
301    
302     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303     % INTERFACIAL CONDUCTANCE
304     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
305     \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
306 gezelter 3803 to interfacial transport}
307 kstocke1 3801
308 gezelter 3803 Interfaces between dissimilar materials have transport properties
309     which can be defined as derivatives of the standard transport
310     coefficients in a direction normal to the interface. For example, the
311     {\it interfacial} thermal conductance ($G$) can be thought of as the
312     change in the thermal conductivity ($\lambda$) across the boundary
313     between materials:
314     \begin{align}
315     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
316     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
317     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
318     \label{derivativeG}
319     \end{align}
320     where $z_0$ is the location of the interface between two materials and
321     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
322     to be the $z$ direction from here on). RNEMD simulations, and
323     particularly the VSS-RNEMD approach, function by applying a momentum
324     or thermal flux and watching the gradient response of the
325     material. This means that the {\it interfacial} conductance is a
326     second derivative property which is subject to significant noise and
327     therefore requires extensive sampling. We have been able to
328     demonstrate the use of the second derivative approach to compute
329     interfacial conductance at chemically-modified metal / solvent
330     interfaces. However, a definition of the interfacial conductance in
331     terms of a temperature difference ($\Delta T$) measured at $z_0$,
332     \begin{displaymath}
333     G = \frac{J_z}{\Delta T_{z_0}},
334     \end{displaymath}
335     is useful once the RNEMD approach has generated a stable temperature
336     gap across the interface.
337    
338     \begin{figure}
339     \includegraphics[width=\linewidth]{figures/resistor_series}
340 gezelter 3822 \caption{The inverse of the interfacial thermal conductance, $G$, is
341     the Kapitza resistance, $R_K$. Because the gold / thiolate/
342     solvent interface extends a significant distance from the metal
343     surface, the interfacial resistance $R_K$ can be computed by
344     summing a series of temperature drops between adjacent temperature
345     bins along the $z$ axis.}
346 gezelter 3803 \label{fig:resistor_series}
347     \end{figure}
348    
349     In the particular case we are studying here, there are two interfaces
350     involved in the transfer of heat from the gold slab to the solvent:
351     the gold/thiolate interface and the thiolate/solvent interface. We
352     could treat the temperature on each side of an interface as discrete,
353     making the interfacial conductance the inverse of the Kaptiza
354     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
355     across multiple interfaces, it is useful to think of the interfaces as
356     a set of resistors wired in series. The total resistance is then
357     additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
358     is the inverse of the total resistance, or $G = \frac{1}{\sum_i
359     R_i}$). In the interfacial region, we treat each bin in the
360     VSS-RNEMD temperature profile as a resistor with resistance
361     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
362     the set of resistors which spans the gold/thiolate interface, thiolate
363     chains, and thiolate/solvent interface simplifies to
364 kstocke1 3801 \begin{equation}
365 gezelter 3825 R_{K} = \frac{T_{n}-T_{1}}{J_z},
366 gezelter 3803 \label{eq:finalG}
367     \end{equation}
368     or the temperature difference between the gold side of the
369     gold/thiolate interface and the solvent side of the thiolate/solvent
370     interface over the applied flux.
371 kstocke1 3801
372    
373     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374     % **COMPUTATIONAL DETAILS**
375     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
376     \section{Computational Details}
377    
378 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
379 gezelter 3819 % FORCE-FIELD PARAMETERS
380     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
381     \subsection{Force-Field Parameters}
382 kstocke1 3801
383 gezelter 3819 Our simulations include a number of chemically distinct components.
384     Figure \ref{fig:structures} demonstrates the sites defined for both
385     the {\it n}-hexane and alkanethiolate ligands present in our
386     simulations. Force field parameters are needed for interactions both
387     between the same type of particles and between particles of different
388     species.
389    
390     \begin{figure}
391     \includegraphics[width=\linewidth]{figures/structures}
392 gezelter 3822 \caption{Topologies of the thiolate capping agents and solvent
393     utilized in the simulations. The chemically-distinct sites (S,
394     \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
395     parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
396     \protect\cite{TraPPE-UA.alkanes} and
397     \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
398     atoms were adapted from references
399     \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
400     \protect\cite{hautman:4994}.}
401 gezelter 3819 \label{fig:structures}
402     \end{figure}
403    
404 kstocke1 3829 The Au-Au interactions in the metal lattice slab are described by the
405 gezelter 3819 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
406     potentials include zero-point quantum corrections and are
407     reparametrized for accurate surface energies compared to the
408     Sutton-Chen potentials.\cite{Chen90}
409    
410     For the {\it n}-hexane solvent molecules, the TraPPE-UA
411     parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
412     sites are located at the carbon centers for alkyl groups. Bonding
413     interactions, including bond stretches and bends and torsions, were
414     used for intra-molecular sites closer than 3 bonds. For non-bonded
415     interactions, Lennard-Jones potentials are used. We have previously
416     utilized both united atom (UA) and all-atom (AA) force fields for
417     thermal conductivity in previous work,\cite{} and since the united
418     atom force fields cannot populate the high-frequency modes that are
419     present in AA force fields, they appear to work better for modeling
420     thermal conductivity. The TraPPE-UA model for alkanes is known to
421     predict a slightly lower boiling point than experimental values. This
422     is one of the reasons we used a lower average temperature (200K) for
423     our simulations.
424    
425     The TraPPE-UA force field includes parameters for thiol
426     molecules\cite{TraPPE-UA.thiols} which were used for the
427     alkanethiolate molecules in our simulations. To derive suitable
428     parameters for butanethiol adsorbed on Au(111) surfaces, we adopted
429     the S parameters from Luedtke and Landman\cite{landman:1998} and
430     modified the parameters for the CTS atom to maintain charge neutrality
431     in the molecule.
432    
433     To describe the interactions between metal (Au) and non-metal atoms,
434     we refer to an adsorption study of alkyl thiols on gold surfaces by
435     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
436     Lennard-Jones form of potential parameters for the interaction between
437     Au and pseudo-atoms CH$_x$ and S based on a well-established and
438     widely-used effective potential of Hautman and Klein for the Au(111)
439     surface.\cite{hautman:4994} As our simulations require the gold slab
440     to be flexible to accommodate thermal excitation, the pair-wise form
441 kstocke1 3832 of potentials they developed was used for our study.
442 kstocke1 3821
443     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
444     % SIMULATION PROTOCOL
445     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
446     \subsection{Simulation Protocol}
447    
448     We have implemented the VSS-RNEMD algorithm in OpenMD, our
449 kstocke1 3842 group molecular dynamics code. A 1188 atom gold slab was
450     equilibrated at 1 atm and 200 K. The periodic box was then expanded
451 kstocke1 3845 in the $z$-direction to expose two Au(111) faces on either side of the 11-atom thick slab.
452 kstocke1 3821
453     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.
454    
455 kstocke1 3845 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.
456 kstocke1 3821
457     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.
458    
459     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.
460    
461     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.
462    
463    
464 kstocke1 3801
465     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
466     % **RESULTS**
467     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468     \section{Results}
469    
470    
471 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
472     % CHAIN LENGTH
473     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
474     \subsection{Effect of Chain Length}
475    
476 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.
477 kstocke1 3830 \begin{longtable}{p{4cm} p{3cm}}
478     \caption{Computed interfacial thermal conductance (G) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
479     \\
480 kstocke1 3831 \centering {\bf Chain Length (n)} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
481 kstocke1 3830 \endhead
482     \hline
483     \endfoot
484 kstocke1 3832 \centering bare gold & \centering\arraybackslash 30.2 \\
485     \centering 4 & \centering\arraybackslash 59.4 \\
486     \centering 6 & \centering\arraybackslash 60.2 \\
487     \centering 8 & \centering\arraybackslash 61.0 \\
488     \centering 10 & \centering\arraybackslash 58.2 \\
489     \centering 12 & \centering\arraybackslash 58.8
490 kstocke1 3830 \label{table:chainlengthG}
491     \end{longtable}
492 kstocke1 3801
493 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
494     % MIXED CHAINS
495     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
496     \subsection{Effect of Mixed Chain Lengths}
497    
498 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.
499 kstocke1 3829 \begin{figure}
500     \includegraphics[width=\linewidth]{figures/Gstacks}
501 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).}
502 kstocke1 3829 \label{fig:Gstacks}
503     \end{figure}
504 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.
505 kstocke1 3815
506 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
507     % **DISCUSSION**
508     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
509     \section{Discussion}
510    
511 gezelter 3819 In the mixed chain-length simulations, solvent molecules can become
512     temporarily trapped or entangled with the thiolate chains. Their
513     residence in close proximity to the higher temperature environment
514     close to the surface allows them to carry heat away from the surface
515     quite efficiently. There are two aspects of this behavior that are
516     relevant to thermal conductance of the interface: the residence time
517     of solvent molecules in the thiolate layer, and the orientational
518     ordering of the C-C chains as a mechanism for transferring vibrational
519 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.
520 gezelter 3819
521     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
522     % RESIDENCE TIME
523     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
524 kstocke1 3841 \subsection{Mobility of solvent in the interfacial layer}
525 gezelter 3819
526     We use a simple survival correlation function, $C(t)$, to quantify the
527     residence time of a solvent molecule in the long thiolate chain
528     layer. This function correlates the identity of all hexane molecules
529     within the $z$-coordinate range of the thiolate layer at two separate
530     times. If the solvent molecule is present at both times, the
531     configuration contributes a $1$, while the absence of the molecule at
532     the later time indicates that the solvent molecule has migrated into
533     the bulk, and this configuration contributes a $0$. A steep decay in
534     $C(t)$ indicates a high turnover rate of solvent molecules from the
535 kstocke1 3837 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
536 kstocke1 3815 \begin{equation}
537 gezelter 3825 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
538 kstocke1 3821 \label{eq:mobility}
539 kstocke1 3815 \end{equation}
540 gezelter 3825 where T is the length of the simulation. This is a direct measure of
541     the rate at which solvent molecules entangled in the thiolate layer
542 kstocke1 3843 can escape into the bulk. As $k_{escape} \rightarrow 0$, the
543 gezelter 3825 solvent has become permanently trapped in the thiolate layer. In
544 kstocke1 3841 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
545 gezelter 3825 decreases as the percentage of long thiolate chains increases.
546 gezelter 3819
547 kstocke1 3832 \begin{figure}
548 kstocke1 3837 \includegraphics[width=\linewidth]{figures/timelapse}
549 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.}
550 kstocke1 3832 \label{fig:timelapse}
551     \end{figure}
552 gezelter 3819
553     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
554     % ORDER PARAMETER
555     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556    
557     \subsection{Vibrational coupling via orientational ordering}
558    
559     As the fraction of long-chain thiolates becomes large, the entrapped
560     solvent molecules must find specific orientations relative to the mean
561     orientation of the thiolate chains. This configuration allows for
562     efficient thermal energy exchange between the thiolate alkyl chain and
563     the solvent molecules.
564    
565 kstocke1 3821 To quantify this cooperative
566 gezelter 3819 ordering, we computed the orientational order parameters and director
567     axes for both the thiolate chains and for the entrapped solvent. The
568     director axis can be easily obtained by diagonalization of the order
569     parameter tensor,
570     \begin{equation}
571     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
572     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
573 kstocke1 3815 \end{equation}
574 gezelter 3819 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
575     the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
576     For both kinds of molecules, the $\mathbf{e}$ vector is defined using
577     the terminal atoms of the chains.
578    
579     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
580     traditionally used to obtain orientational order parameter, while the
581     eigenvector corresponding to the order parameter yields the director
582     axis ($\mathbf{d}(t)$) which defines the average direction of
583     molecular alignment at any time $t$. The overlap between the director
584     axes of the thiolates and the entrapped solvent is time-averaged,
585 kstocke1 3815 \begin{equation}
586 gezelter 3825 \bar{d} = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
587     \mathbf{d}_{solvent} \left( t \right) \rangle_t
588 kstocke1 3815 \label{eq:orientation}
589     \end{equation}
590 kstocke1 3840 and reported in figure \ref{fig:Gstacks}.
591 kstocke1 3815
592 gezelter 3819 Once the solvent molecules have picked up thermal energy from the
593     thiolates, they carry heat away from the gold as they diffuse back
594     into the bulk solvent. When the percentage of long chains decreases,
595     the tails of the long chains are much more disordered and do not
596     provide structured channels for the solvent to fill.
597 kstocke1 3815
598 gezelter 3819 Although the alignment of the chains with the entrapped solvent is one
599     possible mechanism for the non-monotonic increase in the conductance
600     as a function
601    
602    
603 kstocke1 3843 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
604     % **CONCLUSIONS**
605     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
606     \section{Conclusions}
607    
608    
609    
610 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
611     % **ACKNOWLEDGMENTS**
612     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
613     \section*{Acknowledgments}
614     Support for this project was provided by the
615     National Science Foundation under grant CHE-0848243. Computational
616     time was provided by the Center for Research Computing (CRC) at the
617     University of Notre Dame.
618    
619     \newpage
620    
621     \bibliography{thiolsRNEMD}
622    
623     \end{doublespace}
624     \end{document}
625    
626    
627    
628    
629    
630    
631    
632    

Properties

Name Value
svn:executable *