ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3841
Committed: Thu Dec 20 22:04:57 2012 UTC (11 years, 8 months ago) by kstocke1
Content type: application/x-tex
File size: 35159 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 3803 The structural and dynamical details of interfaces between metal
67     nanoparticles and solvents determines how energy flows between these
68     particles and their surroundings. Understanding this energy flow is
69     essential in designing and functionalizing metallic nanoparticles for
70     plasmonic photothermal
71 gezelter 3819 therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw} which rely on the ability of metallic
72 gezelter 3803 nanoparticles to absorb light in the near-IR, a portion of the
73     spectrum in which living tissue is very nearly transparent. The
74     principle of this therapy is to pump the particles at high power at
75     the plasmon resonance and to allow heat dissipation to kill targeted
76     (e.g. cancerous) cells. The relevant physical property controlling
77     this transfer of energy is the interfacial thermal conductance, $G$,
78     which can be somewhat difficult to determine
79     experimentally.\cite{Wilson:2002uq,Plech:2005kx}
80    
81     Metallic particles have also been proposed for use in highly efficient
82     thermal-transfer fluids, although careful experiments by Eapen {\it et al.}
83     have shown that metal-particle-based ``nanofluids'' have thermal
84     conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
85     likely cause of previously reported non-Maxwell
86     behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
87     is percolation networks of nanoparticles exchanging energy via the
88     solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
89     picture of particle-solvent interactions in order to understand the
90     thermal behavior of complex fluids. To date, there have been few
91     reported values (either from theory or experiment) for $G$ for
92     ligand-protected nanoparticles embedded in liquids, and there is a
93     significant gap in knowledge about how chemically distinct ligands or
94     protecting groups will affect heat transport from the particles.
95    
96     The thermal properties of various nanostructured interfaces have been
97     investigated experimentally by a number of groups: Cahill and
98     coworkers studied nanoscale thermal transport from metal
99     nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxides
100     interfaces, and hydrophilic and hydrophobic interfaces between water
101     and solids with different self-assembled
102 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
103 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
104     hydrocarbon monolayers on gold substrate at the individual molecular
105     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
106     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
107     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
108     et al.} studied the cooling dynamics, which is controlled by thermal
109     interface resistance of glass-embedded metal
110     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
111     normally considered barriers for heat transport, Alper {\it et al.}
112     suggested that specific ligands (capping agents) could completely
113     eliminate this barrier
114     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
115    
116     In previous simulations, we applied a variant of reverse
117     non-equilibrium molecular dynamics (RNEMD) to calculate the
118     interfacial thermal conductance at a metal / organic solvent interface
119     that had been chemically protected by butanethiolate groups. Our
120     calculations suggest an explanation for the very large thermal
121     conductivity at alkanethiol-capped metal surfaces when compared with
122     bare metal/solvent interfaces. Specifically, the chemical bond
123     between the metal and the ligand introduces a vibrational overlap that
124     is not present without the protecting group, and the overlap between
125     the vibrational spectra (metal to ligand, ligand to solvent) provides
126     a mechanism for rapid thermal transport across the interface.
127    
128     One interesting result of our previous work was the observation of
129     {\it non-monotonic dependence} of the thermal conductance on the
130     coverage of a metal surface by a chemical protecting group. Our
131     explanation for this behavior was that gaps in surface coverage
132     allowed solvent to penetrate close to the capping molecules that had
133     been heated by the metal surface, to absorb thermal energy from these
134     molecules, and then diffuse away. The effect of surface coverage is
135     relatively difficult to study as the individual protecting groups have
136     lateral mobility on the surface and can aggregate to form domains on
137     the timescale of the simulation.
138    
139     The work reported here involves the use of velocity shearing and
140     scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD) to
141     study surfaces composed of mixed-length chains which collectively form
142     a complete monolayer on the surfaces. These complete (but
143     mixed-chain) surfaces are significantly less prone to surface domain
144     formation, and would allow us to further investigate the mechanism of
145     heat transport to the solvent.
146    
147 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148     % **METHODOLOGY**
149     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
150     \section{Methodology}
151    
152 gezelter 3803 There are many ways to compute bulk transport properties from
153     classical molecular dynamics simulations. Equilibrium Molecular
154     Dynamics (EMD) simulations can be used by computing relevant time
155     correlation functions and assuming that linear response theory
156     holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
157     For some transport properties (notably thermal conductivity), EMD
158     approaches are subject to noise and poor convergence of the relevant
159     correlation functions. Traditional Non-equilibrium Molecular Dynamics
160     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
161     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
162     However, the resulting flux is often difficult to
163     measure. Furthermore, problems arise for NEMD simulations of
164     heterogeneous systems, such as phase-phase boundaries or interfaces,
165     where the type of gradient to enforce at the boundary between
166     materials is unclear.
167    
168     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
169     a different approach in that an unphysical {\it flux} is imposed
170     between different regions or ``slabs'' of the simulation
171     box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
172     system responds by developing a temperature or momentum {\it gradient}
173     between the two regions. Since the amount of the applied flux is known
174     exactly, and the measurement of a gradient is generally less
175     complicated, imposed-flux methods typically take shorter simulation
176     times to obtain converged results for transport properties. The
177     corresponding temperature or velocity gradients which develop in
178     response to the applied flux are then related (via linear response
179     theory) to the transport coefficient of interest. These methods are
180     quite efficient, in that they do not need many trajectories to provide
181     information about transport properties. To date, they have been
182     utilized in computing thermal and mechanical transfer of both
183     homogeneous
184     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
185     well as heterogeneous
186     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
187    
188 gezelter 3822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189     % VSS-RNEMD
190     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191     \subsection{VSS-RNEMD}
192 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
193     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
194     as a sequence of imaginary elastic collisions between particles in
195     regions separated by half of the simulation cell. In each collision,
196     the entire momentum vectors of both particles may be exchanged to
197     generate a thermal flux. Alternatively, a single component of the
198     momentum vectors may be exchanged to generate a shear flux. This
199     algorithm turns out to be quite useful in many simulations of bulk
200     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
201     system away from ideal Maxwell-Boltzmann distributions, and this has
202     undesirable side-effects when the applied flux becomes
203     large.\cite{Maginn:2010}
204    
205     Instead of having momentum exchange between {\it individual particles}
206     in each slab, the NIVS algorithm applies velocity scaling to all of
207     the selected particles in both slabs.\cite{Kuang:2010uq} A combination
208     of linear momentum, kinetic energy, and flux constraint equations
209     governs the amount of velocity scaling performed at each step. NIVS
210     has been shown to be very effective at producing thermal gradients and
211     for computing thermal conductivities, particularly for heterogeneous
212     interfaces. Although the NIVS algorithm can also be applied to impose
213     a directional momentum flux, thermal anisotropy was observed in
214     relatively high flux simulations, and the method is not suitable for
215     imposing a shear flux or for computing shear viscosities.
216    
217     The most useful RNEMD
218     approach developed so far utilizes a series of simultaneous velocity
219     shearing and scaling exchanges between the two
220     slabs.\cite{2012MolPh.110..691K} This method provides a set of
221     conservation constraints while simultaneously creating a desired flux
222     between the two slabs. Satisfying the constraint equations ensures
223     that the new configurations are sampled from the same NVE ensemble.
224    
225     The VSS moves are applied periodically to scale and shift the particle
226     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
227     $C$) which are separated by half of the simulation box,
228     \begin{displaymath}
229     \begin{array}{rclcl}
230    
231     & \underline{\mathrm{shearing}} & &
232     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
233     \mathbf{v}_i \leftarrow &
234     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
235     \rangle\right) + \langle\mathbf{v}_c\rangle \\
236     \mathbf{v}_j \leftarrow &
237     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
238     \rangle\right) + \langle\mathbf{v}_h\rangle .
239    
240     \end{array}
241     \end{displaymath}
242     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
243     the center of mass velocities in the $C$ and $H$ slabs, respectively.
244     Within the two slabs, particles receive incremental changes or a
245     ``shear'' to their velocities. The amount of shear is governed by the
246     imposed momentum flux, $j_z(\mathbf{p})$
247     \begin{eqnarray}
248     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
249     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
250     \end{eqnarray}
251     where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
252     is the interval between two separate operations.
253    
254     To simultaneously impose a thermal flux ($J_z$) between the slabs we
255     use energy conservation constraints,
256     \begin{eqnarray}
257     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
258     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
259     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
260     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
261     \mathbf{a}_h)^2 \label{vss4}.
262     \label{constraint}
263     \end{eqnarray}
264     Simultaneous solution of these quadratic formulae for the scaling
265     coefficients, $c$ and $h$, will ensure that the simulation samples from
266     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
267     instantaneous translational kinetic energy of each slab. At each time
268     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
269     and $\mathbf{a}_h$, subject to the imposed momentum flux,
270     $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
271     operations do not change the kinetic energy due to orientational
272     degrees of freedom or the potential energy of a system, configurations
273     after the VSS move have exactly the same energy ({\it and} linear
274     momentum) as before the move.
275    
276     As the simulation progresses, the VSS moves are performed on a regular
277     basis, and the system develops a thermal or velocity gradient in
278     response to the applied flux. Using the slope of the temperature or
279     velocity gradient, it is quite simple to obtain of thermal
280     conductivity ($\lambda$),
281     \begin{equation}
282     J_z = -\lambda \frac{\partial T}{\partial z},
283     \end{equation}
284     and shear viscosity ($\eta$),
285     \begin{equation}
286     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
287     \end{equation}
288     Here, the quantities on the left hand side are the imposed flux
289     values, while the slopes are obtained from linear fits to the
290     gradients that develop in the simulated system.
291    
292     The VSS-RNEMD approach is versatile in that it may be used to
293     implement both thermal and shear transport either separately or
294     simultaneously. Perturbations of velocities away from the ideal
295     Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
296     kept to a minimum. This ability to generate simultaneous thermal and
297     shear fluxes has been previously utilized to map out the shear
298     viscosity of SPC/E water over a wide range of temperatures (90~K) with
299     a {\it single 1 ns simulation}.\cite{2012MolPh.110..691K}
300    
301 gezelter 3822 \begin{figure}
302     \includegraphics[width=\linewidth]{figures/rnemd}
303     \caption{The VSS-RNEMD approach imposes unphysical transfer of
304     linear momentum or kinetic energy between a ``hot'' slab and a
305     ``cold'' slab in the simulation box. The system responds to this
306     imposed flux by generating velocity or temperature gradients. The
307     slope of the gradients can then be used to compute transport
308     properties (e.g. shear viscosity or thermal conductivity).}
309     \label{fig:rnemd}
310     \end{figure}
311    
312     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
313     % INTERFACIAL CONDUCTANCE
314     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
315     \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
316 gezelter 3803 to interfacial transport}
317 kstocke1 3801
318 gezelter 3803 Interfaces between dissimilar materials have transport properties
319     which can be defined as derivatives of the standard transport
320     coefficients in a direction normal to the interface. For example, the
321     {\it interfacial} thermal conductance ($G$) can be thought of as the
322     change in the thermal conductivity ($\lambda$) across the boundary
323     between materials:
324     \begin{align}
325     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
326     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
327     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
328     \label{derivativeG}
329     \end{align}
330     where $z_0$ is the location of the interface between two materials and
331     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
332     to be the $z$ direction from here on). RNEMD simulations, and
333     particularly the VSS-RNEMD approach, function by applying a momentum
334     or thermal flux and watching the gradient response of the
335     material. This means that the {\it interfacial} conductance is a
336     second derivative property which is subject to significant noise and
337     therefore requires extensive sampling. We have been able to
338     demonstrate the use of the second derivative approach to compute
339     interfacial conductance at chemically-modified metal / solvent
340     interfaces. However, a definition of the interfacial conductance in
341     terms of a temperature difference ($\Delta T$) measured at $z_0$,
342     \begin{displaymath}
343     G = \frac{J_z}{\Delta T_{z_0}},
344     \end{displaymath}
345     is useful once the RNEMD approach has generated a stable temperature
346     gap across the interface.
347    
348     \begin{figure}
349     \includegraphics[width=\linewidth]{figures/resistor_series}
350 gezelter 3822 \caption{The inverse of the interfacial thermal conductance, $G$, is
351     the Kapitza resistance, $R_K$. Because the gold / thiolate/
352     solvent interface extends a significant distance from the metal
353     surface, the interfacial resistance $R_K$ can be computed by
354     summing a series of temperature drops between adjacent temperature
355     bins along the $z$ axis.}
356 gezelter 3803 \label{fig:resistor_series}
357     \end{figure}
358    
359     In the particular case we are studying here, there are two interfaces
360     involved in the transfer of heat from the gold slab to the solvent:
361     the gold/thiolate interface and the thiolate/solvent interface. We
362     could treat the temperature on each side of an interface as discrete,
363     making the interfacial conductance the inverse of the Kaptiza
364     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
365     across multiple interfaces, it is useful to think of the interfaces as
366     a set of resistors wired in series. The total resistance is then
367     additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
368     is the inverse of the total resistance, or $G = \frac{1}{\sum_i
369     R_i}$). In the interfacial region, we treat each bin in the
370     VSS-RNEMD temperature profile as a resistor with resistance
371     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
372     the set of resistors which spans the gold/thiolate interface, thiolate
373     chains, and thiolate/solvent interface simplifies to
374 kstocke1 3801 \begin{equation}
375 gezelter 3825 R_{K} = \frac{T_{n}-T_{1}}{J_z},
376 gezelter 3803 \label{eq:finalG}
377     \end{equation}
378     or the temperature difference between the gold side of the
379     gold/thiolate interface and the solvent side of the thiolate/solvent
380     interface over the applied flux.
381 kstocke1 3801
382    
383     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
384     % **COMPUTATIONAL DETAILS**
385     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
386     \section{Computational Details}
387    
388 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
389 gezelter 3819 % FORCE-FIELD PARAMETERS
390     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
391     \subsection{Force-Field Parameters}
392 kstocke1 3801
393 gezelter 3819 Our simulations include a number of chemically distinct components.
394     Figure \ref{fig:structures} demonstrates the sites defined for both
395     the {\it n}-hexane and alkanethiolate ligands present in our
396     simulations. Force field parameters are needed for interactions both
397     between the same type of particles and between particles of different
398     species.
399    
400     \begin{figure}
401     \includegraphics[width=\linewidth]{figures/structures}
402 gezelter 3822 \caption{Topologies of the thiolate capping agents and solvent
403     utilized in the simulations. The chemically-distinct sites (S,
404     \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
405     parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
406     \protect\cite{TraPPE-UA.alkanes} and
407     \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
408     atoms were adapted from references
409     \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
410     \protect\cite{hautman:4994}.}
411 gezelter 3819 \label{fig:structures}
412     \end{figure}
413    
414 kstocke1 3829 The Au-Au interactions in the metal lattice slab are described by the
415 gezelter 3819 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
416     potentials include zero-point quantum corrections and are
417     reparametrized for accurate surface energies compared to the
418     Sutton-Chen potentials.\cite{Chen90}
419    
420     For the {\it n}-hexane solvent molecules, the TraPPE-UA
421     parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
422     sites are located at the carbon centers for alkyl groups. Bonding
423     interactions, including bond stretches and bends and torsions, were
424     used for intra-molecular sites closer than 3 bonds. For non-bonded
425     interactions, Lennard-Jones potentials are used. We have previously
426     utilized both united atom (UA) and all-atom (AA) force fields for
427     thermal conductivity in previous work,\cite{} and since the united
428     atom force fields cannot populate the high-frequency modes that are
429     present in AA force fields, they appear to work better for modeling
430     thermal conductivity. The TraPPE-UA model for alkanes is known to
431     predict a slightly lower boiling point than experimental values. This
432     is one of the reasons we used a lower average temperature (200K) for
433     our simulations.
434    
435     The TraPPE-UA force field includes parameters for thiol
436     molecules\cite{TraPPE-UA.thiols} which were used for the
437     alkanethiolate molecules in our simulations. To derive suitable
438     parameters for butanethiol adsorbed on Au(111) surfaces, we adopted
439     the S parameters from Luedtke and Landman\cite{landman:1998} and
440     modified the parameters for the CTS atom to maintain charge neutrality
441     in the molecule.
442    
443     To describe the interactions between metal (Au) and non-metal atoms,
444     we refer to an adsorption study of alkyl thiols on gold surfaces by
445     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
446     Lennard-Jones form of potential parameters for the interaction between
447     Au and pseudo-atoms CH$_x$ and S based on a well-established and
448     widely-used effective potential of Hautman and Klein for the Au(111)
449     surface.\cite{hautman:4994} As our simulations require the gold slab
450     to be flexible to accommodate thermal excitation, the pair-wise form
451 kstocke1 3832 of potentials they developed was used for our study.
452 kstocke1 3821
453     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
454     % SIMULATION PROTOCOL
455     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
456     \subsection{Simulation Protocol}
457    
458     We have implemented the VSS-RNEMD algorithm in OpenMD, our
459     group molecular dynamics code. A gold slab 11 atoms thick was
460     equilibrated at 1 atm and 200 K. The periodic box was expanded
461     in the z direction to expose two Au(111) faces.
462    
463     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.
464    
465     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.
466    
467     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.
468    
469     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.
470    
471     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.
472    
473    
474 kstocke1 3801
475     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
476     % **RESULTS**
477     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478     \section{Results}
479    
480    
481 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
482     % CHAIN LENGTH
483     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
484     \subsection{Effect of Chain Length}
485    
486 kstocke1 3830 We examined full coverages of five chain lengths, n = 4, 6, 8, 10, 12. As shown in table \ref{table:chainlengthG}, the trend of interfacial conductance is mostly flat as a function of chain length, indicating that the length of the thiolate alkyl chains does not play a significant role in the transport of heat across the gold/thiolate and thiolate/solvent interfaces. 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. 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.
487     \begin{longtable}{p{4cm} p{3cm}}
488     \caption{Computed interfacial thermal conductance (G) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
489     \\
490 kstocke1 3831 \centering {\bf Chain Length (n)} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
491 kstocke1 3830 \endhead
492     \hline
493     \endfoot
494 kstocke1 3832 \centering bare gold & \centering\arraybackslash 30.2 \\
495     \centering 4 & \centering\arraybackslash 59.4 \\
496     \centering 6 & \centering\arraybackslash 60.2 \\
497     \centering 8 & \centering\arraybackslash 61.0 \\
498     \centering 10 & \centering\arraybackslash 58.2 \\
499     \centering 12 & \centering\arraybackslash 58.8
500 kstocke1 3830 \label{table:chainlengthG}
501     \end{longtable}
502 kstocke1 3801
503 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
504     % MIXED CHAINS
505     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
506     \subsection{Effect of Mixed Chain Lengths}
507    
508 kstocke1 3815 Previous work demonstrated that for butanethiolate monolayers on a Au(111) surface, the interfacial conductance was a non-monotonic function of the percent coverage. This is believed to be due to enhanced solvent-thiolate coupling through greater penetration of solvent molecules into the thiolate layer. At lower coverages, hexane solvent can more easily line up lengthwise with the thiolate tails by fitting into gaps between the thiolates. However, a side effect of low coverages is surface aggregation of the thiolates. To simulate the effect of low coverages while preventing aggregation we maintain 100\% thiolate coverage while varying the proportions of short (butanethiolate, n = 4) and long (decanethiolate, n = 10 or dodecanethiolate, n = 12). In systems where there is a mix of short and long chain thiolates, interfacial conductance is a non-monotonic function of the percent of long chains. The depth of the gaps between the long chains is $n_{long} - n_{short}$, which has implications for the ability of the hexane solvent to fill in the gaps between the long chains.
509 kstocke1 3829
510     \begin{figure}
511     \includegraphics[width=\linewidth]{figures/Gstacks}
512     \caption{}
513     \label{fig:Gstacks}
514     \end{figure}
515 kstocke1 3801
516 kstocke1 3815 \subsubsection{Butanethiolate/Decanethiolate}
517     Mixtures of butanethiolate/decanethiolate (n = 4, 10) have a peak interfacial condutance for 25\%/75\% short/long chains.
518 kstocke1 3801
519 kstocke1 3815 \subsubsection{Butanethiolate/Dodecanethiolate}
520 kstocke1 3821 Mixtures of butanethiolate/dodecanethiolate (n = 4, 12) have a peak interfacial conductance for 12.5\%/87.5\% short/long chains.
521 kstocke1 3815
522    
523 kstocke1 3801
524    
525     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
526     % **DISCUSSION**
527     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528     \section{Discussion}
529    
530 gezelter 3819 In the mixed chain-length simulations, solvent molecules can become
531     temporarily trapped or entangled with the thiolate chains. Their
532     residence in close proximity to the higher temperature environment
533     close to the surface allows them to carry heat away from the surface
534     quite efficiently. There are two aspects of this behavior that are
535     relevant to thermal conductance of the interface: the residence time
536     of solvent molecules in the thiolate layer, and the orientational
537     ordering of the C-C chains as a mechanism for transferring vibrational
538     energy to these entrapped solvent molecules.
539    
540     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541     % RESIDENCE TIME
542     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
543 kstocke1 3841 \subsection{Mobility of solvent in the interfacial layer}
544 gezelter 3819
545     We use a simple survival correlation function, $C(t)$, to quantify the
546     residence time of a solvent molecule in the long thiolate chain
547     layer. This function correlates the identity of all hexane molecules
548     within the $z$-coordinate range of the thiolate layer at two separate
549     times. If the solvent molecule is present at both times, the
550     configuration contributes a $1$, while the absence of the molecule at
551     the later time indicates that the solvent molecule has migrated into
552     the bulk, and this configuration contributes a $0$. A steep decay in
553     $C(t)$ indicates a high turnover rate of solvent molecules from the
554 kstocke1 3837 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
555 kstocke1 3815 \begin{equation}
556 gezelter 3825 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
557 kstocke1 3821 \label{eq:mobility}
558 kstocke1 3815 \end{equation}
559 gezelter 3825 where T is the length of the simulation. This is a direct measure of
560     the rate at which solvent molecules entangled in the thiolate layer
561     can escape into the bulk. As $k_{escape} \rightarrow \infty$, the
562     solvent has become permanently trapped in the thiolate layer. In
563 kstocke1 3841 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
564 gezelter 3825 decreases as the percentage of long thiolate chains increases.
565 gezelter 3819
566 kstocke1 3832 \begin{figure}
567 kstocke1 3837 \includegraphics[width=\linewidth]{figures/timelapse}
568 kstocke1 3832 \caption{}
569     \label{fig:timelapse}
570     \end{figure}
571 gezelter 3819
572     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
573     % ORDER PARAMETER
574     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
575    
576     \subsection{Vibrational coupling via orientational ordering}
577    
578     As the fraction of long-chain thiolates becomes large, the entrapped
579     solvent molecules must find specific orientations relative to the mean
580     orientation of the thiolate chains. This configuration allows for
581     efficient thermal energy exchange between the thiolate alkyl chain and
582     the solvent molecules.
583    
584 kstocke1 3821 To quantify this cooperative
585 gezelter 3819 ordering, we computed the orientational order parameters and director
586     axes for both the thiolate chains and for the entrapped solvent. The
587     director axis can be easily obtained by diagonalization of the order
588     parameter tensor,
589     \begin{equation}
590     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
591     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
592 kstocke1 3815 \end{equation}
593 gezelter 3819 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
594     the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
595     For both kinds of molecules, the $\mathbf{e}$ vector is defined using
596     the terminal atoms of the chains.
597    
598     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
599     traditionally used to obtain orientational order parameter, while the
600     eigenvector corresponding to the order parameter yields the director
601     axis ($\mathbf{d}(t)$) which defines the average direction of
602     molecular alignment at any time $t$. The overlap between the director
603     axes of the thiolates and the entrapped solvent is time-averaged,
604 kstocke1 3815 \begin{equation}
605 gezelter 3825 \bar{d} = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
606     \mathbf{d}_{solvent} \left( t \right) \rangle_t
607 kstocke1 3815 \label{eq:orientation}
608     \end{equation}
609 kstocke1 3840 and reported in figure \ref{fig:Gstacks}.
610 kstocke1 3815
611 gezelter 3819 Once the solvent molecules have picked up thermal energy from the
612     thiolates, they carry heat away from the gold as they diffuse back
613     into the bulk solvent. When the percentage of long chains decreases,
614     the tails of the long chains are much more disordered and do not
615     provide structured channels for the solvent to fill.
616 kstocke1 3815
617 gezelter 3819 Although the alignment of the chains with the entrapped solvent is one
618     possible mechanism for the non-monotonic increase in the conductance
619     as a function
620    
621    
622 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
623     % **ACKNOWLEDGMENTS**
624     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
625     \section*{Acknowledgments}
626     Support for this project was provided by the
627     National Science Foundation under grant CHE-0848243. Computational
628     time was provided by the Center for Research Computing (CRC) at the
629     University of Notre Dame.
630    
631     \newpage
632    
633     \bibliography{thiolsRNEMD}
634    
635     \end{doublespace}
636     \end{document}
637    
638    
639    
640    
641    
642    
643    
644    

Properties

Name Value
svn:executable *