ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3821
Committed: Tue Dec 18 00:41:09 2012 UTC (11 years, 9 months ago) by kstocke1
Content type: application/x-tex
File size: 33202 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     \usepackage{graphicx}
10     \usepackage{multirow}
11     \usepackage[square, comma, sort&compress]{natbib}
12     \usepackage{url}
13     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
14     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
15     9.0in \textwidth 6.5in \brokenpenalty=10000
16    
17     % double space list of tables and figures
18     %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
19     \setlength{\abovecaptionskip}{20 pt}
20     \setlength{\belowcaptionskip}{30 pt}
21    
22     \bibpunct{}{}{,}{s}{}{;}
23     \bibliographystyle{achemso}
24    
25     \begin{document}
26    
27 gezelter 3819 \title{Simulations of heat conduction at thiolate-capped gold
28     surfaces: The role of chain length and solvent penetration}
29 kstocke1 3801
30 gezelter 3803 \author{Kelsey M. Stocker and J. Daniel
31     Gezelter\footnote{Corresponding author. \ Electronic mail:
32     gezelter@nd.edu} \\
33     251 Nieuwland Science Hall, \\
34 kstocke1 3801 Department of Chemistry and Biochemistry,\\
35     University of Notre Dame\\
36     Notre Dame, Indiana 46556}
37    
38     \date{\today}
39    
40     \maketitle
41    
42     \begin{doublespace}
43    
44     \begin{abstract}
45 kstocke1 3815
46 kstocke1 3801 \end{abstract}
47    
48     \newpage
49    
50     %\narrowtext
51    
52     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
53     % **INTRODUCTION**
54     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55     \section{Introduction}
56    
57 gezelter 3803 The structural and dynamical details of interfaces between metal
58     nanoparticles and solvents determines how energy flows between these
59     particles and their surroundings. Understanding this energy flow is
60     essential in designing and functionalizing metallic nanoparticles for
61     plasmonic photothermal
62 gezelter 3819 therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw} which rely on the ability of metallic
63 gezelter 3803 nanoparticles to absorb light in the near-IR, a portion of the
64     spectrum in which living tissue is very nearly transparent. The
65     principle of this therapy is to pump the particles at high power at
66     the plasmon resonance and to allow heat dissipation to kill targeted
67     (e.g. cancerous) cells. The relevant physical property controlling
68     this transfer of energy is the interfacial thermal conductance, $G$,
69     which can be somewhat difficult to determine
70     experimentally.\cite{Wilson:2002uq,Plech:2005kx}
71    
72     Metallic particles have also been proposed for use in highly efficient
73     thermal-transfer fluids, although careful experiments by Eapen {\it et al.}
74     have shown that metal-particle-based ``nanofluids'' have thermal
75     conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
76     likely cause of previously reported non-Maxwell
77     behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
78     is percolation networks of nanoparticles exchanging energy via the
79     solvent,\cite{Eapen:2007mw} so it is vital to get a detailed molecular
80     picture of particle-solvent interactions in order to understand the
81     thermal behavior of complex fluids. To date, there have been few
82     reported values (either from theory or experiment) for $G$ for
83     ligand-protected nanoparticles embedded in liquids, and there is a
84     significant gap in knowledge about how chemically distinct ligands or
85     protecting groups will affect heat transport from the particles.
86    
87     The thermal properties of various nanostructured interfaces have been
88     investigated experimentally by a number of groups: Cahill and
89     coworkers studied nanoscale thermal transport from metal
90     nanoparticle/fluid interfaces, to epitaxial TiN/single crystal oxides
91     interfaces, and hydrophilic and hydrophobic interfaces between water
92     and solids with different self-assembled
93 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
94 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
95     hydrocarbon monolayers on gold substrate at the individual molecular
96     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
97     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
98     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
99     et al.} studied the cooling dynamics, which is controlled by thermal
100     interface resistance of glass-embedded metal
101     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
102     normally considered barriers for heat transport, Alper {\it et al.}
103     suggested that specific ligands (capping agents) could completely
104     eliminate this barrier
105     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
106    
107     In previous simulations, we applied a variant of reverse
108     non-equilibrium molecular dynamics (RNEMD) to calculate the
109     interfacial thermal conductance at a metal / organic solvent interface
110     that had been chemically protected by butanethiolate groups. Our
111     calculations suggest an explanation for the very large thermal
112     conductivity at alkanethiol-capped metal surfaces when compared with
113     bare metal/solvent interfaces. Specifically, the chemical bond
114     between the metal and the ligand introduces a vibrational overlap that
115     is not present without the protecting group, and the overlap between
116     the vibrational spectra (metal to ligand, ligand to solvent) provides
117     a mechanism for rapid thermal transport across the interface.
118    
119     One interesting result of our previous work was the observation of
120     {\it non-monotonic dependence} of the thermal conductance on the
121     coverage of a metal surface by a chemical protecting group. Our
122     explanation for this behavior was that gaps in surface coverage
123     allowed solvent to penetrate close to the capping molecules that had
124     been heated by the metal surface, to absorb thermal energy from these
125     molecules, and then diffuse away. The effect of surface coverage is
126     relatively difficult to study as the individual protecting groups have
127     lateral mobility on the surface and can aggregate to form domains on
128     the timescale of the simulation.
129    
130     The work reported here involves the use of velocity shearing and
131     scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD) to
132     study surfaces composed of mixed-length chains which collectively form
133     a complete monolayer on the surfaces. These complete (but
134     mixed-chain) surfaces are significantly less prone to surface domain
135     formation, and would allow us to further investigate the mechanism of
136     heat transport to the solvent.
137    
138 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
139     % **METHODOLOGY**
140     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
141     \section{Methodology}
142    
143 gezelter 3803 There are many ways to compute bulk transport properties from
144     classical molecular dynamics simulations. Equilibrium Molecular
145     Dynamics (EMD) simulations can be used by computing relevant time
146     correlation functions and assuming that linear response theory
147     holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
148     For some transport properties (notably thermal conductivity), EMD
149     approaches are subject to noise and poor convergence of the relevant
150     correlation functions. Traditional Non-equilibrium Molecular Dynamics
151     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
152     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
153     However, the resulting flux is often difficult to
154     measure. Furthermore, problems arise for NEMD simulations of
155     heterogeneous systems, such as phase-phase boundaries or interfaces,
156     where the type of gradient to enforce at the boundary between
157     materials is unclear.
158    
159     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
160     a different approach in that an unphysical {\it flux} is imposed
161     between different regions or ``slabs'' of the simulation
162     box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang:2010uq} The
163     system responds by developing a temperature or momentum {\it gradient}
164     between the two regions. Since the amount of the applied flux is known
165     exactly, and the measurement of a gradient is generally less
166     complicated, imposed-flux methods typically take shorter simulation
167     times to obtain converged results for transport properties. The
168     corresponding temperature or velocity gradients which develop in
169     response to the applied flux are then related (via linear response
170     theory) to the transport coefficient of interest. These methods are
171     quite efficient, in that they do not need many trajectories to provide
172     information about transport properties. To date, they have been
173     utilized in computing thermal and mechanical transfer of both
174     homogeneous
175     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
176     well as heterogeneous
177     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
178    
179 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
180     % VSS-RNEMD
181     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
182     \subsection{VSS-RNEMD}
183 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
184     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
185     as a sequence of imaginary elastic collisions between particles in
186     regions separated by half of the simulation cell. In each collision,
187     the entire momentum vectors of both particles may be exchanged to
188     generate a thermal flux. Alternatively, a single component of the
189     momentum vectors may be exchanged to generate a shear flux. This
190     algorithm turns out to be quite useful in many simulations of bulk
191     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
192     system away from ideal Maxwell-Boltzmann distributions, and this has
193     undesirable side-effects when the applied flux becomes
194     large.\cite{Maginn:2010}
195    
196     Instead of having momentum exchange between {\it individual particles}
197     in each slab, the NIVS algorithm applies velocity scaling to all of
198     the selected particles in both slabs.\cite{Kuang:2010uq} A combination
199     of linear momentum, kinetic energy, and flux constraint equations
200     governs the amount of velocity scaling performed at each step. NIVS
201     has been shown to be very effective at producing thermal gradients and
202     for computing thermal conductivities, particularly for heterogeneous
203     interfaces. Although the NIVS algorithm can also be applied to impose
204     a directional momentum flux, thermal anisotropy was observed in
205     relatively high flux simulations, and the method is not suitable for
206     imposing a shear flux or for computing shear viscosities.
207    
208     The most useful RNEMD
209     approach developed so far utilizes a series of simultaneous velocity
210     shearing and scaling exchanges between the two
211     slabs.\cite{2012MolPh.110..691K} This method provides a set of
212     conservation constraints while simultaneously creating a desired flux
213     between the two slabs. Satisfying the constraint equations ensures
214     that the new configurations are sampled from the same NVE ensemble.
215    
216     The VSS moves are applied periodically to scale and shift the particle
217     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
218     $C$) which are separated by half of the simulation box,
219     \begin{displaymath}
220     \begin{array}{rclcl}
221    
222     & \underline{\mathrm{shearing}} & &
223     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
224     \mathbf{v}_i \leftarrow &
225     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
226     \rangle\right) + \langle\mathbf{v}_c\rangle \\
227     \mathbf{v}_j \leftarrow &
228     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
229     \rangle\right) + \langle\mathbf{v}_h\rangle .
230    
231     \end{array}
232     \end{displaymath}
233     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
234     the center of mass velocities in the $C$ and $H$ slabs, respectively.
235     Within the two slabs, particles receive incremental changes or a
236     ``shear'' to their velocities. The amount of shear is governed by the
237     imposed momentum flux, $j_z(\mathbf{p})$
238     \begin{eqnarray}
239     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
240     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
241     \end{eqnarray}
242     where $M_{\{c,h\}}$ is total mass of particles within each slab and $\Delta t$
243     is the interval between two separate operations.
244    
245     To simultaneously impose a thermal flux ($J_z$) between the slabs we
246     use energy conservation constraints,
247     \begin{eqnarray}
248     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
249     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
250     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
251     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
252     \mathbf{a}_h)^2 \label{vss4}.
253     \label{constraint}
254     \end{eqnarray}
255     Simultaneous solution of these quadratic formulae for the scaling
256     coefficients, $c$ and $h$, will ensure that the simulation samples from
257     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
258     instantaneous translational kinetic energy of each slab. At each time
259     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
260     and $\mathbf{a}_h$, subject to the imposed momentum flux,
261     $j_z(\mathbf{p})$, and thermal flux, $J_z$ values. Since the VSS
262     operations do not change the kinetic energy due to orientational
263     degrees of freedom or the potential energy of a system, configurations
264     after the VSS move have exactly the same energy ({\it and} linear
265     momentum) as before the move.
266    
267     As the simulation progresses, the VSS moves are performed on a regular
268     basis, and the system develops a thermal or velocity gradient in
269     response to the applied flux. Using the slope of the temperature or
270     velocity gradient, it is quite simple to obtain of thermal
271     conductivity ($\lambda$),
272     \begin{equation}
273     J_z = -\lambda \frac{\partial T}{\partial z},
274     \end{equation}
275     and shear viscosity ($\eta$),
276     \begin{equation}
277     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
278     \end{equation}
279     Here, the quantities on the left hand side are the imposed flux
280     values, while the slopes are obtained from linear fits to the
281     gradients that develop in the simulated system.
282    
283     The VSS-RNEMD approach is versatile in that it may be used to
284     implement both thermal and shear transport either separately or
285     simultaneously. Perturbations of velocities away from the ideal
286     Maxwell-Boltzmann distributions are minimal, and thermal anisotropy is
287     kept to a minimum. This ability to generate simultaneous thermal and
288     shear fluxes has been previously utilized to map out the shear
289     viscosity of SPC/E water over a wide range of temperatures (90~K) with
290     a {\it single 1 ns simulation}.\cite{2012MolPh.110..691K}
291    
292 kstocke1 3801 \begin{figure}
293     \includegraphics[width=\linewidth]{figures/rnemd}
294     \caption{VSS-RNEMD}
295     \label{fig:rnemd}
296     \end{figure}
297    
298    
299    
300     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
301     % INTERFACIAL CONDUCTANCE
302     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303 gezelter 3803 \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
304     to interfacial transport}
305 kstocke1 3801
306 gezelter 3803 Interfaces between dissimilar materials have transport properties
307     which can be defined as derivatives of the standard transport
308     coefficients in a direction normal to the interface. For example, the
309     {\it interfacial} thermal conductance ($G$) can be thought of as the
310     change in the thermal conductivity ($\lambda$) across the boundary
311     between materials:
312     \begin{align}
313     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
314     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
315     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
316     \label{derivativeG}
317     \end{align}
318     where $z_0$ is the location of the interface between two materials and
319     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
320     to be the $z$ direction from here on). RNEMD simulations, and
321     particularly the VSS-RNEMD approach, function by applying a momentum
322     or thermal flux and watching the gradient response of the
323     material. This means that the {\it interfacial} conductance is a
324     second derivative property which is subject to significant noise and
325     therefore requires extensive sampling. We have been able to
326     demonstrate the use of the second derivative approach to compute
327     interfacial conductance at chemically-modified metal / solvent
328     interfaces. However, a definition of the interfacial conductance in
329     terms of a temperature difference ($\Delta T$) measured at $z_0$,
330     \begin{displaymath}
331     G = \frac{J_z}{\Delta T_{z_0}},
332     \end{displaymath}
333     is useful once the RNEMD approach has generated a stable temperature
334     gap across the interface.
335    
336     \begin{figure}
337     \includegraphics[width=\linewidth]{figures/resistor_series}
338     \caption{RESISTOR SERIES}
339     \label{fig:resistor_series}
340     \end{figure}
341    
342     In the particular case we are studying here, there are two interfaces
343     involved in the transfer of heat from the gold slab to the solvent:
344     the gold/thiolate interface and the thiolate/solvent interface. We
345     could treat the temperature on each side of an interface as discrete,
346     making the interfacial conductance the inverse of the Kaptiza
347     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
348     across multiple interfaces, it is useful to think of the interfaces as
349     a set of resistors wired in series. The total resistance is then
350     additive, $R_{total} = \sum_i R_{i}$ and the interfacial conductance
351     is the inverse of the total resistance, or $G = \frac{1}{\sum_i
352     R_i}$). In the interfacial region, we treat each bin in the
353     VSS-RNEMD temperature profile as a resistor with resistance
354     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
355     the set of resistors which spans the gold/thiolate interface, thiolate
356     chains, and thiolate/solvent interface simplifies to
357 kstocke1 3801 \begin{equation}
358 gezelter 3803 \frac{T_{n}-T_{1}}{J_z},
359     \label{eq:finalG}
360     \end{equation}
361     or the temperature difference between the gold side of the
362     gold/thiolate interface and the solvent side of the thiolate/solvent
363     interface over the applied flux.
364 kstocke1 3801
365    
366     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
367     % **COMPUTATIONAL DETAILS**
368     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
369     \section{Computational Details}
370    
371 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372 gezelter 3819 % FORCE-FIELD PARAMETERS
373     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374     \subsection{Force-Field Parameters}
375 kstocke1 3801
376 gezelter 3819 Our simulations include a number of chemically distinct components.
377     Figure \ref{fig:structures} demonstrates the sites defined for both
378     the {\it n}-hexane and alkanethiolate ligands present in our
379     simulations. Force field parameters are needed for interactions both
380     between the same type of particles and between particles of different
381     species.
382    
383     \begin{figure}
384     \includegraphics[width=\linewidth]{figures/structures}
385     \caption{STRUCTURES}
386     \label{fig:structures}
387     \end{figure}
388    
389     The Au-Au interactions in metal lattice slab is described by the
390     quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
391     potentials include zero-point quantum corrections and are
392     reparametrized for accurate surface energies compared to the
393     Sutton-Chen potentials.\cite{Chen90}
394    
395     For the {\it n}-hexane solvent molecules, the TraPPE-UA
396     parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
397     sites are located at the carbon centers for alkyl groups. Bonding
398     interactions, including bond stretches and bends and torsions, were
399     used for intra-molecular sites closer than 3 bonds. For non-bonded
400     interactions, Lennard-Jones potentials are used. We have previously
401     utilized both united atom (UA) and all-atom (AA) force fields for
402     thermal conductivity in previous work,\cite{} and since the united
403     atom force fields cannot populate the high-frequency modes that are
404     present in AA force fields, they appear to work better for modeling
405     thermal conductivity. The TraPPE-UA model for alkanes is known to
406     predict a slightly lower boiling point than experimental values. This
407     is one of the reasons we used a lower average temperature (200K) for
408     our simulations.
409    
410     The TraPPE-UA force field includes parameters for thiol
411     molecules\cite{TraPPE-UA.thiols} which were used for the
412     alkanethiolate molecules in our simulations. To derive suitable
413     parameters for butanethiol adsorbed on Au(111) surfaces, we adopted
414     the S parameters from Luedtke and Landman\cite{landman:1998} and
415     modified the parameters for the CTS atom to maintain charge neutrality
416     in the molecule.
417    
418     To describe the interactions between metal (Au) and non-metal atoms,
419     we refer to an adsorption study of alkyl thiols on gold surfaces by
420     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
421     Lennard-Jones form of potential parameters for the interaction between
422     Au and pseudo-atoms CH$_x$ and S based on a well-established and
423     widely-used effective potential of Hautman and Klein for the Au(111)
424     surface.\cite{hautman:4994} As our simulations require the gold slab
425     to be flexible to accommodate thermal excitation, the pair-wise form
426     of potentials they developed was used for our study. Table
427     \ref{table:pars} in the supporting information summarizes the
428     ``metal/non-metal'' parameters utilized in our simulations.
429 kstocke1 3821
430     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
431     % SIMULATION PROTOCOL
432     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
433     \subsection{Simulation Protocol}
434    
435     We have implemented the VSS-RNEMD algorithm in OpenMD, our
436     group molecular dynamics code. A gold slab 11 atoms thick was
437     equilibrated at 1 atm and 200 K. The periodic box was expanded
438     in the z direction to expose two Au(111) faces.
439    
440     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.
441    
442     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.
443    
444     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.
445    
446     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.
447    
448     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.
449    
450    
451 kstocke1 3801
452     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
453     % **RESULTS**
454     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
455     \section{Results}
456    
457    
458 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
459     % CHAIN LENGTH
460     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
461     \subsection{Effect of Chain Length}
462    
463 kstocke1 3815 We examined full coverages of five chain lengths, n = 4, 6, 8, 10, 12. In all cases, the hexane solvent was unable to penetrate into the thiolate layer, leading to a persistent 2-4 \AA \, gap between the solvent region and the thiolates. The trend of interfacial conductance is mostly flat as a function of chain length, indicating that the length of the thiolate alkyl chains does not play a significant role in the transport of heat across the gold/thiolate and thiolate/solvent interfaces. There is, however, a peak in conductance for a chain length of 6 (hexanethiolate). This may be due to the equivalent chain lengths of the hexane solvent and the alkyl chain of the capping agent, leading to an especially high degree of vibrational overlap between the thiolate and solvent. Strong vibrational overlap would allow for efficient thermal energy transfer across the thiolate/solvent interface.
464 kstocke1 3801
465 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
466     % MIXED CHAINS
467     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468     \subsection{Effect of Mixed Chain Lengths}
469    
470 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.
471 kstocke1 3801
472 kstocke1 3815 \subsubsection{Butanethiolate/Decanethiolate}
473     Mixtures of butanethiolate/decanethiolate (n = 4, 10) have a peak interfacial condutance for 25\%/75\% short/long chains.
474 kstocke1 3801
475 kstocke1 3815 \subsubsection{Butanethiolate/Dodecanethiolate}
476 kstocke1 3821 Mixtures of butanethiolate/dodecanethiolate (n = 4, 12) have a peak interfacial conductance for 12.5\%/87.5\% short/long chains.
477 kstocke1 3815
478    
479 kstocke1 3801
480    
481     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
482     % **DISCUSSION**
483     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
484     \section{Discussion}
485    
486 gezelter 3819 In the mixed chain-length simulations, solvent molecules can become
487     temporarily trapped or entangled with the thiolate chains. Their
488     residence in close proximity to the higher temperature environment
489     close to the surface allows them to carry heat away from the surface
490     quite efficiently. There are two aspects of this behavior that are
491     relevant to thermal conductance of the interface: the residence time
492     of solvent molecules in the thiolate layer, and the orientational
493     ordering of the C-C chains as a mechanism for transferring vibrational
494     energy to these entrapped solvent molecules.
495    
496     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
497     % RESIDENCE TIME
498     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499 kstocke1 3821 \subsection{Mobility for solvent in the interfacial layer}
500 gezelter 3819
501     We use a simple survival correlation function, $C(t)$, to quantify the
502     residence time of a solvent molecule in the long thiolate chain
503     layer. This function correlates the identity of all hexane molecules
504     within the $z$-coordinate range of the thiolate layer at two separate
505     times. If the solvent molecule is present at both times, the
506     configuration contributes a $1$, while the absence of the molecule at
507     the later time indicates that the solvent molecule has migrated into
508     the bulk, and this configuration contributes a $0$. A steep decay in
509     $C(t)$ indicates a high turnover rate of solvent molecules from the
510 kstocke1 3821 chain region to the bulk. % The correlation function is easily fit
511     % using a biexponential,
512     % \begin{equation}
513     % C(t) = A \, e^{-t/\tau_{short}} + (1-A) e^{-t/\tau_{long}}
514     % \label{eq:biexponential}
515     % \end{equation}
516     % to determine short and long residence timescales and the relative populations of solvent molecules that can escape rapidly.
517     We define the mobility of solvent molecules at the interface as
518 kstocke1 3815 \begin{equation}
519 kstocke1 3821 M = \int_{0}^{T} 1 - C(t) dt,
520     \label{eq:mobility}
521 kstocke1 3815 \end{equation}
522 kstocke1 3821 where T is the length of the simulation.
523     In figure \ref{figure:res} we show that interfacial solvent mobility decreases as the percentage of long thiolate chains increases.
524 gezelter 3819
525    
526     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
527     % ORDER PARAMETER
528     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529    
530     \subsection{Vibrational coupling via orientational ordering}
531    
532     As the fraction of long-chain thiolates becomes large, the entrapped
533     solvent molecules must find specific orientations relative to the mean
534     orientation of the thiolate chains. This configuration allows for
535     efficient thermal energy exchange between the thiolate alkyl chain and
536     the solvent molecules.
537    
538 kstocke1 3821 To quantify this cooperative
539 gezelter 3819 ordering, we computed the orientational order parameters and director
540     axes for both the thiolate chains and for the entrapped solvent. The
541     director axis can be easily obtained by diagonalization of the order
542     parameter tensor,
543     \begin{equation}
544     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
545     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
546 kstocke1 3815 \end{equation}
547 gezelter 3819 where $\mathbf{e}_{i \alpha}$ was the $\alpha = x,y,z$ component of
548     the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
549     For both kinds of molecules, the $\mathbf{e}$ vector is defined using
550     the terminal atoms of the chains.
551    
552     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
553     traditionally used to obtain orientational order parameter, while the
554     eigenvector corresponding to the order parameter yields the director
555     axis ($\mathbf{d}(t)$) which defines the average direction of
556     molecular alignment at any time $t$. The overlap between the director
557     axes of the thiolates and the entrapped solvent is time-averaged,
558 kstocke1 3815 \begin{equation}
559 gezelter 3819 \left \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
560     \mathbf{d}_{solvent} \left( t \right) \right \rangle,
561 kstocke1 3815 \label{eq:orientation}
562     \end{equation}
563 gezelter 3819 and reported in table \ref{table:ordering}.
564 kstocke1 3815
565 gezelter 3819 Once the solvent molecules have picked up thermal energy from the
566     thiolates, they carry heat away from the gold as they diffuse back
567     into the bulk solvent. When the percentage of long chains decreases,
568     the tails of the long chains are much more disordered and do not
569     provide structured channels for the solvent to fill.
570 kstocke1 3815
571 gezelter 3819 Although the alignment of the chains with the entrapped solvent is one
572     possible mechanism for the non-monotonic increase in the conductance
573     as a function
574    
575    
576 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
577     % **ACKNOWLEDGMENTS**
578     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
579     \section*{Acknowledgments}
580     Support for this project was provided by the
581     National Science Foundation under grant CHE-0848243. Computational
582     time was provided by the Center for Research Computing (CRC) at the
583     University of Notre Dame.
584    
585     \newpage
586    
587     \bibliography{thiolsRNEMD}
588    
589     \end{doublespace}
590     \end{document}
591    
592    
593    
594    
595    
596    
597    
598    

Properties

Name Value
svn:executable *