ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chainLength/GoldThiolsPaper.tex
Revision: 3855
Committed: Tue Dec 25 17:58:26 2012 UTC (11 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 40233 byte(s)
Log Message:
Added Hase papers, fixed a few grammatical issues

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 gezelter 3855 We report on simulations of heat conduction through Au(111) / hexane
56     interfaces in which the surface has been protected by a mixture of
57     short and long chain alkanethiolate ligands. Reverse
58     non-equilibrium molecular dynamics (RNEMD) was used to create a
59     thermal flux between the metal and solvent, and thermal conductance
60     was computed using the resulting thermal profiles near the
61     interface. We find a non-monotonic dependence of the interfacial
62     thermal conductance on the fraction of long chains present at the
63     interface, and correlate this behavior to both solvent ordering and
64     the rate of solvent escape from the thiolate layer immediately in
65     contact with the metal surface. Our results suggest that a mixed
66     vibrational transfer / convection model is necessary to explain the
67     features of heat transfer at this interface. The alignment of the
68     solvent chains with the ordered ligand allows rapid transfer of
69     energy to the trapped solvent and is the dominant feature for
70     ordered ligand layers. Diffusion of the vibrationally excited
71     solvent into the bulk also plays a significant role when the ligands
72     are less tightly packed.
73 kstocke1 3851
74 kstocke1 3801 \end{abstract}
75    
76     \newpage
77    
78     %\narrowtext
79    
80     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
81     % **INTRODUCTION**
82     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83     \section{Introduction}
84    
85 gezelter 3848 The structural details of the interfaces of metal nanoparticles
86 gezelter 3855 determine how energy flows between these particles and their
87 gezelter 3848 surroundings. Understanding this energy flow is essential in designing
88 kstocke1 3854 and functionalizing metallic nanoparticles for use in plasmonic photothermal
89 gezelter 3848 therapies,\cite{Jain:2007ux,Petrova:2007ad,Gnyawali:2008lp,Mazzaglia:2008to,Huff:2007ye,Larson:2007hw}
90     which rely on the ability of metallic nanoparticles to absorb light in
91     the near-IR, a portion of the spectrum in which living tissue is very
92     nearly transparent. The relevant physical property controlling the
93     transfer of this energy as heat into the surrounding tissue is the
94     interfacial thermal conductance, $G$, which can be somewhat difficult
95     to determine experimentally.\cite{Wilson:2002uq,Plech:2005kx}
96 gezelter 3803
97 gezelter 3855 Metallic particles have also been proposed for use in efficient
98 gezelter 3848 thermal-transfer fluids, although careful experiments by Eapen {\it et
99 gezelter 3850 al.} have shown that metal-particle-based nanofluids have thermal
100     conductivities that match Maxwell predictions.\cite{Eapen:2007th} The
101     likely cause of previously reported non-Maxwell
102 gezelter 3803 behavior\cite{Eastman:2001wb,Keblinski:2002bx,Lee:1999ct,Xue:2003ya,Xue:2004oa}
103     is percolation networks of nanoparticles exchanging energy via the
104 gezelter 3850 solvent,\cite{Eapen:2007mw} so it is important to get a detailed
105     molecular picture of particle-ligand and ligand-solvent interactions
106     in order to understand the thermal behavior of complex fluids. To
107     date, there have been few reported values (either from theory or
108 kstocke1 3854 experiment) of $G$ for ligand-protected nanoparticles embedded in
109 gezelter 3850 liquids, and there is a significant gap in knowledge about how
110     chemically distinct ligands or protecting groups will affect heat
111     transport from the particles.
112 gezelter 3803
113 gezelter 3848 Experimentally, the thermal properties of various nanostructured
114     interfaces have been investigated by a number of groups. Cahill and
115 gezelter 3850 coworkers studied thermal transport from metal nanoparticle/fluid
116     interfaces, epitaxial TiN/single crystal oxide interfaces, and
117     hydrophilic and hydrophobic interfaces between water and solids with
118     different self-assembled
119 gezelter 3819 monolayers.\cite{cahill:793,Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
120 gezelter 3803 Wang {\it et al.} studied heat transport through long-chain
121     hydrocarbon monolayers on gold substrate at the individual molecular
122     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
123     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
124     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
125     et al.} studied the cooling dynamics, which is controlled by thermal
126     interface resistance of glass-embedded metal
127     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
128     normally considered barriers for heat transport, Alper {\it et al.}
129 gezelter 3850 have suggested that specific ligands (capping agents) could completely
130 gezelter 3803 eliminate this barrier
131     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
132    
133 gezelter 3855 Recently, Hase and coworkers employed Non-Equilibrium Molecular
134     Dynamics (NEMD) simulations to study thermal transport from hot
135     Au(111) substrate to a self-assembled monolayer of alkylthiol with
136     relatively long chain (8-20 carbon atoms).\cite{hase:2010,hase:2011}
137     These simulations explained many of the features of the experiments of
138     Wang {\it et al.} However, ensemble averaged measurements for heat
139     conductance of interfaces between the capping monolayer on Au and a
140     solvent phase have yet to be studied with their approach. In previous
141     simulations, our group applied a variant of reverse non-equilibrium
142     molecular dynamics (RNEMD) to calculate the interfacial thermal
143     conductance at a metal / organic solvent interface that had been
144     chemically protected by butanethiolate groups.\cite{kuang:AuThl} Our
145 gezelter 3850 calculations suggested an explanation for the very large thermal
146 gezelter 3803 conductivity at alkanethiol-capped metal surfaces when compared with
147     bare metal/solvent interfaces. Specifically, the chemical bond
148     between the metal and the ligand introduces a vibrational overlap that
149     is not present without the protecting group, and the overlap between
150     the vibrational spectra (metal to ligand, ligand to solvent) provides
151     a mechanism for rapid thermal transport across the interface.
152    
153 gezelter 3850 A notable result of the previous simulations was the non-monotonic
154     dependence of $G$ on the fractional coverage of the metal surface by
155     the chemical protecting group. Gaps in surface coverage allow the
156 gezelter 3855 solvent molecules to come into direct contact with ligands that have
157     been heated by the metal surface, absorb thermal energy from the
158     ligands, and then diffuse away. Quantifying the role of overall
159     surface coverage was difficult because the ligands have lateral
160     mobility on the surface and can aggregate to form domains on the
161     timescale of the simulation.
162 gezelter 3803
163 gezelter 3855 To isolate the effect of ligand/solvent coupling while avoiding
164     lateral mobility of the surface ligands, the current work utilizes
165     monolayers of mixed chain-lengths in which the length mismatch between
166     long and short chains is sufficient to accommodate solvent
167     molecules. These completely covered (but mixed-chain) surfaces are
168     significantly less prone to surface domain formation, and allow us to
169     further investigate the mechanism of heat transport to the solvent. A
170     thermal flux is created using velocity shearing and scaling reverse
171     non-equilibrium molecular dynamics (VSS-RNEMD), and the resulting
172     temperature profiles are analyzed to yield information about the
173     interfacial thermal conductance.
174 gezelter 3803
175 gezelter 3848
176 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
177     % **METHODOLOGY**
178     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
179     \section{Methodology}
180    
181 gezelter 3803 There are many ways to compute bulk transport properties from
182     classical molecular dynamics simulations. Equilibrium Molecular
183 gezelter 3855 Dynamics (EMD) simulations can be used to compute the relevant time
184     correlation functions and transport coefficients can be calculated
185     assuming that linear response theory
186 gezelter 3803 holds.\cite{PhysRevB.37.5677,MASSOBRIO:1984bl,PhysRev.119.1,Viscardy:2007rp,che:6888,kinaci:014106}
187     For some transport properties (notably thermal conductivity), EMD
188     approaches are subject to noise and poor convergence of the relevant
189     correlation functions. Traditional Non-equilibrium Molecular Dynamics
190     (NEMD) methods impose a gradient (e.g. thermal or momentum) on a
191     simulation.\cite{ASHURST:1975tg,Evans:1982zk,ERPENBECK:1984sp,MAGINN:1993hc,Berthier:2002ij,Evans:2002ai,Schelling:2002dp,PhysRevA.34.1449,JiangHao_jp802942v}
192     However, the resulting flux is often difficult to
193     measure. Furthermore, problems arise for NEMD simulations of
194     heterogeneous systems, such as phase-phase boundaries or interfaces,
195     where the type of gradient to enforce at the boundary between
196     materials is unclear.
197    
198     {\it Reverse} Non-Equilibrium Molecular Dynamics (RNEMD) methods adopt
199     a different approach in that an unphysical {\it flux} is imposed
200     between different regions or ``slabs'' of the simulation
201 gezelter 3850 box.\cite{MullerPlathe:1997xw,ISI:000080382700030,Kuang2010} The
202 gezelter 3803 system responds by developing a temperature or momentum {\it gradient}
203     between the two regions. Since the amount of the applied flux is known
204     exactly, and the measurement of a gradient is generally less
205     complicated, imposed-flux methods typically take shorter simulation
206     times to obtain converged results for transport properties. The
207     corresponding temperature or velocity gradients which develop in
208     response to the applied flux are then related (via linear response
209     theory) to the transport coefficient of interest. These methods are
210     quite efficient, in that they do not need many trajectories to provide
211     information about transport properties. To date, they have been
212     utilized in computing thermal and mechanical transfer of both
213     homogeneous
214     liquids~\cite{MullerPlathe:1997xw,ISI:000080382700030,Maginn:2010} as
215     well as heterogeneous
216     systems.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
217    
218 gezelter 3822 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
219     % VSS-RNEMD
220     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
221     \subsection{VSS-RNEMD}
222 gezelter 3803 The original ``swapping'' approaches by M\"{u}ller-Plathe {\it et
223     al.}\cite{ISI:000080382700030,MullerPlathe:1997xw} can be understood
224     as a sequence of imaginary elastic collisions between particles in
225     regions separated by half of the simulation cell. In each collision,
226     the entire momentum vectors of both particles may be exchanged to
227     generate a thermal flux. Alternatively, a single component of the
228     momentum vectors may be exchanged to generate a shear flux. This
229     algorithm turns out to be quite useful in many simulations of bulk
230     liquids. However, the M\"{u}ller-Plathe swapping approach perturbs the
231 kstocke1 3854 system away from ideal Maxwell-Boltzmann distributions, which has
232 gezelter 3803 undesirable side-effects when the applied flux becomes
233     large.\cite{Maginn:2010}
234    
235 gezelter 3848 The most useful alternative RNEMD approach developed so far utilizes a
236 gezelter 3855 series of simultaneous velocity shearing and scaling (VSS) exchanges between
237 gezelter 3850 the two slabs.\cite{Kuang2012} This method provides a set of
238 gezelter 3803 conservation constraints while simultaneously creating a desired flux
239     between the two slabs. Satisfying the constraint equations ensures
240     that the new configurations are sampled from the same NVE ensemble.
241    
242     The VSS moves are applied periodically to scale and shift the particle
243     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
244     $C$) which are separated by half of the simulation box,
245     \begin{displaymath}
246     \begin{array}{rclcl}
247    
248     & \underline{\mathrm{shearing}} & &
249     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\ \\
250     \mathbf{v}_i \leftarrow &
251     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
252     \rangle\right) + \langle\mathbf{v}_c\rangle \\
253     \mathbf{v}_j \leftarrow &
254     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
255     \rangle\right) + \langle\mathbf{v}_h\rangle .
256    
257     \end{array}
258     \end{displaymath}
259     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
260     the center of mass velocities in the $C$ and $H$ slabs, respectively.
261     Within the two slabs, particles receive incremental changes or a
262     ``shear'' to their velocities. The amount of shear is governed by the
263 gezelter 3855 imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
264 gezelter 3803 \begin{eqnarray}
265     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
266     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
267     \end{eqnarray}
268 gezelter 3855 where $M_{\{c,h\}}$ is the total mass of particles within each of the
269     slabs and $\Delta t$ is the interval between two separate operations.
270 gezelter 3803
271     To simultaneously impose a thermal flux ($J_z$) between the slabs we
272     use energy conservation constraints,
273     \begin{eqnarray}
274     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
275     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
276     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
277     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
278     \mathbf{a}_h)^2 \label{vss4}.
279     \label{constraint}
280     \end{eqnarray}
281     Simultaneous solution of these quadratic formulae for the scaling
282     coefficients, $c$ and $h$, will ensure that the simulation samples from
283     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
284     instantaneous translational kinetic energy of each slab. At each time
285     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
286     and $\mathbf{a}_h$, subject to the imposed momentum flux,
287 kstocke1 3854 $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
288 gezelter 3803 operations do not change the kinetic energy due to orientational
289     degrees of freedom or the potential energy of a system, configurations
290 gezelter 3855 after the VSS move have exactly the same energy (and linear
291 gezelter 3803 momentum) as before the move.
292    
293     As the simulation progresses, the VSS moves are performed on a regular
294     basis, and the system develops a thermal or velocity gradient in
295     response to the applied flux. Using the slope of the temperature or
296 kstocke1 3854 velocity gradient, it is quite simple to obtain the thermal
297 gezelter 3803 conductivity ($\lambda$),
298     \begin{equation}
299     J_z = -\lambda \frac{\partial T}{\partial z},
300     \end{equation}
301     and shear viscosity ($\eta$),
302     \begin{equation}
303     j_z(p_x) = -\eta \frac{\partial \langle v_x \rangle}{\partial z}.
304     \end{equation}
305     Here, the quantities on the left hand side are the imposed flux
306     values, while the slopes are obtained from linear fits to the
307     gradients that develop in the simulated system.
308    
309     The VSS-RNEMD approach is versatile in that it may be used to
310     implement both thermal and shear transport either separately or
311     simultaneously. Perturbations of velocities away from the ideal
312 gezelter 3855 Maxwell-Boltzmann distributions are minimal, as is thermal anisotropy.
313     This ability to generate simultaneous thermal and shear fluxes has
314     been previously utilized to map out the shear viscosity of SPC/E water
315     over a wide range of temperatures (90~K) with a single 1 ns
316     simulation.\cite{Kuang2012}
317 gezelter 3803
318 gezelter 3822 \begin{figure}
319     \includegraphics[width=\linewidth]{figures/rnemd}
320     \caption{The VSS-RNEMD approach imposes unphysical transfer of
321     linear momentum or kinetic energy between a ``hot'' slab and a
322     ``cold'' slab in the simulation box. The system responds to this
323     imposed flux by generating velocity or temperature gradients. The
324     slope of the gradients can then be used to compute transport
325     properties (e.g. shear viscosity or thermal conductivity).}
326     \label{fig:rnemd}
327     \end{figure}
328    
329     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330     % INTERFACIAL CONDUCTANCE
331     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
332     \subsection{Reverse Non-Equilibrium Molecular Dynamics approaches
333 gezelter 3803 to interfacial transport}
334 kstocke1 3801
335 gezelter 3803 Interfaces between dissimilar materials have transport properties
336     which can be defined as derivatives of the standard transport
337     coefficients in a direction normal to the interface. For example, the
338     {\it interfacial} thermal conductance ($G$) can be thought of as the
339     change in the thermal conductivity ($\lambda$) across the boundary
340     between materials:
341     \begin{align}
342     G &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
343     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
344     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2.
345     \label{derivativeG}
346     \end{align}
347     where $z_0$ is the location of the interface between two materials and
348     $\mathbf{\hat{n}}$ is a unit vector normal to the interface (assumed
349     to be the $z$ direction from here on). RNEMD simulations, and
350     particularly the VSS-RNEMD approach, function by applying a momentum
351     or thermal flux and watching the gradient response of the
352     material. This means that the {\it interfacial} conductance is a
353     second derivative property which is subject to significant noise and
354 gezelter 3855 therefore requires extensive sampling. Previous work has demonstrated
355     the use of the second derivative approach to compute interfacial
356     conductance at chemically-modified metal / solvent interfaces.
357     However, a definition of the interfacial conductance in terms of a
358     temperature difference ($\Delta T$) measured at $z_0$,
359 gezelter 3803 \begin{displaymath}
360     G = \frac{J_z}{\Delta T_{z_0}},
361     \end{displaymath}
362     is useful once the RNEMD approach has generated a stable temperature
363     gap across the interface.
364    
365     \begin{figure}
366     \includegraphics[width=\linewidth]{figures/resistor_series}
367 gezelter 3822 \caption{The inverse of the interfacial thermal conductance, $G$, is
368 kstocke1 3854 the Kapitza resistance, $R_K$. Because the gold / thiolate /
369 gezelter 3822 solvent interface extends a significant distance from the metal
370     surface, the interfacial resistance $R_K$ can be computed by
371     summing a series of temperature drops between adjacent temperature
372     bins along the $z$ axis.}
373 gezelter 3803 \label{fig:resistor_series}
374     \end{figure}
375    
376     In the particular case we are studying here, there are two interfaces
377     involved in the transfer of heat from the gold slab to the solvent:
378 kstocke1 3854 the metal / thiolate interface and the thiolate / solvent interface. We
379     can treat the temperature on each side of an interface as discrete,
380 gezelter 3803 making the interfacial conductance the inverse of the Kaptiza
381     resistance, or $G = \frac{1}{R_k}$. To model the total conductance
382     across multiple interfaces, it is useful to think of the interfaces as
383     a set of resistors wired in series. The total resistance is then
384 kstocke1 3854 additive, $R_{total} = \sum_i R_{i}$, and the interfacial conductance
385 gezelter 3803 is the inverse of the total resistance, or $G = \frac{1}{\sum_i
386     R_i}$). In the interfacial region, we treat each bin in the
387     VSS-RNEMD temperature profile as a resistor with resistance
388     $\frac{T_{2}-T_{1}}{J_z}$, $\frac{T_{3}-T_{2}}{J_z}$, etc. The sum of
389 kstocke1 3854 the set of resistors which spans the gold / thiolate interface, thiolate
390     chains, and thiolate / solvent interface simplifies to
391 kstocke1 3801 \begin{equation}
392 gezelter 3825 R_{K} = \frac{T_{n}-T_{1}}{J_z},
393 gezelter 3803 \label{eq:finalG}
394     \end{equation}
395     or the temperature difference between the gold side of the
396 kstocke1 3854 gold / thiolate interface and the solvent side of the thiolate / solvent
397 gezelter 3803 interface over the applied flux.
398 kstocke1 3801
399    
400     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
401     % **COMPUTATIONAL DETAILS**
402     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403     \section{Computational Details}
404    
405 gezelter 3803 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
406 gezelter 3819 % FORCE-FIELD PARAMETERS
407     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
408     \subsection{Force-Field Parameters}
409 kstocke1 3801
410 gezelter 3819 Our simulations include a number of chemically distinct components.
411     Figure \ref{fig:structures} demonstrates the sites defined for both
412     the {\it n}-hexane and alkanethiolate ligands present in our
413 gezelter 3850 simulations.
414 gezelter 3819
415     \begin{figure}
416     \includegraphics[width=\linewidth]{figures/structures}
417 gezelter 3822 \caption{Topologies of the thiolate capping agents and solvent
418     utilized in the simulations. The chemically-distinct sites (S,
419     \ce{CH2}, and \ce{CH3}) are treated as united atoms. Most
420     parameters are taken from references \bibpunct{}{}{,}{n}{}{,}
421     \protect\cite{TraPPE-UA.alkanes} and
422     \protect\cite{TraPPE-UA.thiols}. Cross-interactions with the Au
423     atoms were adapted from references
424     \protect\cite{landman:1998},~\protect\cite{vlugt:cpc2007154},~and
425     \protect\cite{hautman:4994}.}
426 gezelter 3819 \label{fig:structures}
427     \end{figure}
428    
429 kstocke1 3854 The Au-Au interactions in the metal lattice slab were described by the
430 gezelter 3819 quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
431     potentials include zero-point quantum corrections and are
432     reparametrized for accurate surface energies compared to the
433     Sutton-Chen potentials.\cite{Chen90}
434    
435     For the {\it n}-hexane solvent molecules, the TraPPE-UA
436     parameters\cite{TraPPE-UA.alkanes} were utilized. In this model,
437     sites are located at the carbon centers for alkyl groups. Bonding
438     interactions, including bond stretches and bends and torsions, were
439     used for intra-molecular sites closer than 3 bonds. For non-bonded
440 gezelter 3850 interactions, Lennard-Jones potentials were used. We have previously
441 gezelter 3819 utilized both united atom (UA) and all-atom (AA) force fields for
442 gezelter 3850 thermal conductivity,\cite{kuang:AuThl,Kuang2012} and since the united
443 gezelter 3819 atom force fields cannot populate the high-frequency modes that are
444     present in AA force fields, they appear to work better for modeling
445     thermal conductivity. The TraPPE-UA model for alkanes is known to
446     predict a slightly lower boiling point than experimental values. This
447 kstocke1 3854 is one of the reasons we used a lower average temperature (220 K) for
448 gezelter 3850 our simulations.
449 gezelter 3819
450     The TraPPE-UA force field includes parameters for thiol
451     molecules\cite{TraPPE-UA.thiols} which were used for the
452     alkanethiolate molecules in our simulations. To derive suitable
453 kstocke1 3854 parameters for butanethiolate adsorbed on Au(111) surfaces, we adopted
454 gezelter 3819 the S parameters from Luedtke and Landman\cite{landman:1998} and
455     modified the parameters for the CTS atom to maintain charge neutrality
456     in the molecule.
457    
458     To describe the interactions between metal (Au) and non-metal atoms,
459     we refer to an adsorption study of alkyl thiols on gold surfaces by
460     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
461     Lennard-Jones form of potential parameters for the interaction between
462     Au and pseudo-atoms CH$_x$ and S based on a well-established and
463     widely-used effective potential of Hautman and Klein for the Au(111)
464     surface.\cite{hautman:4994} As our simulations require the gold slab
465     to be flexible to accommodate thermal excitation, the pair-wise form
466 kstocke1 3832 of potentials they developed was used for our study.
467 kstocke1 3821
468     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
469     % SIMULATION PROTOCOL
470     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
471     \subsection{Simulation Protocol}
472    
473 gezelter 3850 We have implemented the VSS-RNEMD algorithm in OpenMD, our group
474     molecular dynamics code.\cite{openmd} An 1188 atom gold slab was
475 gezelter 3855 prepared and equilibrated at 1 atm and 200 K. The periodic box was
476     then expanded in the $z$ direction to expose two Au(111) faces on
477     either side of the 11-layer slab.
478 kstocke1 3821
479 gezelter 3850 A full monolayer of thiolates (1/3 the number of surface gold atoms)
480     was placed on three-fold hollow sites on the gold interfaces. The
481     effect of thiolate binding sites on the thermal conductance was tested
482     by placing thiolates at both fcc and hcp hollow sites. No appreciable
483 kstocke1 3854 difference in the temperature profile due to the location of
484     thiolate binding was noted.
485 kstocke1 3821
486 gezelter 3850 To test the role of thiolate chain length on interfacial thermal
487     conductance, full coverages of each of five chain lengths were tested:
488     butanethiolate (C$_4$), hexanethiolate (C$_6$), octanethiolate
489     (C$_8$), decanethiolate (C$_{10}$), and dodecanethiolate
490     (C$_{12}$). To test the effect of mixed chain lengths, full coverages
491     of C$_4$ / C$_{10}$ and C$_4$ / C$_{12}$ mixtures were created in
492     short/long chain percentages of 25/75, 50/50, 62.5/37.5, 75/25, and
493     87.5/12.5. The short and long chains were placed on the surface hollow
494     sites in a random configuration.
495 kstocke1 3821
496 gezelter 3850 The simulation box $z$-dimension was set to roughly double the length
497 kstocke1 3854 of the gold / thiolate block. Hexane solvent molecules were placed in
498 gezelter 3855 the vacant portion of the box using the packmol
499     algorithm.\cite{packmol} Figure \ref{fig:timelapse} shows two of the
500     mixed chain length interfaces both before and after the RNEMD simulation.
501 kstocke1 3821
502 kstocke1 3851 \begin{figure}
503     \includegraphics[width=\linewidth]{figures/timelapse}
504 kstocke1 3854 \caption{Images of 25\%~C$_4$~/~75\%~C$_{12}$ (top panel) and 75\%~C$_4$~/~25\%~C$_{12}$ (bottom panel) interfaces at the beginning and end of 3 ns simulations. Solvent molecules that were initially present in the thiolate layer are colored light blue. Diffusion of the initially-trapped solvent into the bulk is apparent in the interface with fewer long chains. Trapped solvent is orientationally locked to the ordered ligands (and is less able to diffuse into the bulk) when the fraction of long chains increases.}
505 kstocke1 3851 \label{fig:timelapse}
506     \end{figure}
507    
508 gezelter 3850 The system was equilibrated to 220 K in the canonical (NVT) ensemble,
509     allowing the thiolates and solvent to warm gradually. Pressure
510     correction to 1 atm was done using an isobaric-isothermal (NPT)
511     integrator that allowed expansion or contraction only in the $z$
512     direction, maintaining the crystalline structure of the gold as close
513     to the bulk result as possible. The diagonal elements of the pressure
514     tensor were monitored during the pressure equilibration stage. If the
515     $xx$ and/or $yy$ elements had a mean above zero throughout the
516     simulation -- indicating residual surface tension in the plane of the
517     gold slab -- an additional short NPT equilibration step was performed
518     allowing all box dimensions to change. Once the pressure was stable
519     at 1 atm, a final equilibration stage was performed at constant
520     temperature. All systems were equilibrated in the microcanonical (NVE)
521     ensemble before proceeding with the VSS-RNEMD and data collection
522     stages.
523 kstocke1 3821
524 gezelter 3855 A kinetic energy flux was applied using VSS-RNEMD during a data
525     collection period of 3 nanoseconds, with velocity scaling moves
526     occurring every 10 femtoseconds. The ``hot'' slab was centered in the
527     gold and the ``cold'' slab was placed in the center of the solvent
528     region. The entire system had a (time-averaged) temperature of 220 K,
529     with a temperature difference between the hot and cold slabs of
530     approximately 30 K. The average temperature and kinetic energy flux
531     were selected to avoid solvent freezing (or glass formation) and to
532     prevent the thiolates from burying in the gold slab. The Au-S
533     interaction has a deep potential energy well, which allows sulfur
534     atoms to embed into the gold slab at temperatures above 250 K.
535     Simulation conditions were chosen to avoid both of these
536     situations.
537 kstocke1 3821
538 gezelter 3850 Temperature profiles of the system were created by dividing the box
539     into $\sim$ 3 \AA \, bins along the $z$ axis and recording the average
540     temperature of each bin.
541 kstocke1 3821
542 gezelter 3850
543 kstocke1 3801
544     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
545     % **RESULTS**
546     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
547     \section{Results}
548    
549 kstocke1 3851 The solvent, hexane, is a straight chain flexible alkane that is structurally
550     similar to the thiolate alkane tails. Previous work has shown that UA models
551     of hexane and butanethiolate have a high degree of vibrational
552     overlap.\cite{kuang:AuThl} This overlap provides a mechanism for thermal
553     energy conduction from the thiolates to the solvent. Indeed, we observe that
554     the interfacial conductance is twice as large with the thiolate monolayers (of
555     all chain lengths) than with the bare metal surface.
556 kstocke1 3801
557 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
558     % CHAIN LENGTH
559     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560     \subsection{Effect of Chain Length}
561    
562 kstocke1 3854 We examined full coverages of five alkyl chain lengths, C$_{4}$, C$_{6}$, C$_{8}$, C$_{10}$, and C$_{12}$. As shown in table \ref{table:chainlengthG}, the interfacial conductance is roughly constant as a function of chain length, indicating that the length of the thiolate alkyl chains does not play a significant role in the transport of heat across the gold/thiolate and thiolate/solvent interfaces. In all cases, the hexane solvent was unable to penetrate into the thiolate layer, leading to a persistent 2-4 \AA \, gap between the solvent region and the thiolates. However, while the identity of the alkyl thiolate capping agent has little effect on the interfacial thermal conductance, the presence of a full monolayer of capping agent provides a two-fold increase in the G value relative to a bare gold surface.
563 kstocke1 3830 \begin{longtable}{p{4cm} p{3cm}}
564 kstocke1 3851 \caption{Computed interfacial thermal conductance ($G$) values for bare gold and 100\% coverages of various thiolate alkyl chain lengths.}
565 kstocke1 3830 \\
566 kstocke1 3854 \centering {\bf Chain Length} & \centering\arraybackslash {\bf G (MW/m$^2$/K)} \\ \hline
567 kstocke1 3830 \endhead
568     \hline
569     \endfoot
570 kstocke1 3851 \centering bare metal & \centering\arraybackslash 30.2 \\
571 kstocke1 3854 $~~~~~~~~~~~~~~~~~~$ C$_{4}$ & \centering\arraybackslash 59.4 \\
572     $~~~~~~~~~~~~~~~~~~$ C$_{6}$ & \centering\arraybackslash 60.2 \\
573     $~~~~~~~~~~~~~~~~~~$ C$_{8}$ & \centering\arraybackslash 61.0 \\
574     $~~~~~~~~~~~~~~~~~~$ C$_{10}$ & \centering\arraybackslash 58.2 \\
575     $~~~~~~~~~~~~~~~~~~$ C$_{12}$ & \centering\arraybackslash 58.8
576 kstocke1 3830 \label{table:chainlengthG}
577     \end{longtable}
578 kstocke1 3801
579 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
580     % MIXED CHAINS
581     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
582     \subsection{Effect of Mixed Chain Lengths}
583    
584 kstocke1 3854 Previous simulations have demonstrated non-monotonic behavior for $G$ as a function of the surface coverage. One difficulty with the previous study was the ability of butanethiolate ligands to migrate on the Au(111) surface and to form segregated domains. To simulate the effect of low coverages while preventing thiolate domain formation, we maintain 100\% thiolate coverage while varying the proportions of short (butanethiolate, C$_4$) and long (decanethiolate, C$_{10}$, or dodecanethiolate, C$_{12}$) alkyl chains. Data on the conductance trend as the fraction of long chains was varied is shown in figure \ref{fig:Gstacks}. Note that as in the previous study, $G$ is dependent upon solvent accessibility to thermally excited ligands. Our simulations indicate a similar (but less dramatic) non-monotonic dependence on the fraction of long chains.
585 kstocke1 3829 \begin{figure}
586     \includegraphics[width=\linewidth]{figures/Gstacks}
587 kstocke1 3851 \caption{Interfacial thermal conductivity of mixed-chains has a non-monotonic dependence on the fraction of long chains (lower panels). At low fractions of long chains, the solvent escape rate ($k_{escape}$) dominates the heat transfer process, while the solvent-thiolate orientational ordering dominates in systems with higher fractions of long chains (upper panels).}
588 kstocke1 3829 \label{fig:Gstacks}
589     \end{figure}
590 kstocke1 3815
591 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
592 kstocke1 3854 % **DISCUSSION**
593 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
594 kstocke1 3854 \section{Discussion}
595 gezelter 3819
596 kstocke1 3854 In the mixed chain-length simulations, solvent molecules
597     can become temporarily trapped or entangled with the thiolate chains. Their
598     residence in close proximity to the higher temperature environment close to
599     the surface allows them to carry heat away from the surface quite efficiently.
600     There are two aspects of this behavior that are relevant to thermal
601     conductance of the interface: the residence time of solvent molecules in the
602     thiolate layer, and the alignment of solvent molecules with the ligand alkyl
603     chains as a mechanism for transferring vibrational energy to these entrapped
604     solvent molecules. To quantify these competing effects, we have computed
605     solvent escape rates from the thiolate layer as well as a joint orientational
606     order parameter between the trapped solvent and the thiolate ligands.
607    
608 gezelter 3819 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
609     % RESIDENCE TIME
610     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
611 kstocke1 3841 \subsection{Mobility of solvent in the interfacial layer}
612 gezelter 3819
613 kstocke1 3851 We use a simple survival correlation function, $C(t)$, to measure the
614     residence time of a solvent molecule in the thiolate
615 gezelter 3819 layer. This function correlates the identity of all hexane molecules
616     within the $z$-coordinate range of the thiolate layer at two separate
617     times. If the solvent molecule is present at both times, the
618     configuration contributes a $1$, while the absence of the molecule at
619     the later time indicates that the solvent molecule has migrated into
620     the bulk, and this configuration contributes a $0$. A steep decay in
621     $C(t)$ indicates a high turnover rate of solvent molecules from the
622 kstocke1 3837 chain region to the bulk. We define the escape rate for trapped solvent molecules at the interface as
623 kstocke1 3815 \begin{equation}
624 gezelter 3825 k_{escape} = \left( \int_0^T C(t) dt \right)^{-1}
625 kstocke1 3821 \label{eq:mobility}
626 kstocke1 3815 \end{equation}
627 gezelter 3825 where T is the length of the simulation. This is a direct measure of
628 kstocke1 3851 the rate at which solvent molecules initially entangled in the thiolate layer
629 kstocke1 3843 can escape into the bulk. As $k_{escape} \rightarrow 0$, the
630 kstocke1 3854 solvent becomes permanently trapped in the thiolate layer. In
631 kstocke1 3841 figure \ref{fig:Gstacks} we show that interfacial solvent mobility
632 kstocke1 3854 decreases as the fraction of long thiolate chains increases.
633 gezelter 3819
634     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
635     % ORDER PARAMETER
636     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
637    
638     \subsection{Vibrational coupling via orientational ordering}
639    
640 kstocke1 3851 As the fraction of long-chain thiolates increases, the entrapped
641 gezelter 3819 solvent molecules must find specific orientations relative to the mean
642 kstocke1 3851 orientation of the thiolate chains. This alignment allows for
643 gezelter 3819 efficient thermal energy exchange between the thiolate alkyl chain and
644     the solvent molecules.
645    
646 kstocke1 3854 Once the interfacial solvent molecules have picked up thermal energy from the
647     thiolates, they carry heat away from the gold as they diffuse back
648     into the bulk solvent. When the percentage of long chains decreases,
649     the tails of the long chains are much more disordered and do not
650     provide structured channels for the solvent to fill.
651    
652     To measure this cooperative ordering, we compute the orientational order
653 kstocke1 3851 parameters and director axes for both the thiolate chains and for the
654     entrapped solvent. The director axis can be easily obtained by diagonalization
655     of the order parameter tensor,
656 gezelter 3819 \begin{equation}
657     \mathsf{Q}_{\alpha \beta} = \frac{1}{2 N} \sum_{i=1}^{N} \left( 3 \mathbf{e}_{i
658     \alpha} \mathbf{e}_{i \beta} - \delta_{\alpha \beta} \right)
659 kstocke1 3815 \end{equation}
660 kstocke1 3854 where $\mathbf{e}_{i \alpha}$ is the $\alpha = x,y,z$ component of
661 gezelter 3819 the unit vector $\mathbf{e}_{i}$ along the long axis of molecule $i$.
662 kstocke1 3854 For both the solvent and the ligand, the $\mathbf{e}$ vector is defined using
663     the terminal atoms of the molecule.
664 gezelter 3819
665     The largest eigenvalue of $\overleftrightarrow{\mathsf{Q}}$ is
666 kstocke1 3854 traditionally used to obtain the orientational order parameter, while the
667 gezelter 3819 eigenvector corresponding to the order parameter yields the director
668 kstocke1 3854 axis ($\mathbf{d}(t)$), which defines the average direction of
669 gezelter 3819 molecular alignment at any time $t$. The overlap between the director
670     axes of the thiolates and the entrapped solvent is time-averaged,
671 kstocke1 3815 \begin{equation}
672 kstocke1 3851 \langle d \rangle = \langle \mathbf{d}_{thiolates} \left( t \right) \cdot
673 gezelter 3825 \mathbf{d}_{solvent} \left( t \right) \rangle_t
674 kstocke1 3815 \label{eq:orientation}
675     \end{equation}
676 kstocke1 3854 and reported in figure \ref{fig:Gstacks}. Values of $\langle d \rangle$ range from $0$ (solvent molecules in the ligand layer are perpendicular to the thiolate chains) to $1$ (solvent and ligand chains are aligned parallel to each other).
677 kstocke1 3815
678 kstocke1 3854 C$_4$ / C$_{10}$ mixed monolayers have a peak interfacial conductance with 75\% long chains. At this fraction of long chains, the cooperative orientational ordering of the solvent molecules and chains becomes the dominant effect while the solvent escape rate is quite slow. C$_4$ / C$_{12}$ mixtures have a peak interfacial conductance for 87.5\% long chains. The solvent-thiolate orientational ordering reaches its maximum value at this long chain fraction. Long chain fractions of over $0.5$ for the C$_4$ / C$_{12}$ system are well ordered, but this effect is tempered by the exceptionally slow solvent escape rate ($\sim$ 1 molecule / 2 ns).
679 kstocke1 3851
680 kstocke1 3843 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
681     % **CONCLUSIONS**
682     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
683     \section{Conclusions}
684 gezelter 3855 Our results suggest that a mixed vibrational transfer / convection
685     model may be necessary to explain the features of heat transfer at
686     this interface. The alignment of the solvent chains with the ordered
687     ligand allows rapid transfer of energy to the trapped solvent and
688     becomes the dominant feature for ordered ligand layers. Diffusion of
689     the vibrationally excited solvent into the bulk also plays a
690     significant role when the ligands are less tightly packed.
691 kstocke1 3843
692 kstocke1 3851 In the language of earlier continuum approaches to interfacial
693     conductance,\cite{RevModPhys.61.605} the alignment of the chains is an
694     important factor in the transfer of phonons from the thiolate layer to the
695     trapped solvent. The aligned solvent and thiolate chains have nearly identical
696     acoustic impedances and the phonons can scatter directly into a solvent
697     molecule that has been forced into alignment. When the entrapped solvent has
698     more configurations available, the likelihood of an impedance mismatch is
699     higher, and the phonon scatters into the solvent with lower
700     efficiency. The fractional coverage of the long chains is therefore a simple
701     way of tuning the acoustic mismatch between the thiolate layer and the hexane
702     solvent.
703 kstocke1 3843
704 kstocke1 3851 Efficient heat transfer also can be accomplished via convective or diffusive motion of vibrationally excited solvent back into the bulk. Once the entrapped solvent becomes too tightly aligned with the ligands, however, the convective avenue of heat transfer is cut off.
705 kstocke1 3843
706 kstocke1 3854 Our simulations suggest a number of routes to make interfaces with high thermal conductance. If it is possible to create an interface which forces the solvent into alignment with a ligand that shares many of the solvent's vibrational modes, while simultaneously preserving the ability of the solvent to diffuse back into the bulk, we would expect a significant jump in the interfacial conductance. One possible way to accomplish this is to use polyene ligands with alternating unsaturated bonds. These are significantly more rigid than long alkanes, and could force solvent alignment (even at low relative coverages) while preserving mobility of solvent molecules within the ligand layer.
707 kstocke1 3851
708 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
709 kstocke1 3851 % **ACKNOWLEDGMENTS**
710 kstocke1 3801 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
711 kstocke1 3851 \section*{Acknowledgments}
712 kstocke1 3801
713 kstocke1 3851 We gratefully acknowledge conversations with Dr. Shenyu Kuang. Support for
714     this project was provided by the National Science Foundation under grant
715     CHE-0848243. Computational time was provided by the Center for Research
716     Computing (CRC) at the University of Notre Dame.
717    
718 kstocke1 3801 \newpage
719    
720     \bibliography{thiolsRNEMD}
721    
722     \end{doublespace}
723     \end{document}
724    
725    
726    
727    
728    
729    
730    
731    

Properties

Name Value
svn:executable *