ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3770
Committed: Fri Dec 2 20:14:03 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 52519 byte(s)
Log Message:
start from methodology right onto interfacial paper. add a bib file.

File Contents

# User Rev Content
1 gezelter 3769 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5     \usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{tabularx}
8     \usepackage{graphicx}
9     \usepackage{multirow}
10     %\usepackage{booktabs}
11     %\usepackage{bibentry}
12     %\usepackage{mathrsfs}
13     %\usepackage[ref]{overcite}
14     \usepackage[square, comma, sort&compress]{natbib}
15     \usepackage{url}
16     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19    
20     % double space list of tables and figures
21     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22     \setlength{\abovecaptionskip}{20 pt}
23     \setlength{\belowcaptionskip}{30 pt}
24    
25     %\renewcommand\citemid{\ } % no comma in optional reference note
26     \bibpunct{[}{]}{,}{n}{}{;}
27     \bibliographystyle{achemso}
28    
29     \begin{document}
30    
31 skuang 3770 \title{ENTER TITLE HERE}
32 gezelter 3769
33     \author{Shenyu Kuang and J. Daniel
34     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
35     Department of Chemistry and Biochemistry,\\
36     University of Notre Dame\\
37     Notre Dame, Indiana 46556}
38    
39     \date{\today}
40    
41     \maketitle
42    
43     \begin{doublespace}
44    
45     \begin{abstract}
46 skuang 3770 REPLACE ABSTRACT HERE
47 gezelter 3769 With the Non-Isotropic Velocity Scaling (NIVS) approach to Reverse
48     Non-Equilibrium Molecular Dynamics (RNEMD) it is possible to impose
49     an unphysical thermal flux between different regions of
50     inhomogeneous systems such as solid / liquid interfaces. We have
51     applied NIVS to compute the interfacial thermal conductance at a
52     metal / organic solvent interface that has been chemically capped by
53     butanethiol molecules. Our calculations suggest that coupling
54     between the metal and liquid phases is enhanced by the capping
55     agents, leading to a greatly enhanced conductivity at the interface.
56     Specifically, the chemical bond between the metal and the capping
57     agent introduces a vibrational overlap that is not present without
58     the capping agent, and the overlap between the vibrational spectra
59     (metal to cap, cap to solvent) provides a mechanism for rapid
60     thermal transport across the interface. Our calculations also
61     suggest that this is a non-monotonic function of the fractional
62     coverage of the surface, as moderate coverages allow diffusive heat
63     transport of solvent molecules that have been in close contact with
64     the capping agent.
65    
66     \end{abstract}
67    
68     \newpage
69    
70     %\narrowtext
71    
72     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73     % BODY OF TEXT
74     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
75    
76     \section{Introduction}
77 skuang 3770 [DO THIS LATER]
78    
79     [IMPORTANCE OF NANOSCALE TRANSPORT PROPERTIES STUDIES]
80    
81 gezelter 3769 Due to the importance of heat flow (and heat removal) in
82     nanotechnology, interfacial thermal conductance has been studied
83     extensively both experimentally and computationally.\cite{cahill:793}
84     Nanoscale materials have a significant fraction of their atoms at
85     interfaces, and the chemical details of these interfaces govern the
86     thermal transport properties. Furthermore, the interfaces are often
87     heterogeneous (e.g. solid - liquid), which provides a challenge to
88     computational methods which have been developed for homogeneous or
89     bulk systems.
90    
91     Experimentally, the thermal properties of a number of interfaces have
92     been investigated. Cahill and coworkers studied nanoscale thermal
93     transport from metal nanoparticle/fluid interfaces, to epitaxial
94     TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
95     interfaces between water and solids with different self-assembled
96     monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
97     Wang {\it et al.} studied heat transport through long-chain
98     hydrocarbon monolayers on gold substrate at individual molecular
99     level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
100     cetyltrimethylammonium bromide (CTAB) on the thermal transport between
101     gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
102     et al.} studied the cooling dynamics, which is controlled by thermal
103     interface resistance of glass-embedded metal
104     nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
105     normally considered barriers for heat transport, Alper {\it et al.}
106     suggested that specific ligands (capping agents) could completely
107     eliminate this barrier
108     ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
109    
110     The acoustic mismatch model for interfacial conductance utilizes the
111     acoustic impedance ($Z_a = \rho_a v^s_a$) on both sides of the
112     interface.\cite{swartz1989} Here, $\rho_a$ and $v^s_a$ are the density
113     and speed of sound in material $a$. The phonon transmission
114     probability at the $a-b$ interface is
115     \begin{equation}
116     t_{ab} = \frac{4 Z_a Z_b}{\left(Z_a + Z_b \right)^2},
117     \end{equation}
118     and the interfacial conductance can then be approximated as
119     \begin{equation}
120     G_{ab} \approx \frac{1}{4} C_D v_D t_{ab}
121     \end{equation}
122     where $C_D$ is the Debye heat capacity of the hot side, and $v_D$ is
123     the Debye phonon velocity ($1/v_D^3 = 1/3v_L^3 + 2/3 v_T^3$) where
124     $v_L$ and $v_T$ are the longitudinal and transverse speeds of sound,
125     respectively. For the Au/hexane and Au/toluene interfaces, the
126     acoustic mismatch model predicts room-temperature $G \approx 87 \mbox{
127     and } 129$ MW m$^{-2}$ K$^{-1}$, respectively. However, it is not
128     clear how to apply the acoustic mismatch model to a
129     chemically-modified surface, particularly when the acoustic properties
130     of a monolayer film may not be well characterized.
131    
132 skuang 3770 [PREVIOUS METHODS INCLUDING NIVS AND THEIR LIMITATIONS]
133     [DIFFICULTY TO GENERATE JZKE AND JZP SIMUTANEOUSLY]
134    
135 gezelter 3769 More precise computational models have also been used to study the
136     interfacial thermal transport in order to gain an understanding of
137     this phenomena at the molecular level. Recently, Hase and coworkers
138     employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
139     study thermal transport from hot Au(111) substrate to a self-assembled
140     monolayer of alkylthiol with relatively long chain (8-20 carbon
141     atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
142     measurements for heat conductance of interfaces between the capping
143     monolayer on Au and a solvent phase have yet to be studied with their
144     approach. The comparatively low thermal flux through interfaces is
145     difficult to measure with Equilibrium
146     MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
147     methods. Therefore, the Reverse NEMD (RNEMD)
148     methods\cite{MullerPlathe:1997xw,kuang:164101} would be advantageous
149     in that they {\it apply} the difficult to measure quantity (flux),
150     while {\it measuring} the easily-computed quantity (the thermal
151     gradient). This is particularly true for inhomogeneous interfaces
152     where it would not be clear how to apply a gradient {\it a priori}.
153     Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
154     this approach to various liquid interfaces and studied how thermal
155     conductance (or resistance) is dependent on chemical details of a
156     number of hydrophobic and hydrophilic aqueous interfaces. And
157     recently, Luo {\it et al.} studied the thermal conductance of
158     Au-SAM-Au junctions using the same approach, comparing to a constant
159     temperature difference method.\cite{Luo20101} While this latter
160     approach establishes more ideal Maxwell-Boltzmann distributions than
161     previous RNEMD methods, it does not guarantee momentum or kinetic
162     energy conservation.
163    
164     Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
165     algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
166     retains the desirable features of RNEMD (conservation of linear
167     momentum and total energy, compatibility with periodic boundary
168     conditions) while establishing true thermal distributions in each of
169     the two slabs. Furthermore, it allows effective thermal exchange
170     between particles of different identities, and thus makes the study of
171     interfacial conductance much simpler.
172    
173 skuang 3770 [WHAT IS COVERED IN THIS MANUSCRIPT]
174     [MAY PUT FIGURE 1 HERE]
175 gezelter 3769 The work presented here deals with the Au(111) surface covered to
176     varying degrees by butanethiol, a capping agent with short carbon
177     chain, and solvated with organic solvents of different molecular
178     properties. Different models were used for both the capping agent and
179     the solvent force field parameters. Using the NIVS algorithm, the
180     thermal transport across these interfaces was studied and the
181     underlying mechanism for the phenomena was investigated.
182    
183     \section{Methodology}
184 skuang 3770 Similar to the NIVS methodology,\cite{kuang:164101} we consider a
185     periodic system divided into a series of slabs along a certain axis
186     (e.g. $z$). The unphysical thermal and/or momentum flux is designated
187     from the center slab to one of the end slabs, and thus the center slab
188     would have a lower temperature than the end slab (unless the thermal
189     flux is negative). Therefore, the center slab is denoted as ``$c$''
190     while the end slab as ``$h$''.
191    
192     To impose these fluxes, we periodically apply separate operations to
193     velocities of particles {$i$} within the center slab and of particles
194     {$j$} within the end slab:
195     \begin{eqnarray}
196     \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
197     \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
198     \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
199     \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
200     \end{eqnarray}
201     where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes
202     the instantaneous bulk velocity of slabs $c$ and $h$ respectively
203     before an operation occurs. When a momentum flux $\vec{j}_z(\vec{p})$
204     presents, these bulk velocities would have a corresponding change
205     ($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's
206     second law:
207     \begin{eqnarray}
208     M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\
209     M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
210     \end{eqnarray}
211     where
212     \begin{eqnarray}
213     M_c & = & \sum_{i = 1}^{N_c} m_i \\
214     M_h & = & \sum_{j = 1}^{N_h} m_j
215     \end{eqnarray}
216     and $\Delta t$ is the interval between two operations.
217    
218     The above operations conserve the linear momentum of a periodic
219     system. To satisfy total energy conservation as well as to impose a
220     thermal flux $J_z$, one would have
221     %SUPPORT INFO MIGHT BE NECESSARY TO PUT EXTRA MATH IN
222     \begin{eqnarray}
223     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
224     \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
225     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
226     \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2
227     \end{eqnarray}
228     where $K_c$ and $K_h$ denotes translational kinetic energy of slabs
229     $c$ and $h$ respectively before an operation occurs. These
230     translational kinetic energy conservation equations are sufficient to
231     ensure total energy conservation, as the operations applied do not
232     change the potential energy of a system, given that the potential
233     energy does not depend on particle velocity.
234    
235     The above sets of equations are sufficient to determine the velocity
236     scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
237     $\vec{a}_h$. Note that two roots of $c$ and $h$ exist
238     respectively. However, to avoid dramatic perturbations to a system,
239     the positive roots (which are closer to 1) are chosen.
240    
241     By implementing these operations at a certain frequency, a steady
242     thermal and/or momentum flux can be applied and the corresponding
243     temperature and/or momentum gradients can be established.
244     [REFER TO NIVS PAPER]
245     [ADVANTAGES]
246    
247 gezelter 3769 Steady state MD simulations have an advantage in that not many
248     trajectories are needed to study the relationship between thermal flux
249     and thermal gradients. For systems with low interfacial conductance,
250     one must have a method capable of generating or measuring relatively
251     small fluxes, compared to those required for bulk conductivity. This
252     requirement makes the calculation even more difficult for
253     slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
254     NEMD methods impose a gradient (and measure a flux), but at interfaces
255     it is not clear what behavior should be imposed at the boundaries
256     between materials. Imposed-flux reverse non-equilibrium
257     methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
258     the thermal response becomes an easy-to-measure quantity. Although
259     M\"{u}ller-Plathe's original momentum swapping approach can be used
260     for exchanging energy between particles of different identity, the
261     kinetic energy transfer efficiency is affected by the mass difference
262     between the particles, which limits its application on heterogeneous
263     interfacial systems.
264    
265     The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
266     to non-equilibrium MD simulations is able to impose a wide range of
267     kinetic energy fluxes without obvious perturbation to the velocity
268     distributions of the simulated systems. Furthermore, this approach has
269     the advantage in heterogeneous interfaces in that kinetic energy flux
270     can be applied between regions of particles of arbitrary identity, and
271     the flux will not be restricted by difference in particle mass.
272    
273     The NIVS algorithm scales the velocity vectors in two separate regions
274     of a simulation system with respective diagonal scaling matrices. To
275     determine these scaling factors in the matrices, a set of equations
276     including linear momentum conservation and kinetic energy conservation
277     constraints and target energy flux satisfaction is solved. With the
278     scaling operation applied to the system in a set frequency, bulk
279     temperature gradients can be easily established, and these can be used
280     for computing thermal conductivities. The NIVS algorithm conserves
281     momenta and energy and does not depend on an external thermostat.
282    
283     \subsection{Defining Interfacial Thermal Conductivity ($G$)}
284    
285     For an interface with relatively low interfacial conductance, and a
286     thermal flux between two distinct bulk regions, the regions on either
287     side of the interface rapidly come to a state in which the two phases
288     have relatively homogeneous (but distinct) temperatures. The
289     interfacial thermal conductivity $G$ can therefore be approximated as:
290     \begin{equation}
291     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
292     \langle T_\mathrm{cold}\rangle \right)}
293     \label{lowG}
294     \end{equation}
295     where ${E_{total}}$ is the total imposed non-physical kinetic energy
296     transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
297     and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
298     temperature of the two separated phases. For an applied flux $J_z$
299     operating over a simulation time $t$ on a periodically-replicated slab
300     of dimensions $L_x \times L_y$, $E_{total} = 2 J_z t L_x L_y$.
301    
302     When the interfacial conductance is {\it not} small, there are two
303     ways to define $G$. One common way is to assume the temperature is
304     discrete on the two sides of the interface. $G$ can be calculated
305     using the applied thermal flux $J$ and the maximum temperature
306     difference measured along the thermal gradient max($\Delta T$), which
307     occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
308     known as the Kapitza conductance, which is the inverse of the Kapitza
309     resistance.
310     \begin{equation}
311     G=\frac{J}{\Delta T}
312     \label{discreteG}
313     \end{equation}
314    
315     \begin{figure}
316     \includegraphics[width=\linewidth]{method}
317     \caption{Interfacial conductance can be calculated by applying an
318     (unphysical) kinetic energy flux between two slabs, one located
319     within the metal and another on the edge of the periodic box. The
320     system responds by forming a thermal gradient. In bulk liquids,
321     this gradient typically has a single slope, but in interfacial
322     systems, there are distinct thermal conductivity domains. The
323     interfacial conductance, $G$ is found by measuring the temperature
324     gap at the Gibbs dividing surface, or by using second derivatives of
325     the thermal profile.}
326     \label{demoPic}
327     \end{figure}
328    
329     Another approach is to assume that the temperature is continuous and
330     differentiable throughout the space. Given that $\lambda$ is also
331     differentiable, $G$ can be defined as its gradient ($\nabla\lambda$)
332     projected along a vector normal to the interface ($\mathbf{\hat{n}}$)
333     and evaluated at the interface location ($z_0$). This quantity,
334     \begin{align}
335     G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
336     &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
337     \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
338     &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
339     \left(\frac{\partial T}{\partial z}\right)_{z_0}^2 \label{derivativeG}
340     \end{align}
341     has the same units as the common definition for $G$, and the maximum
342     of its magnitude denotes where thermal conductivity has the largest
343     change, i.e. the interface. In the geometry used in this study, the
344     vector normal to the interface points along the $z$ axis, as do
345     $\vec{J}$ and the thermal gradient. This yields the simplified
346     expressions in Eq. \ref{derivativeG}.
347    
348     With temperature profiles obtained from simulation, one is able to
349     approximate the first and second derivatives of $T$ with finite
350     difference methods and calculate $G^\prime$. In what follows, both
351     definitions have been used, and are compared in the results.
352    
353     To investigate the interfacial conductivity at metal / solvent
354     interfaces, we have modeled a metal slab with its (111) surfaces
355     perpendicular to the $z$-axis of our simulation cells. The metal slab
356     has been prepared both with and without capping agents on the exposed
357     surface, and has been solvated with simple organic solvents, as
358     illustrated in Figure \ref{gradT}.
359    
360     With the simulation cell described above, we are able to equilibrate
361     the system and impose an unphysical thermal flux between the liquid
362     and the metal phase using the NIVS algorithm. By periodically applying
363     the unphysical flux, we obtained a temperature profile and its spatial
364     derivatives. Figure \ref{gradT} shows how an applied thermal flux can
365     be used to obtain the 1st and 2nd derivatives of the temperature
366     profile.
367    
368     \begin{figure}
369     \includegraphics[width=\linewidth]{gradT}
370     \caption{A sample of Au (111) / butanethiol / hexane interfacial
371     system with the temperature profile after a kinetic energy flux has
372     been imposed. Note that the largest temperature jump in the thermal
373     profile (corresponding to the lowest interfacial conductance) is at
374     the interface between the butanethiol molecules (blue) and the
375     solvent (grey). First and second derivatives of the temperature
376     profile are obtained using a finite difference approximation (lower
377     panel).}
378     \label{gradT}
379     \end{figure}
380    
381     \section{Computational Details}
382     \subsection{Simulation Protocol}
383     The NIVS algorithm has been implemented in our MD simulation code,
384     OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
385     Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
386     under atmospheric pressure (1 atm) and 200K. After equilibration,
387     butanethiol capping agents were placed at three-fold hollow sites on
388     the Au(111) surfaces. These sites are either {\it fcc} or {\it
389     hcp} sites, although Hase {\it et al.} found that they are
390     equivalent in a heat transfer process,\cite{hase:2010} so we did not
391     distinguish between these sites in our study. The maximum butanethiol
392     capacity on Au surface is $1/3$ of the total number of surface Au
393     atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
394     structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
395     series of lower coverages was also prepared by eliminating
396     butanethiols from the higher coverage surface in a regular manner. The
397     lower coverages were prepared in order to study the relation between
398     coverage and interfacial conductance.
399    
400     The capping agent molecules were allowed to migrate during the
401     simulations. They distributed themselves uniformly and sampled a
402     number of three-fold sites throughout out study. Therefore, the
403     initial configuration does not noticeably affect the sampling of a
404     variety of configurations of the same coverage, and the final
405     conductance measurement would be an average effect of these
406     configurations explored in the simulations.
407    
408     After the modified Au-butanethiol surface systems were equilibrated in
409     the canonical (NVT) ensemble, organic solvent molecules were packed in
410     the previously empty part of the simulation cells.\cite{packmol} Two
411     solvents were investigated, one which has little vibrational overlap
412     with the alkanethiol and which has a planar shape (toluene), and one
413     which has similar vibrational frequencies to the capping agent and
414     chain-like shape ({\it n}-hexane).
415    
416     The simulation cells were not particularly extensive along the
417     $z$-axis, as a very long length scale for the thermal gradient may
418     cause excessively hot or cold temperatures in the middle of the
419     solvent region and lead to undesired phenomena such as solvent boiling
420     or freezing when a thermal flux is applied. Conversely, too few
421     solvent molecules would change the normal behavior of the liquid
422     phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
423     these extreme cases did not happen to our simulations. The spacing
424     between periodic images of the gold interfaces is $45 \sim 75$\AA in
425     our simulations.
426    
427     The initial configurations generated are further equilibrated with the
428     $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
429     change. This is to ensure that the equilibration of liquid phase does
430     not affect the metal's crystalline structure. Comparisons were made
431     with simulations that allowed changes of $L_x$ and $L_y$ during NPT
432     equilibration. No substantial changes in the box geometry were noticed
433     in these simulations. After ensuring the liquid phase reaches
434     equilibrium at atmospheric pressure (1 atm), further equilibration was
435     carried out under canonical (NVT) and microcanonical (NVE) ensembles.
436    
437     After the systems reach equilibrium, NIVS was used to impose an
438     unphysical thermal flux between the metal and the liquid phases. Most
439     of our simulations were done under an average temperature of
440     $\sim$200K. Therefore, thermal flux usually came from the metal to the
441     liquid so that the liquid has a higher temperature and would not
442     freeze due to lowered temperatures. After this induced temperature
443     gradient had stabilized, the temperature profile of the simulation cell
444     was recorded. To do this, the simulation cell is divided evenly into
445     $N$ slabs along the $z$-axis. The average temperatures of each slab
446     are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
447     the same, the derivatives of $T$ with respect to slab number $n$ can
448     be directly used for $G^\prime$ calculations: \begin{equation}
449     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
450     \Big/\left(\frac{\partial T}{\partial z}\right)^2
451     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
452     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
453     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
454     \Big/\left(\frac{\partial T}{\partial n}\right)^2
455     \label{derivativeG2}
456     \end{equation}
457     The absolute values in Eq. \ref{derivativeG2} appear because the
458     direction of the flux $\vec{J}$ is in an opposing direction on either
459     side of the metal slab.
460    
461     All of the above simulation procedures use a time step of 1 fs. Each
462     equilibration stage took a minimum of 100 ps, although in some cases,
463     longer equilibration stages were utilized.
464    
465     \subsection{Force Field Parameters}
466     Our simulations include a number of chemically distinct components.
467     Figure \ref{demoMol} demonstrates the sites defined for both
468     United-Atom and All-Atom models of the organic solvent and capping
469     agents in our simulations. Force field parameters are needed for
470     interactions both between the same type of particles and between
471     particles of different species.
472    
473     \begin{figure}
474     \includegraphics[width=\linewidth]{structures}
475     \caption{Structures of the capping agent and solvents utilized in
476     these simulations. The chemically-distinct sites (a-e) are expanded
477     in terms of constituent atoms for both United Atom (UA) and All Atom
478     (AA) force fields. Most parameters are from References
479     \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
480     (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
481     atoms are given in Table 1 in the supporting information.}
482     \label{demoMol}
483     \end{figure}
484    
485     The Au-Au interactions in metal lattice slab is described by the
486     quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
487     potentials include zero-point quantum corrections and are
488     reparametrized for accurate surface energies compared to the
489     Sutton-Chen potentials.\cite{Chen90}
490    
491     For the two solvent molecules, {\it n}-hexane and toluene, two
492     different atomistic models were utilized. Both solvents were modeled
493     using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
494     parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
495     for our UA solvent molecules. In these models, sites are located at
496     the carbon centers for alkyl groups. Bonding interactions, including
497     bond stretches and bends and torsions, were used for intra-molecular
498     sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
499     potentials are used.
500    
501     By eliminating explicit hydrogen atoms, the TraPPE-UA models are
502     simple and computationally efficient, while maintaining good accuracy.
503     However, the TraPPE-UA model for alkanes is known to predict a slightly
504     lower boiling point than experimental values. This is one of the
505     reasons we used a lower average temperature (200K) for our
506     simulations. If heat is transferred to the liquid phase during the
507     NIVS simulation, the liquid in the hot slab can actually be
508     substantially warmer than the mean temperature in the simulation. The
509     lower mean temperatures therefore prevent solvent boiling.
510    
511     For UA-toluene, the non-bonded potentials between intermolecular sites
512     have a similar Lennard-Jones formulation. The toluene molecules were
513     treated as a single rigid body, so there was no need for
514     intramolecular interactions (including bonds, bends, or torsions) in
515     this solvent model.
516    
517     Besides the TraPPE-UA models, AA models for both organic solvents are
518     included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
519     were used. For hexane, additional explicit hydrogen sites were
520     included. Besides bonding and non-bonded site-site interactions,
521     partial charges and the electrostatic interactions were added to each
522     CT and HC site. For toluene, a flexible model for the toluene molecule
523     was utilized which included bond, bend, torsion, and inversion
524     potentials to enforce ring planarity.
525    
526     The butanethiol capping agent in our simulations, were also modeled
527     with both UA and AA model. The TraPPE-UA force field includes
528     parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
529     UA butanethiol model in our simulations. The OPLS-AA also provides
530     parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
531     surfaces do not have the hydrogen atom bonded to sulfur. To derive
532     suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
533     adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
534     modify the parameters for the CTS atom to maintain charge neutrality
535     in the molecule. Note that the model choice (UA or AA) for the capping
536     agent can be different from the solvent. Regardless of model choice,
537     the force field parameters for interactions between capping agent and
538     solvent can be derived using Lorentz-Berthelot Mixing Rule:
539     \begin{eqnarray}
540     \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
541     \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
542     \end{eqnarray}
543    
544     To describe the interactions between metal (Au) and non-metal atoms,
545     we refer to an adsorption study of alkyl thiols on gold surfaces by
546     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
547     Lennard-Jones form of potential parameters for the interaction between
548     Au and pseudo-atoms CH$_x$ and S based on a well-established and
549     widely-used effective potential of Hautman and Klein for the Au(111)
550     surface.\cite{hautman:4994} As our simulations require the gold slab
551     to be flexible to accommodate thermal excitation, the pair-wise form
552     of potentials they developed was used for our study.
553    
554     The potentials developed from {\it ab initio} calculations by Leng
555     {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
556     interactions between Au and aromatic C/H atoms in toluene. However,
557     the Lennard-Jones parameters between Au and other types of particles,
558     (e.g. AA alkanes) have not yet been established. For these
559     interactions, the Lorentz-Berthelot mixing rule can be used to derive
560     effective single-atom LJ parameters for the metal using the fit values
561     for toluene. These are then used to construct reasonable mixing
562     parameters for the interactions between the gold and other atoms.
563     Table 1 in the supporting information summarizes the
564     ``metal/non-metal'' parameters utilized in our simulations.
565    
566     \section{Results}
567 skuang 3770 [L-J COMPARED TO RENMD NIVS; WATER COMPARED TO RNEMD NIVS;
568     SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
569    
570 gezelter 3769 There are many factors contributing to the measured interfacial
571     conductance; some of these factors are physically motivated
572     (e.g. coverage of the surface by the capping agent coverage and
573     solvent identity), while some are governed by parameters of the
574     methodology (e.g. applied flux and the formulas used to obtain the
575     conductance). In this section we discuss the major physical and
576     calculational effects on the computed conductivity.
577    
578     \subsection{Effects due to capping agent coverage}
579    
580     A series of different initial conditions with a range of surface
581     coverages was prepared and solvated with various with both of the
582     solvent molecules. These systems were then equilibrated and their
583     interfacial thermal conductivity was measured with the NIVS
584     algorithm. Figure \ref{coverage} demonstrates the trend of conductance
585     with respect to surface coverage.
586    
587     \begin{figure}
588     \includegraphics[width=\linewidth]{coverage}
589     \caption{The interfacial thermal conductivity ($G$) has a
590     non-monotonic dependence on the degree of surface capping. This
591     data is for the Au(111) / butanethiol / solvent interface with
592     various UA force fields at $\langle T\rangle \sim $200K.}
593     \label{coverage}
594     \end{figure}
595    
596     In partially covered surfaces, the derivative definition for
597     $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
598     location of maximum change of $\lambda$ becomes washed out. The
599     discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
600     Gibbs dividing surface is still well-defined. Therefore, $G$ (not
601     $G^\prime$) was used in this section.
602    
603     From Figure \ref{coverage}, one can see the significance of the
604     presence of capping agents. When even a small fraction of the Au(111)
605     surface sites are covered with butanethiols, the conductivity exhibits
606     an enhancement by at least a factor of 3. Capping agents are clearly
607     playing a major role in thermal transport at metal / organic solvent
608     surfaces.
609    
610     We note a non-monotonic behavior in the interfacial conductance as a
611     function of surface coverage. The maximum conductance (largest $G$)
612     happens when the surfaces are about 75\% covered with butanethiol
613     caps. The reason for this behavior is not entirely clear. One
614     explanation is that incomplete butanethiol coverage allows small gaps
615     between butanethiols to form. These gaps can be filled by transient
616     solvent molecules. These solvent molecules couple very strongly with
617     the hot capping agent molecules near the surface, and can then carry
618     away (diffusively) the excess thermal energy from the surface.
619    
620     There appears to be a competition between the conduction of the
621     thermal energy away from the surface by the capping agents (enhanced
622     by greater coverage) and the coupling of the capping agents with the
623     solvent (enhanced by interdigitation at lower coverages). This
624     competition would lead to the non-monotonic coverage behavior observed
625     here.
626    
627     Results for rigid body toluene solvent, as well as the UA hexane, are
628     within the ranges expected from prior experimental
629     work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
630     that explicit hydrogen atoms might not be required for modeling
631     thermal transport in these systems. C-H vibrational modes do not see
632     significant excited state population at low temperatures, and are not
633     likely to carry lower frequency excitations from the solid layer into
634     the bulk liquid.
635    
636     The toluene solvent does not exhibit the same behavior as hexane in
637     that $G$ remains at approximately the same magnitude when the capping
638     coverage increases from 25\% to 75\%. Toluene, as a rigid planar
639     molecule, cannot occupy the relatively small gaps between the capping
640     agents as easily as the chain-like {\it n}-hexane. The effect of
641     solvent coupling to the capping agent is therefore weaker in toluene
642     except at the very lowest coverage levels. This effect counters the
643     coverage-dependent conduction of heat away from the metal surface,
644     leading to a much flatter $G$ vs. coverage trend than is observed in
645     {\it n}-hexane.
646    
647     \subsection{Effects due to Solvent \& Solvent Models}
648     In addition to UA solvent and capping agent models, AA models have
649     also been included in our simulations. In most of this work, the same
650     (UA or AA) model for solvent and capping agent was used, but it is
651     also possible to utilize different models for different components.
652     We have also included isotopic substitutions (Hydrogen to Deuterium)
653     to decrease the explicit vibrational overlap between solvent and
654     capping agent. Table \ref{modelTest} summarizes the results of these
655     studies.
656    
657     \begin{table*}
658     \begin{minipage}{\linewidth}
659     \begin{center}
660    
661     \caption{Computed interfacial thermal conductance ($G$ and
662     $G^\prime$) values for interfaces using various models for
663     solvent and capping agent (or without capping agent) at
664     $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
665     solvent or capping agent molecules. Error estimates are
666     indicated in parentheses.}
667    
668     \begin{tabular}{llccc}
669     \hline\hline
670     Butanethiol model & Solvent & $G$ & $G^\prime$ \\
671     (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
672     \hline
673     UA & UA hexane & 131(9) & 87(10) \\
674     & UA hexane(D) & 153(5) & 136(13) \\
675     & AA hexane & 131(6) & 122(10) \\
676     & UA toluene & 187(16) & 151(11) \\
677     & AA toluene & 200(36) & 149(53) \\
678     \hline
679     AA & UA hexane & 116(9) & 129(8) \\
680     & AA hexane & 442(14) & 356(31) \\
681     & AA hexane(D) & 222(12) & 234(54) \\
682     & UA toluene & 125(25) & 97(60) \\
683     & AA toluene & 487(56) & 290(42) \\
684     \hline
685     AA(D) & UA hexane & 158(25) & 172(4) \\
686     & AA hexane & 243(29) & 191(11) \\
687     & AA toluene & 364(36) & 322(67) \\
688     \hline
689     bare & UA hexane & 46.5(3.2) & 49.4(4.5) \\
690     & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
691     & AA hexane & 31.0(1.4) & 29.4(1.3) \\
692     & UA toluene & 70.1(1.3) & 65.8(0.5) \\
693     \hline\hline
694     \end{tabular}
695     \label{modelTest}
696     \end{center}
697     \end{minipage}
698     \end{table*}
699    
700     To facilitate direct comparison between force fields, systems with the
701     same capping agent and solvent were prepared with the same length
702     scales for the simulation cells.
703    
704     On bare metal / solvent surfaces, different force field models for
705     hexane yield similar results for both $G$ and $G^\prime$, and these
706     two definitions agree with each other very well. This is primarily an
707     indicator of weak interactions between the metal and the solvent.
708    
709     For the fully-covered surfaces, the choice of force field for the
710     capping agent and solvent has a large impact on the calculated values
711     of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
712     much larger than their UA to UA counterparts, and these values exceed
713     the experimental estimates by a large measure. The AA force field
714     allows significant energy to go into C-H (or C-D) stretching modes,
715     and since these modes are high frequency, this non-quantum behavior is
716     likely responsible for the overestimate of the conductivity. Compared
717     to the AA model, the UA model yields more reasonable conductivity
718     values with much higher computational efficiency.
719    
720     \subsubsection{Are electronic excitations in the metal important?}
721     Because they lack electronic excitations, the QSC and related embedded
722     atom method (EAM) models for gold are known to predict unreasonably
723     low values for bulk conductivity
724     ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
725     conductance between the phases ($G$) is governed primarily by phonon
726     excitation (and not electronic degrees of freedom), one would expect a
727     classical model to capture most of the interfacial thermal
728     conductance. Our results for $G$ and $G^\prime$ indicate that this is
729     indeed the case, and suggest that the modeling of interfacial thermal
730     transport depends primarily on the description of the interactions
731     between the various components at the interface. When the metal is
732     chemically capped, the primary barrier to thermal conductivity appears
733     to be the interface between the capping agent and the surrounding
734     solvent, so the excitations in the metal have little impact on the
735     value of $G$.
736    
737     \subsection{Effects due to methodology and simulation parameters}
738    
739     We have varied the parameters of the simulations in order to
740     investigate how these factors would affect the computation of $G$. Of
741     particular interest are: 1) the length scale for the applied thermal
742     gradient (modified by increasing the amount of solvent in the system),
743     2) the sign and magnitude of the applied thermal flux, 3) the average
744     temperature of the simulation (which alters the solvent density during
745     equilibration), and 4) the definition of the interfacial conductance
746     (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
747     calculation.
748    
749     Systems of different lengths were prepared by altering the number of
750     solvent molecules and extending the length of the box along the $z$
751     axis to accomodate the extra solvent. Equilibration at the same
752     temperature and pressure conditions led to nearly identical surface
753     areas ($L_x$ and $L_y$) available to the metal and capping agent,
754     while the extra solvent served mainly to lengthen the axis that was
755     used to apply the thermal flux. For a given value of the applied
756     flux, the different $z$ length scale has only a weak effect on the
757     computed conductivities.
758    
759     \subsubsection{Effects of applied flux}
760     The NIVS algorithm allows changes in both the sign and magnitude of
761     the applied flux. It is possible to reverse the direction of heat
762     flow simply by changing the sign of the flux, and thermal gradients
763     which would be difficult to obtain experimentally ($5$ K/\AA) can be
764     easily simulated. However, the magnitude of the applied flux is not
765     arbitrary if one aims to obtain a stable and reliable thermal gradient.
766     A temperature gradient can be lost in the noise if $|J_z|$ is too
767     small, and excessive $|J_z|$ values can cause phase transitions if the
768     extremes of the simulation cell become widely separated in
769     temperature. Also, if $|J_z|$ is too large for the bulk conductivity
770     of the materials, the thermal gradient will never reach a stable
771     state.
772    
773     Within a reasonable range of $J_z$ values, we were able to study how
774     $G$ changes as a function of this flux. In what follows, we use
775     positive $J_z$ values to denote the case where energy is being
776     transferred by the method from the metal phase and into the liquid.
777     The resulting gradient therefore has a higher temperature in the
778     liquid phase. Negative flux values reverse this transfer, and result
779     in higher temperature metal phases. The conductance measured under
780     different applied $J_z$ values is listed in Tables 2 and 3 in the
781     supporting information. These results do not indicate that $G$ depends
782     strongly on $J_z$ within this flux range. The linear response of flux
783     to thermal gradient simplifies our investigations in that we can rely
784     on $G$ measurement with only a small number $J_z$ values.
785    
786     The sign of $J_z$ is a different matter, however, as this can alter
787     the temperature on the two sides of the interface. The average
788     temperature values reported are for the entire system, and not for the
789     liquid phase, so at a given $\langle T \rangle$, the system with
790     positive $J_z$ has a warmer liquid phase. This means that if the
791     liquid carries thermal energy via diffusive transport, {\it positive}
792     $J_z$ values will result in increased molecular motion on the liquid
793     side of the interface, and this will increase the measured
794     conductivity.
795    
796     \subsubsection{Effects due to average temperature}
797    
798     We also studied the effect of average system temperature on the
799     interfacial conductance. The simulations are first equilibrated in
800     the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
801     predict a lower boiling point (and liquid state density) than
802     experiments. This lower-density liquid phase leads to reduced contact
803     between the hexane and butanethiol, and this accounts for our
804     observation of lower conductance at higher temperatures. In raising
805     the average temperature from 200K to 250K, the density drop of
806     $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
807     conductance.
808    
809     Similar behavior is observed in the TraPPE-UA model for toluene,
810     although this model has better agreement with the experimental
811     densities of toluene. The expansion of the toluene liquid phase is
812     not as significant as that of the hexane (8.3\% over 100K), and this
813     limits the effect to $\sim$20\% drop in thermal conductivity.
814    
815     Although we have not mapped out the behavior at a large number of
816     temperatures, is clear that there will be a strong temperature
817     dependence in the interfacial conductance when the physical properties
818     of one side of the interface (notably the density) change rapidly as a
819     function of temperature.
820    
821     Besides the lower interfacial thermal conductance, surfaces at
822     relatively high temperatures are susceptible to reconstructions,
823     particularly when butanethiols fully cover the Au(111) surface. These
824     reconstructions include surface Au atoms which migrate outward to the
825     S atom layer, and butanethiol molecules which embed into the surface
826     Au layer. The driving force for this behavior is the strong Au-S
827     interactions which are modeled here with a deep Lennard-Jones
828     potential. This phenomenon agrees with reconstructions that have been
829     experimentally
830     observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
831     {\it et al.} kept their Au(111) slab rigid so that their simulations
832     could reach 300K without surface
833     reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
834     blur the interface, the measurement of $G$ becomes more difficult to
835     conduct at higher temperatures. For this reason, most of our
836     measurements are undertaken at $\langle T\rangle\sim$200K where
837     reconstruction is minimized.
838    
839     However, when the surface is not completely covered by butanethiols,
840     the simulated system appears to be more resistent to the
841     reconstruction. Our Au / butanethiol / toluene system had the Au(111)
842     surfaces 90\% covered by butanethiols, but did not see this above
843     phenomena even at $\langle T\rangle\sim$300K. That said, we did
844     observe butanethiols migrating to neighboring three-fold sites during
845     a simulation. Since the interface persisted in these simulations, we
846     were able to obtain $G$'s for these interfaces even at a relatively
847     high temperature without being affected by surface reconstructions.
848    
849     \section{Discussion}
850 skuang 3770 [COMBINE W. RESULTS]
851 gezelter 3769 The primary result of this work is that the capping agent acts as an
852     efficient thermal coupler between solid and solvent phases. One of
853     the ways the capping agent can carry out this role is to down-shift
854     between the phonon vibrations in the solid (which carry the heat from
855     the gold) and the molecular vibrations in the liquid (which carry some
856     of the heat in the solvent).
857    
858     To investigate the mechanism of interfacial thermal conductance, the
859     vibrational power spectrum was computed. Power spectra were taken for
860     individual components in different simulations. To obtain these
861     spectra, simulations were run after equilibration in the
862     microcanonical (NVE) ensemble and without a thermal
863     gradient. Snapshots of configurations were collected at a frequency
864     that is higher than that of the fastest vibrations occurring in the
865     simulations. With these configurations, the velocity auto-correlation
866     functions can be computed:
867     \begin{equation}
868     C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
869     \label{vCorr}
870     \end{equation}
871     The power spectrum is constructed via a Fourier transform of the
872     symmetrized velocity autocorrelation function,
873     \begin{equation}
874     \hat{f}(\omega) =
875     \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
876     \label{fourier}
877     \end{equation}
878    
879     \subsection{The role of specific vibrations}
880     The vibrational spectra for gold slabs in different environments are
881     shown as in Figure \ref{specAu}. Regardless of the presence of
882     solvent, the gold surfaces which are covered by butanethiol molecules
883     exhibit an additional peak observed at a frequency of
884     $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
885     vibration. This vibration enables efficient thermal coupling of the
886     surface Au layer to the capping agents. Therefore, in our simulations,
887     the Au / S interfaces do not appear to be the primary barrier to
888     thermal transport when compared with the butanethiol / solvent
889     interfaces. This supports the results of Luo {\it et
890     al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
891     twice as large as what we have computed for the thiol-liquid
892     interfaces.
893    
894     \begin{figure}
895     \includegraphics[width=\linewidth]{vibration}
896     \caption{The vibrational power spectrum for thiol-capped gold has an
897     additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
898     surfaces (both with and without a solvent over-layer) are missing
899     this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
900     the vibrational power spectrum for the butanethiol capping agents.}
901     \label{specAu}
902     \end{figure}
903    
904     Also in this figure, we show the vibrational power spectrum for the
905     bound butanethiol molecules, which also exhibits the same
906     $\sim$165cm$^{-1}$ peak.
907    
908     \subsection{Overlap of power spectra}
909     A comparison of the results obtained from the two different organic
910     solvents can also provide useful information of the interfacial
911     thermal transport process. In particular, the vibrational overlap
912     between the butanethiol and the organic solvents suggests a highly
913     efficient thermal exchange between these components. Very high
914     thermal conductivity was observed when AA models were used and C-H
915     vibrations were treated classically. The presence of extra degrees of
916     freedom in the AA force field yields higher heat exchange rates
917     between the two phases and results in a much higher conductivity than
918     in the UA force field. The all-atom classical models include high
919     frequency modes which should be unpopulated at our relatively low
920     temperatures. This artifact is likely the cause of the high thermal
921     conductance in all-atom MD simulations.
922    
923     The similarity in the vibrational modes available to solvent and
924     capping agent can be reduced by deuterating one of the two components
925     (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
926     are deuterated, one can observe a significantly lower $G$ and
927     $G^\prime$ values (Table \ref{modelTest}).
928    
929     \begin{figure}
930     \includegraphics[width=\linewidth]{aahxntln}
931     \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
932     systems. When butanethiol is deuterated (lower left), its
933     vibrational overlap with hexane decreases significantly. Since
934     aromatic molecules and the butanethiol are vibrationally dissimilar,
935     the change is not as dramatic when toluene is the solvent (right).}
936     \label{aahxntln}
937     \end{figure}
938    
939     For the Au / butanethiol / toluene interfaces, having the AA
940     butanethiol deuterated did not yield a significant change in the
941     measured conductance. Compared to the C-H vibrational overlap between
942     hexane and butanethiol, both of which have alkyl chains, the overlap
943     between toluene and butanethiol is not as significant and thus does
944     not contribute as much to the heat exchange process.
945    
946     Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
947     that the {\it intra}molecular heat transport due to alkylthiols is
948     highly efficient. Combining our observations with those of Zhang {\it
949     et al.}, it appears that butanethiol acts as a channel to expedite
950     heat flow from the gold surface and into the alkyl chain. The
951     vibrational coupling between the metal and the liquid phase can
952     therefore be enhanced with the presence of suitable capping agents.
953    
954     Deuterated models in the UA force field did not decouple the thermal
955     transport as well as in the AA force field. The UA models, even
956     though they have eliminated the high frequency C-H vibrational
957     overlap, still have significant overlap in the lower-frequency
958     portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
959     the UA models did not decouple the low frequency region enough to
960     produce an observable difference for the results of $G$ (Table
961     \ref{modelTest}).
962    
963     \begin{figure}
964     \includegraphics[width=\linewidth]{uahxnua}
965     \caption{Vibrational power spectra for UA models for the butanethiol
966     and hexane solvent (upper panel) show the high degree of overlap
967     between these two molecules, particularly at lower frequencies.
968     Deuterating a UA model for the solvent (lower panel) does not
969     decouple the two spectra to the same degree as in the AA force
970     field (see Fig \ref{aahxntln}).}
971     \label{uahxnua}
972     \end{figure}
973    
974     \section{Conclusions}
975     The NIVS algorithm has been applied to simulations of
976     butanethiol-capped Au(111) surfaces in the presence of organic
977     solvents. This algorithm allows the application of unphysical thermal
978     flux to transfer heat between the metal and the liquid phase. With the
979     flux applied, we were able to measure the corresponding thermal
980     gradients and to obtain interfacial thermal conductivities. Under
981     steady states, 2-3 ns trajectory simulations are sufficient for
982     computation of this quantity.
983    
984     Our simulations have seen significant conductance enhancement in the
985     presence of capping agent, compared with the bare gold / liquid
986     interfaces. The vibrational coupling between the metal and the liquid
987     phase is enhanced by a chemically-bonded capping agent. Furthermore,
988     the coverage percentage of the capping agent plays an important role
989     in the interfacial thermal transport process. Moderately low coverages
990     allow higher contact between capping agent and solvent, and thus could
991     further enhance the heat transfer process, giving a non-monotonic
992     behavior of conductance with increasing coverage.
993    
994     Our results, particularly using the UA models, agree well with
995     available experimental data. The AA models tend to overestimate the
996     interfacial thermal conductance in that the classically treated C-H
997     vibrations become too easily populated. Compared to the AA models, the
998     UA models have higher computational efficiency with satisfactory
999     accuracy, and thus are preferable in modeling interfacial thermal
1000     transport.
1001    
1002     Of the two definitions for $G$, the discrete form
1003     (Eq. \ref{discreteG}) was easier to use and gives out relatively
1004     consistent results, while the derivative form (Eq. \ref{derivativeG})
1005     is not as versatile. Although $G^\prime$ gives out comparable results
1006     and follows similar trend with $G$ when measuring close to fully
1007     covered or bare surfaces, the spatial resolution of $T$ profile
1008     required for the use of a derivative form is limited by the number of
1009     bins and the sampling required to obtain thermal gradient information.
1010    
1011     Vlugt {\it et al.} have investigated the surface thiol structures for
1012     nanocrystalline gold and pointed out that they differ from those of
1013     the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
1014     difference could also cause differences in the interfacial thermal
1015     transport behavior. To investigate this problem, one would need an
1016     effective method for applying thermal gradients in non-planar
1017     (i.e. spherical) geometries.
1018    
1019     \section{Acknowledgments}
1020     Support for this project was provided by the National Science
1021     Foundation under grant CHE-0848243. Computational time was provided by
1022     the Center for Research Computing (CRC) at the University of Notre
1023     Dame.
1024    
1025     \newpage
1026    
1027 skuang 3770 \bibliography{stokes}
1028 gezelter 3769
1029     \end{doublespace}
1030     \end{document}
1031