ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3774
Committed: Wed Dec 7 03:39:41 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 38671 byte(s)
Log Message:
more stuff in simulation details

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 3771 [REFINE LATER, ADD MORE REF.S]
78     Imposed-flux methods in Molecular Dynamics (MD)
79     simulations\cite{MullerPlathe:1997xw} can establish steady state
80     systems with a set applied flux vs a corresponding gradient that can
81     be measured. These methods does not need many trajectories to provide
82     information of transport properties of a given system. Thus, they are
83     utilized in computing thermal and mechanical transfer of homogeneous
84     or bulk systems as well as heterogeneous systems such as liquid-solid
85     interfaces.\cite{kuang:AuThl}
86 skuang 3770
87 skuang 3771 The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that
88     satisfy linear momentum and total energy conservation of a system when
89     imposing fluxes in a simulation. Thus they are compatible with various
90     ensembles, including the micro-canonical (NVE) ensemble, without the
91     need of an external thermostat. The original approaches by
92     M\"{u}ller-Plathe {\it et
93     al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
94     momentum swapping for generating energy/momentum fluxes, which is also
95     compatible with particles of different identities. Although simple to
96     implement in a simulation, this approach can create nonthermal
97     velocity distributions, as discovered by Tenney and
98     Maginn\cite{Maginn:2010}. Furthermore, this approach to kinetic energy
99     transfer between particles of different identities is less efficient
100     when the mass difference between the particles becomes significant,
101     which also limits its application on heterogeneous interfacial
102     systems.
103 skuang 3770
104 skuang 3771 Recently, we developed a different approach, using Non-Isotropic
105     Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose
106     fluxes. Compared to the momentum swapping move, it scales the velocity
107     vectors in two separate regions of a simulated system with respective
108     diagonal scaling matrices. These matrices are determined by solving a
109     set of equations including linear momentum and kinetic energy
110     conservation constraints and target flux satisfaction. This method is
111     able to effectively impose a wide range of kinetic energy fluxes
112     without obvious perturbation to the velocity distributions of the
113     simulated systems, regardless of the presence of heterogeneous
114     interfaces. We have successfully applied this approach in studying the
115     interfacial thermal conductance at metal-solvent
116     interfaces.\cite{kuang:AuThl}
117 gezelter 3769
118 skuang 3771 However, the NIVS approach limits its application in imposing momentum
119     fluxes. Temperature anisotropy can happen under high momentum fluxes,
120     due to the nature of the algorithm. Thus, combining thermal and
121     momentum flux is also difficult to implement with this
122     approach. However, such combination may provide a means to simulate
123     thermal/momentum gradient coupled processes such as freeze
124     desalination. Therefore, developing novel approaches to extend the
125     application of imposed-flux method is desired.
126 gezelter 3769
127 skuang 3771 In this paper, we improve the NIVS method and propose a novel approach
128     to impose fluxes. This approach separate the means of applying
129     momentum and thermal flux with operations in one time step and thus is
130     able to simutaneously impose thermal and momentum flux. Furthermore,
131     the approach retains desirable features of previous RNEMD approaches
132     and is simpler to implement compared to the NIVS method. In what
133     follows, we first present the method to implement the method in a
134     simulation. Then we compare the method on bulk fluids to previous
135     methods. Also, interfacial frictions are computed for a series of
136     interfaces.
137 gezelter 3769
138     \section{Methodology}
139 skuang 3770 Similar to the NIVS methodology,\cite{kuang:164101} we consider a
140     periodic system divided into a series of slabs along a certain axis
141     (e.g. $z$). The unphysical thermal and/or momentum flux is designated
142     from the center slab to one of the end slabs, and thus the center slab
143     would have a lower temperature than the end slab (unless the thermal
144     flux is negative). Therefore, the center slab is denoted as ``$c$''
145     while the end slab as ``$h$''.
146    
147     To impose these fluxes, we periodically apply separate operations to
148     velocities of particles {$i$} within the center slab and of particles
149     {$j$} within the end slab:
150     \begin{eqnarray}
151     \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
152     \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
153     \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
154     \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
155     \end{eqnarray}
156     where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes
157     the instantaneous bulk velocity of slabs $c$ and $h$ respectively
158     before an operation occurs. When a momentum flux $\vec{j}_z(\vec{p})$
159     presents, these bulk velocities would have a corresponding change
160     ($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's
161     second law:
162     \begin{eqnarray}
163     M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\
164     M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
165     \end{eqnarray}
166     where
167     \begin{eqnarray}
168     M_c & = & \sum_{i = 1}^{N_c} m_i \\
169     M_h & = & \sum_{j = 1}^{N_h} m_j
170     \end{eqnarray}
171     and $\Delta t$ is the interval between two operations.
172    
173     The above operations conserve the linear momentum of a periodic
174     system. To satisfy total energy conservation as well as to impose a
175     thermal flux $J_z$, one would have
176 skuang 3771 [SUPPORT INFO MIGHT BE NECESSARY TO PUT EXTRA MATH IN]
177 skuang 3770 \begin{eqnarray}
178     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
179     \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
180     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
181     \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2
182     \end{eqnarray}
183     where $K_c$ and $K_h$ denotes translational kinetic energy of slabs
184     $c$ and $h$ respectively before an operation occurs. These
185     translational kinetic energy conservation equations are sufficient to
186     ensure total energy conservation, as the operations applied do not
187     change the potential energy of a system, given that the potential
188     energy does not depend on particle velocity.
189    
190     The above sets of equations are sufficient to determine the velocity
191     scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
192     $\vec{a}_h$. Note that two roots of $c$ and $h$ exist
193     respectively. However, to avoid dramatic perturbations to a system,
194 skuang 3772 the positive roots (which are closer to 1) are chosen. Figure
195     \ref{method} illustrates the implementation of this algorithm in an
196     individual step.
197 skuang 3770
198 skuang 3772 \begin{figure}
199     \includegraphics[width=\linewidth]{method}
200     \caption{Illustration of the implementation of the algorithm in a
201     single step. Starting from an ideal velocity distribution, the
202     transformation is used to apply both thermal and momentum flux from
203     the ``c'' slab to the ``h'' slab. As the figure shows, the thermal
204     distributions preserve after this operation.}
205     \label{method}
206     \end{figure}
207    
208 skuang 3770 By implementing these operations at a certain frequency, a steady
209     thermal and/or momentum flux can be applied and the corresponding
210     temperature and/or momentum gradients can be established.
211    
212 skuang 3771 This approach is more computationaly efficient compared to the
213     previous NIVS method, in that only quadratic equations are involved,
214     while the NIVS method needs to solve a quartic equations. Furthermore,
215     the method implements isotropic scaling of velocities in respective
216     slabs, unlike the NIVS, where an extra criteria function is necessary
217     to choose a set of coefficients that performs the most isotropic
218     scaling. More importantly, separating the momentum flux imposing from
219     velocity scaling avoids the underlying cause that NIVS produced
220     thermal anisotropy when applying a momentum flux.
221     %NEW METHOD DOESN'T CAUSE UNDESIRED CONCOMITENT MOMENTUM FLUX WHEN
222     %IMPOSING A THERMAL FLUX
223 gezelter 3769
224 skuang 3772 The advantages of the approach over the original momentum swapping
225     approach lies in its nature to preserve a Gaussian
226     distribution. Because the momentum swapping tends to render a
227     nonthermal distribution, when the imposed flux is relatively large,
228     diffusion of the neighboring slabs could no longer remedy this effect,
229     and nonthermal distributions would be observed. Results in later
230     section will illustrate this effect.
231 skuang 3773 %NEW METHOD (AND NIVS) HAVE LESS PERTURBATION THAN MOMENTUM SWAPPING
232 gezelter 3769
233 skuang 3772 \section{Computational Details}
234 skuang 3773 The algorithm has been implemented in our MD simulation code,
235     OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with
236     previous RNEMD methods or equilibrium MD methods in homogeneous fluids
237     (Lennard-Jones and SPC/E water). And taking advantage of the method,
238     we simulate the interfacial friction of different heterogeneous
239     interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid
240     water).
241 gezelter 3769
242 skuang 3773 \subsection{Simulation Protocols}
243     The systems to be investigated are set up in a orthorhombic simulation
244     cell with periodic boundary conditions in all three dimensions. The
245 skuang 3774 $z$ axis of these cells were longer and was set as the gradient axis
246 skuang 3773 of temperature and/or momentum. Thus the cells were divided into $N$
247     slabs along this axis, with various $N$ depending on individual
248     system. The $x$ and $y$ axis were usually of the same length in
249     homogeneous systems or close to each other where interfaces
250     presents. In all cases, before introducing a nonequilibrium method to
251     establish steady thermal and/or momentum gradients for further
252     measurements and calculations, canonical ensemble with a Nos\'e-Hoover
253     thermostat\cite{hoover85} and microcanonical ensemble equilibrations
254     were used to prepare systems ready for data
255     collections. Isobaric-isothermal equilibrations are performed before
256     this for SPC/E water systems to reach normal pressure (1 bar), while
257     similar equilibrations are used for interfacial systems to relax the
258     surface tensions.
259 skuang 3772
260 skuang 3773 While homogeneous fluid systems can be set up with random
261     configurations, our interfacial systems needs extra steps to ensure
262 skuang 3774 the interfaces be established properly for computations. The
263     preparation and equilibration of butanethiol covered gold (111)
264     surface and further solvation and equilibration process is described
265     as in reference \ref{kuang:AuThl}.
266 skuang 3773
267 skuang 3774 As for the ice/liquid water interfaces, the basal surface of ice
268     lattice was first constructed. [REF JPCB PAPER] explored the energeics
269     of ice lattices with different proton orders. We refer to their
270     results and choose the configuration of the lowest energy after
271     relaxations as the unit cells of our ice lattices. Although
272     experimental solid/liquid coexistant temperature near normal pressure
273     is 273K, [REF HAYMET] simulations of ice/liquid water interfaces with
274     different models suggest that for SPC/E, stable interfaces could only
275     exist no higher than 225K. Therefore, all our ice/liquid water
276     simulations were carried out under 225K. To have extra protection of
277     the ice lattice during initial equilibration (when the randomly
278     generated liquid configuration could release large amount of energy in
279     relaxation), a constraint method (REF?) was adopted until the high
280     energy configuration was relaxed.
281 gezelter 3769
282     \subsection{Force Field Parameters}
283 skuang 3774 For comparison of our new method with previous work, we retain our
284     force field parameters consistent with the results we will compare
285     with. The Lennard-Jones Fluid used here is argon, and reduced unit
286     results are reported when it is favorable for comparison purpose.
287 gezelter 3769
288 skuang 3774 As for our water simulations, SPC/E model is used throughout this work
289     for consistency. Previous work for transport properties of SPC/E water
290     model is available so that unnecessary repetition of previous methods
291     can be avoided.
292 gezelter 3769
293 skuang 3774 The Au-Au interaction parameters in all simulations are described by
294     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
295     QSC potentials include zero-point quantum corrections and are
296 gezelter 3769 reparametrized for accurate surface energies compared to the
297 skuang 3774 Sutton-Chen potentials.\cite{Chen90} When Au-H$_2$O interactions are
298     involved, the Spohr potential was adopted.[CITE NIVS REF.46]
299 gezelter 3769
300 skuang 3774 The small organic molecules included in our simulations are the Au
301     surface capping agent butanethiol and liquid hexane and toluene. The
302     United-Atom
303     models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
304     for these components were used in this work for better computational
305     efficiency, while maintaining good accuracy. We refer readers to our
306     previous work\cite{kuang:AuThl} for further details of these models,
307     as well as the interactions between Au and the above organic molecule
308     components.
309 gezelter 3769
310 skuang 3774 \subsection{Thermal conductivities}
311     \subsection{Shear viscosities}
312     \subsection{Interfacial friction and Slip length}
313 gezelter 3769
314    
315     \section{Results}
316 skuang 3773 [L-J COMPARED TO RNEMD NIVS; WATER COMPARED TO RNEMD NIVS AND EMD;
317 skuang 3770 SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
318    
319 gezelter 3769 There are many factors contributing to the measured interfacial
320     conductance; some of these factors are physically motivated
321     (e.g. coverage of the surface by the capping agent coverage and
322     solvent identity), while some are governed by parameters of the
323     methodology (e.g. applied flux and the formulas used to obtain the
324     conductance). In this section we discuss the major physical and
325     calculational effects on the computed conductivity.
326    
327     \subsection{Effects due to capping agent coverage}
328    
329     A series of different initial conditions with a range of surface
330     coverages was prepared and solvated with various with both of the
331     solvent molecules. These systems were then equilibrated and their
332     interfacial thermal conductivity was measured with the NIVS
333     algorithm. Figure \ref{coverage} demonstrates the trend of conductance
334     with respect to surface coverage.
335    
336     \begin{figure}
337     \includegraphics[width=\linewidth]{coverage}
338     \caption{The interfacial thermal conductivity ($G$) has a
339     non-monotonic dependence on the degree of surface capping. This
340     data is for the Au(111) / butanethiol / solvent interface with
341     various UA force fields at $\langle T\rangle \sim $200K.}
342     \label{coverage}
343     \end{figure}
344    
345     In partially covered surfaces, the derivative definition for
346     $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
347     location of maximum change of $\lambda$ becomes washed out. The
348     discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
349     Gibbs dividing surface is still well-defined. Therefore, $G$ (not
350     $G^\prime$) was used in this section.
351    
352     From Figure \ref{coverage}, one can see the significance of the
353     presence of capping agents. When even a small fraction of the Au(111)
354     surface sites are covered with butanethiols, the conductivity exhibits
355     an enhancement by at least a factor of 3. Capping agents are clearly
356     playing a major role in thermal transport at metal / organic solvent
357     surfaces.
358    
359     We note a non-monotonic behavior in the interfacial conductance as a
360     function of surface coverage. The maximum conductance (largest $G$)
361     happens when the surfaces are about 75\% covered with butanethiol
362     caps. The reason for this behavior is not entirely clear. One
363     explanation is that incomplete butanethiol coverage allows small gaps
364     between butanethiols to form. These gaps can be filled by transient
365     solvent molecules. These solvent molecules couple very strongly with
366     the hot capping agent molecules near the surface, and can then carry
367     away (diffusively) the excess thermal energy from the surface.
368    
369     There appears to be a competition between the conduction of the
370     thermal energy away from the surface by the capping agents (enhanced
371     by greater coverage) and the coupling of the capping agents with the
372     solvent (enhanced by interdigitation at lower coverages). This
373     competition would lead to the non-monotonic coverage behavior observed
374     here.
375    
376     Results for rigid body toluene solvent, as well as the UA hexane, are
377     within the ranges expected from prior experimental
378     work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
379     that explicit hydrogen atoms might not be required for modeling
380     thermal transport in these systems. C-H vibrational modes do not see
381     significant excited state population at low temperatures, and are not
382     likely to carry lower frequency excitations from the solid layer into
383     the bulk liquid.
384    
385     The toluene solvent does not exhibit the same behavior as hexane in
386     that $G$ remains at approximately the same magnitude when the capping
387     coverage increases from 25\% to 75\%. Toluene, as a rigid planar
388     molecule, cannot occupy the relatively small gaps between the capping
389     agents as easily as the chain-like {\it n}-hexane. The effect of
390     solvent coupling to the capping agent is therefore weaker in toluene
391     except at the very lowest coverage levels. This effect counters the
392     coverage-dependent conduction of heat away from the metal surface,
393     leading to a much flatter $G$ vs. coverage trend than is observed in
394     {\it n}-hexane.
395    
396     \subsection{Effects due to Solvent \& Solvent Models}
397     In addition to UA solvent and capping agent models, AA models have
398     also been included in our simulations. In most of this work, the same
399     (UA or AA) model for solvent and capping agent was used, but it is
400     also possible to utilize different models for different components.
401     We have also included isotopic substitutions (Hydrogen to Deuterium)
402     to decrease the explicit vibrational overlap between solvent and
403     capping agent. Table \ref{modelTest} summarizes the results of these
404     studies.
405    
406     \begin{table*}
407     \begin{minipage}{\linewidth}
408     \begin{center}
409    
410     \caption{Computed interfacial thermal conductance ($G$ and
411     $G^\prime$) values for interfaces using various models for
412     solvent and capping agent (or without capping agent) at
413     $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
414     solvent or capping agent molecules. Error estimates are
415     indicated in parentheses.}
416    
417     \begin{tabular}{llccc}
418     \hline\hline
419     Butanethiol model & Solvent & $G$ & $G^\prime$ \\
420     (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
421     \hline
422     UA & UA hexane & 131(9) & 87(10) \\
423     & UA hexane(D) & 153(5) & 136(13) \\
424     & AA hexane & 131(6) & 122(10) \\
425     & UA toluene & 187(16) & 151(11) \\
426     & AA toluene & 200(36) & 149(53) \\
427     \hline
428     AA & UA hexane & 116(9) & 129(8) \\
429     & AA hexane & 442(14) & 356(31) \\
430     & AA hexane(D) & 222(12) & 234(54) \\
431     & UA toluene & 125(25) & 97(60) \\
432     & AA toluene & 487(56) & 290(42) \\
433     \hline
434     AA(D) & UA hexane & 158(25) & 172(4) \\
435     & AA hexane & 243(29) & 191(11) \\
436     & AA toluene & 364(36) & 322(67) \\
437     \hline
438     bare & UA hexane & 46.5(3.2) & 49.4(4.5) \\
439     & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
440     & AA hexane & 31.0(1.4) & 29.4(1.3) \\
441     & UA toluene & 70.1(1.3) & 65.8(0.5) \\
442     \hline\hline
443     \end{tabular}
444     \label{modelTest}
445     \end{center}
446     \end{minipage}
447     \end{table*}
448    
449     To facilitate direct comparison between force fields, systems with the
450     same capping agent and solvent were prepared with the same length
451     scales for the simulation cells.
452    
453     On bare metal / solvent surfaces, different force field models for
454     hexane yield similar results for both $G$ and $G^\prime$, and these
455     two definitions agree with each other very well. This is primarily an
456     indicator of weak interactions between the metal and the solvent.
457    
458     For the fully-covered surfaces, the choice of force field for the
459     capping agent and solvent has a large impact on the calculated values
460     of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
461     much larger than their UA to UA counterparts, and these values exceed
462     the experimental estimates by a large measure. The AA force field
463     allows significant energy to go into C-H (or C-D) stretching modes,
464     and since these modes are high frequency, this non-quantum behavior is
465     likely responsible for the overestimate of the conductivity. Compared
466     to the AA model, the UA model yields more reasonable conductivity
467     values with much higher computational efficiency.
468    
469     \subsubsection{Are electronic excitations in the metal important?}
470     Because they lack electronic excitations, the QSC and related embedded
471     atom method (EAM) models for gold are known to predict unreasonably
472     low values for bulk conductivity
473     ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
474     conductance between the phases ($G$) is governed primarily by phonon
475     excitation (and not electronic degrees of freedom), one would expect a
476     classical model to capture most of the interfacial thermal
477     conductance. Our results for $G$ and $G^\prime$ indicate that this is
478     indeed the case, and suggest that the modeling of interfacial thermal
479     transport depends primarily on the description of the interactions
480     between the various components at the interface. When the metal is
481     chemically capped, the primary barrier to thermal conductivity appears
482     to be the interface between the capping agent and the surrounding
483     solvent, so the excitations in the metal have little impact on the
484     value of $G$.
485    
486     \subsection{Effects due to methodology and simulation parameters}
487    
488     We have varied the parameters of the simulations in order to
489     investigate how these factors would affect the computation of $G$. Of
490     particular interest are: 1) the length scale for the applied thermal
491     gradient (modified by increasing the amount of solvent in the system),
492     2) the sign and magnitude of the applied thermal flux, 3) the average
493     temperature of the simulation (which alters the solvent density during
494     equilibration), and 4) the definition of the interfacial conductance
495     (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
496     calculation.
497    
498     Systems of different lengths were prepared by altering the number of
499     solvent molecules and extending the length of the box along the $z$
500     axis to accomodate the extra solvent. Equilibration at the same
501     temperature and pressure conditions led to nearly identical surface
502     areas ($L_x$ and $L_y$) available to the metal and capping agent,
503     while the extra solvent served mainly to lengthen the axis that was
504     used to apply the thermal flux. For a given value of the applied
505     flux, the different $z$ length scale has only a weak effect on the
506     computed conductivities.
507    
508     \subsubsection{Effects of applied flux}
509     The NIVS algorithm allows changes in both the sign and magnitude of
510     the applied flux. It is possible to reverse the direction of heat
511     flow simply by changing the sign of the flux, and thermal gradients
512     which would be difficult to obtain experimentally ($5$ K/\AA) can be
513     easily simulated. However, the magnitude of the applied flux is not
514     arbitrary if one aims to obtain a stable and reliable thermal gradient.
515     A temperature gradient can be lost in the noise if $|J_z|$ is too
516     small, and excessive $|J_z|$ values can cause phase transitions if the
517     extremes of the simulation cell become widely separated in
518     temperature. Also, if $|J_z|$ is too large for the bulk conductivity
519     of the materials, the thermal gradient will never reach a stable
520     state.
521    
522     Within a reasonable range of $J_z$ values, we were able to study how
523     $G$ changes as a function of this flux. In what follows, we use
524     positive $J_z$ values to denote the case where energy is being
525     transferred by the method from the metal phase and into the liquid.
526     The resulting gradient therefore has a higher temperature in the
527     liquid phase. Negative flux values reverse this transfer, and result
528     in higher temperature metal phases. The conductance measured under
529     different applied $J_z$ values is listed in Tables 2 and 3 in the
530     supporting information. These results do not indicate that $G$ depends
531     strongly on $J_z$ within this flux range. The linear response of flux
532     to thermal gradient simplifies our investigations in that we can rely
533     on $G$ measurement with only a small number $J_z$ values.
534    
535     The sign of $J_z$ is a different matter, however, as this can alter
536     the temperature on the two sides of the interface. The average
537     temperature values reported are for the entire system, and not for the
538     liquid phase, so at a given $\langle T \rangle$, the system with
539     positive $J_z$ has a warmer liquid phase. This means that if the
540     liquid carries thermal energy via diffusive transport, {\it positive}
541     $J_z$ values will result in increased molecular motion on the liquid
542     side of the interface, and this will increase the measured
543     conductivity.
544    
545     \subsubsection{Effects due to average temperature}
546    
547     We also studied the effect of average system temperature on the
548     interfacial conductance. The simulations are first equilibrated in
549     the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
550     predict a lower boiling point (and liquid state density) than
551     experiments. This lower-density liquid phase leads to reduced contact
552     between the hexane and butanethiol, and this accounts for our
553     observation of lower conductance at higher temperatures. In raising
554     the average temperature from 200K to 250K, the density drop of
555     $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
556     conductance.
557    
558     Similar behavior is observed in the TraPPE-UA model for toluene,
559     although this model has better agreement with the experimental
560     densities of toluene. The expansion of the toluene liquid phase is
561     not as significant as that of the hexane (8.3\% over 100K), and this
562     limits the effect to $\sim$20\% drop in thermal conductivity.
563    
564     Although we have not mapped out the behavior at a large number of
565     temperatures, is clear that there will be a strong temperature
566     dependence in the interfacial conductance when the physical properties
567     of one side of the interface (notably the density) change rapidly as a
568     function of temperature.
569    
570     Besides the lower interfacial thermal conductance, surfaces at
571     relatively high temperatures are susceptible to reconstructions,
572     particularly when butanethiols fully cover the Au(111) surface. These
573     reconstructions include surface Au atoms which migrate outward to the
574     S atom layer, and butanethiol molecules which embed into the surface
575     Au layer. The driving force for this behavior is the strong Au-S
576     interactions which are modeled here with a deep Lennard-Jones
577     potential. This phenomenon agrees with reconstructions that have been
578     experimentally
579     observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
580     {\it et al.} kept their Au(111) slab rigid so that their simulations
581     could reach 300K without surface
582     reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
583     blur the interface, the measurement of $G$ becomes more difficult to
584     conduct at higher temperatures. For this reason, most of our
585     measurements are undertaken at $\langle T\rangle\sim$200K where
586     reconstruction is minimized.
587    
588     However, when the surface is not completely covered by butanethiols,
589     the simulated system appears to be more resistent to the
590     reconstruction. Our Au / butanethiol / toluene system had the Au(111)
591     surfaces 90\% covered by butanethiols, but did not see this above
592     phenomena even at $\langle T\rangle\sim$300K. That said, we did
593     observe butanethiols migrating to neighboring three-fold sites during
594     a simulation. Since the interface persisted in these simulations, we
595     were able to obtain $G$'s for these interfaces even at a relatively
596     high temperature without being affected by surface reconstructions.
597    
598     \section{Discussion}
599 skuang 3770 [COMBINE W. RESULTS]
600 gezelter 3769 The primary result of this work is that the capping agent acts as an
601     efficient thermal coupler between solid and solvent phases. One of
602     the ways the capping agent can carry out this role is to down-shift
603     between the phonon vibrations in the solid (which carry the heat from
604     the gold) and the molecular vibrations in the liquid (which carry some
605     of the heat in the solvent).
606    
607     To investigate the mechanism of interfacial thermal conductance, the
608     vibrational power spectrum was computed. Power spectra were taken for
609     individual components in different simulations. To obtain these
610     spectra, simulations were run after equilibration in the
611     microcanonical (NVE) ensemble and without a thermal
612     gradient. Snapshots of configurations were collected at a frequency
613     that is higher than that of the fastest vibrations occurring in the
614     simulations. With these configurations, the velocity auto-correlation
615     functions can be computed:
616     \begin{equation}
617     C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
618     \label{vCorr}
619     \end{equation}
620     The power spectrum is constructed via a Fourier transform of the
621     symmetrized velocity autocorrelation function,
622     \begin{equation}
623     \hat{f}(\omega) =
624     \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
625     \label{fourier}
626     \end{equation}
627    
628     \subsection{The role of specific vibrations}
629     The vibrational spectra for gold slabs in different environments are
630     shown as in Figure \ref{specAu}. Regardless of the presence of
631     solvent, the gold surfaces which are covered by butanethiol molecules
632     exhibit an additional peak observed at a frequency of
633     $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
634     vibration. This vibration enables efficient thermal coupling of the
635     surface Au layer to the capping agents. Therefore, in our simulations,
636     the Au / S interfaces do not appear to be the primary barrier to
637     thermal transport when compared with the butanethiol / solvent
638     interfaces. This supports the results of Luo {\it et
639     al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
640     twice as large as what we have computed for the thiol-liquid
641     interfaces.
642    
643     \begin{figure}
644     \includegraphics[width=\linewidth]{vibration}
645     \caption{The vibrational power spectrum for thiol-capped gold has an
646     additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
647     surfaces (both with and without a solvent over-layer) are missing
648     this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
649     the vibrational power spectrum for the butanethiol capping agents.}
650     \label{specAu}
651     \end{figure}
652    
653     Also in this figure, we show the vibrational power spectrum for the
654     bound butanethiol molecules, which also exhibits the same
655     $\sim$165cm$^{-1}$ peak.
656    
657     \subsection{Overlap of power spectra}
658     A comparison of the results obtained from the two different organic
659     solvents can also provide useful information of the interfacial
660     thermal transport process. In particular, the vibrational overlap
661     between the butanethiol and the organic solvents suggests a highly
662     efficient thermal exchange between these components. Very high
663     thermal conductivity was observed when AA models were used and C-H
664     vibrations were treated classically. The presence of extra degrees of
665     freedom in the AA force field yields higher heat exchange rates
666     between the two phases and results in a much higher conductivity than
667     in the UA force field. The all-atom classical models include high
668     frequency modes which should be unpopulated at our relatively low
669     temperatures. This artifact is likely the cause of the high thermal
670     conductance in all-atom MD simulations.
671    
672     The similarity in the vibrational modes available to solvent and
673     capping agent can be reduced by deuterating one of the two components
674     (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
675     are deuterated, one can observe a significantly lower $G$ and
676     $G^\prime$ values (Table \ref{modelTest}).
677    
678     \begin{figure}
679     \includegraphics[width=\linewidth]{aahxntln}
680     \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
681     systems. When butanethiol is deuterated (lower left), its
682     vibrational overlap with hexane decreases significantly. Since
683     aromatic molecules and the butanethiol are vibrationally dissimilar,
684     the change is not as dramatic when toluene is the solvent (right).}
685     \label{aahxntln}
686     \end{figure}
687    
688     For the Au / butanethiol / toluene interfaces, having the AA
689     butanethiol deuterated did not yield a significant change in the
690     measured conductance. Compared to the C-H vibrational overlap between
691     hexane and butanethiol, both of which have alkyl chains, the overlap
692     between toluene and butanethiol is not as significant and thus does
693     not contribute as much to the heat exchange process.
694    
695     Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
696     that the {\it intra}molecular heat transport due to alkylthiols is
697     highly efficient. Combining our observations with those of Zhang {\it
698     et al.}, it appears that butanethiol acts as a channel to expedite
699     heat flow from the gold surface and into the alkyl chain. The
700     vibrational coupling between the metal and the liquid phase can
701     therefore be enhanced with the presence of suitable capping agents.
702    
703     Deuterated models in the UA force field did not decouple the thermal
704     transport as well as in the AA force field. The UA models, even
705     though they have eliminated the high frequency C-H vibrational
706     overlap, still have significant overlap in the lower-frequency
707     portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
708     the UA models did not decouple the low frequency region enough to
709     produce an observable difference for the results of $G$ (Table
710     \ref{modelTest}).
711    
712     \begin{figure}
713     \includegraphics[width=\linewidth]{uahxnua}
714     \caption{Vibrational power spectra for UA models for the butanethiol
715     and hexane solvent (upper panel) show the high degree of overlap
716     between these two molecules, particularly at lower frequencies.
717     Deuterating a UA model for the solvent (lower panel) does not
718     decouple the two spectra to the same degree as in the AA force
719     field (see Fig \ref{aahxntln}).}
720     \label{uahxnua}
721     \end{figure}
722    
723     \section{Conclusions}
724     The NIVS algorithm has been applied to simulations of
725     butanethiol-capped Au(111) surfaces in the presence of organic
726     solvents. This algorithm allows the application of unphysical thermal
727     flux to transfer heat between the metal and the liquid phase. With the
728     flux applied, we were able to measure the corresponding thermal
729     gradients and to obtain interfacial thermal conductivities. Under
730     steady states, 2-3 ns trajectory simulations are sufficient for
731     computation of this quantity.
732    
733     Our simulations have seen significant conductance enhancement in the
734     presence of capping agent, compared with the bare gold / liquid
735     interfaces. The vibrational coupling between the metal and the liquid
736     phase is enhanced by a chemically-bonded capping agent. Furthermore,
737     the coverage percentage of the capping agent plays an important role
738     in the interfacial thermal transport process. Moderately low coverages
739     allow higher contact between capping agent and solvent, and thus could
740     further enhance the heat transfer process, giving a non-monotonic
741     behavior of conductance with increasing coverage.
742    
743     Our results, particularly using the UA models, agree well with
744     available experimental data. The AA models tend to overestimate the
745     interfacial thermal conductance in that the classically treated C-H
746     vibrations become too easily populated. Compared to the AA models, the
747     UA models have higher computational efficiency with satisfactory
748     accuracy, and thus are preferable in modeling interfacial thermal
749     transport.
750    
751     Of the two definitions for $G$, the discrete form
752     (Eq. \ref{discreteG}) was easier to use and gives out relatively
753     consistent results, while the derivative form (Eq. \ref{derivativeG})
754     is not as versatile. Although $G^\prime$ gives out comparable results
755     and follows similar trend with $G$ when measuring close to fully
756     covered or bare surfaces, the spatial resolution of $T$ profile
757     required for the use of a derivative form is limited by the number of
758     bins and the sampling required to obtain thermal gradient information.
759    
760     Vlugt {\it et al.} have investigated the surface thiol structures for
761     nanocrystalline gold and pointed out that they differ from those of
762     the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
763     difference could also cause differences in the interfacial thermal
764     transport behavior. To investigate this problem, one would need an
765     effective method for applying thermal gradients in non-planar
766     (i.e. spherical) geometries.
767    
768     \section{Acknowledgments}
769     Support for this project was provided by the National Science
770     Foundation under grant CHE-0848243. Computational time was provided by
771     the Center for Research Computing (CRC) at the University of Notre
772     Dame.
773    
774     \newpage
775    
776 skuang 3770 \bibliography{stokes}
777 gezelter 3769
778     \end{doublespace}
779     \end{document}
780