ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3773
Committed: Tue Dec 6 19:43:33 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 46267 byte(s)
Log Message:
just start 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     $z$ axis of these cells were longer and was used as the gradient axis
246     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     the interfaces be established properly for computations.
263     [AU(THIOL)ORGANIC SOLVENTS: REFER TO JPCC]
264     [ICE-WATER REFER TO OTHER REF.S]
265    
266 gezelter 3769 Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
267     under atmospheric pressure (1 atm) and 200K. After equilibration,
268     butanethiol capping agents were placed at three-fold hollow sites on
269     the Au(111) surfaces. These sites are either {\it fcc} or {\it
270     hcp} sites, although Hase {\it et al.} found that they are
271     equivalent in a heat transfer process,\cite{hase:2010} so we did not
272     distinguish between these sites in our study. The maximum butanethiol
273     capacity on Au surface is $1/3$ of the total number of surface Au
274     atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
275     structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
276     series of lower coverages was also prepared by eliminating
277     butanethiols from the higher coverage surface in a regular manner. The
278     lower coverages were prepared in order to study the relation between
279     coverage and interfacial conductance.
280    
281     The capping agent molecules were allowed to migrate during the
282     simulations. They distributed themselves uniformly and sampled a
283     number of three-fold sites throughout out study. Therefore, the
284     initial configuration does not noticeably affect the sampling of a
285     variety of configurations of the same coverage, and the final
286     conductance measurement would be an average effect of these
287     configurations explored in the simulations.
288    
289     After the modified Au-butanethiol surface systems were equilibrated in
290     the canonical (NVT) ensemble, organic solvent molecules were packed in
291     the previously empty part of the simulation cells.\cite{packmol} Two
292     solvents were investigated, one which has little vibrational overlap
293     with the alkanethiol and which has a planar shape (toluene), and one
294     which has similar vibrational frequencies to the capping agent and
295     chain-like shape ({\it n}-hexane).
296    
297     The simulation cells were not particularly extensive along the
298     $z$-axis, as a very long length scale for the thermal gradient may
299     cause excessively hot or cold temperatures in the middle of the
300     solvent region and lead to undesired phenomena such as solvent boiling
301     or freezing when a thermal flux is applied. Conversely, too few
302     solvent molecules would change the normal behavior of the liquid
303     phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
304     these extreme cases did not happen to our simulations. The spacing
305     between periodic images of the gold interfaces is $45 \sim 75$\AA in
306     our simulations.
307    
308     The initial configurations generated are further equilibrated with the
309     $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
310     change. This is to ensure that the equilibration of liquid phase does
311     not affect the metal's crystalline structure. Comparisons were made
312     with simulations that allowed changes of $L_x$ and $L_y$ during NPT
313     equilibration. No substantial changes in the box geometry were noticed
314     in these simulations. After ensuring the liquid phase reaches
315     equilibrium at atmospheric pressure (1 atm), further equilibration was
316     carried out under canonical (NVT) and microcanonical (NVE) ensembles.
317    
318     After the systems reach equilibrium, NIVS was used to impose an
319     unphysical thermal flux between the metal and the liquid phases. Most
320     of our simulations were done under an average temperature of
321     $\sim$200K. Therefore, thermal flux usually came from the metal to the
322     liquid so that the liquid has a higher temperature and would not
323     freeze due to lowered temperatures. After this induced temperature
324     gradient had stabilized, the temperature profile of the simulation cell
325     was recorded. To do this, the simulation cell is divided evenly into
326     $N$ slabs along the $z$-axis. The average temperatures of each slab
327     are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
328     the same, the derivatives of $T$ with respect to slab number $n$ can
329     be directly used for $G^\prime$ calculations: \begin{equation}
330     G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
331     \Big/\left(\frac{\partial T}{\partial z}\right)^2
332     = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
333     \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
334     = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
335     \Big/\left(\frac{\partial T}{\partial n}\right)^2
336     \label{derivativeG2}
337     \end{equation}
338     The absolute values in Eq. \ref{derivativeG2} appear because the
339     direction of the flux $\vec{J}$ is in an opposing direction on either
340     side of the metal slab.
341    
342     All of the above simulation procedures use a time step of 1 fs. Each
343     equilibration stage took a minimum of 100 ps, although in some cases,
344     longer equilibration stages were utilized.
345    
346     \subsection{Force Field Parameters}
347     Our simulations include a number of chemically distinct components.
348     Figure \ref{demoMol} demonstrates the sites defined for both
349     United-Atom and All-Atom models of the organic solvent and capping
350     agents in our simulations. Force field parameters are needed for
351     interactions both between the same type of particles and between
352     particles of different species.
353    
354     \begin{figure}
355     \includegraphics[width=\linewidth]{structures}
356     \caption{Structures of the capping agent and solvents utilized in
357     these simulations. The chemically-distinct sites (a-e) are expanded
358     in terms of constituent atoms for both United Atom (UA) and All Atom
359     (AA) force fields. Most parameters are from References
360     \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
361     (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
362     atoms are given in Table 1 in the supporting information.}
363     \label{demoMol}
364     \end{figure}
365    
366     The Au-Au interactions in metal lattice slab is described by the
367     quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
368     potentials include zero-point quantum corrections and are
369     reparametrized for accurate surface energies compared to the
370     Sutton-Chen potentials.\cite{Chen90}
371    
372     For the two solvent molecules, {\it n}-hexane and toluene, two
373     different atomistic models were utilized. Both solvents were modeled
374     using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
375     parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
376     for our UA solvent molecules. In these models, sites are located at
377     the carbon centers for alkyl groups. Bonding interactions, including
378     bond stretches and bends and torsions, were used for intra-molecular
379     sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
380     potentials are used.
381    
382     By eliminating explicit hydrogen atoms, the TraPPE-UA models are
383     simple and computationally efficient, while maintaining good accuracy.
384     However, the TraPPE-UA model for alkanes is known to predict a slightly
385     lower boiling point than experimental values. This is one of the
386     reasons we used a lower average temperature (200K) for our
387     simulations. If heat is transferred to the liquid phase during the
388     NIVS simulation, the liquid in the hot slab can actually be
389     substantially warmer than the mean temperature in the simulation. The
390     lower mean temperatures therefore prevent solvent boiling.
391    
392     For UA-toluene, the non-bonded potentials between intermolecular sites
393     have a similar Lennard-Jones formulation. The toluene molecules were
394     treated as a single rigid body, so there was no need for
395     intramolecular interactions (including bonds, bends, or torsions) in
396     this solvent model.
397    
398     Besides the TraPPE-UA models, AA models for both organic solvents are
399     included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
400     were used. For hexane, additional explicit hydrogen sites were
401     included. Besides bonding and non-bonded site-site interactions,
402     partial charges and the electrostatic interactions were added to each
403     CT and HC site. For toluene, a flexible model for the toluene molecule
404     was utilized which included bond, bend, torsion, and inversion
405     potentials to enforce ring planarity.
406    
407     The butanethiol capping agent in our simulations, were also modeled
408     with both UA and AA model. The TraPPE-UA force field includes
409     parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
410     UA butanethiol model in our simulations. The OPLS-AA also provides
411     parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
412     surfaces do not have the hydrogen atom bonded to sulfur. To derive
413     suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
414     adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
415     modify the parameters for the CTS atom to maintain charge neutrality
416     in the molecule. Note that the model choice (UA or AA) for the capping
417     agent can be different from the solvent. Regardless of model choice,
418     the force field parameters for interactions between capping agent and
419     solvent can be derived using Lorentz-Berthelot Mixing Rule:
420     \begin{eqnarray}
421     \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
422     \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
423     \end{eqnarray}
424    
425     To describe the interactions between metal (Au) and non-metal atoms,
426     we refer to an adsorption study of alkyl thiols on gold surfaces by
427     Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
428     Lennard-Jones form of potential parameters for the interaction between
429     Au and pseudo-atoms CH$_x$ and S based on a well-established and
430     widely-used effective potential of Hautman and Klein for the Au(111)
431     surface.\cite{hautman:4994} As our simulations require the gold slab
432     to be flexible to accommodate thermal excitation, the pair-wise form
433     of potentials they developed was used for our study.
434    
435     The potentials developed from {\it ab initio} calculations by Leng
436     {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
437     interactions between Au and aromatic C/H atoms in toluene. However,
438     the Lennard-Jones parameters between Au and other types of particles,
439     (e.g. AA alkanes) have not yet been established. For these
440     interactions, the Lorentz-Berthelot mixing rule can be used to derive
441     effective single-atom LJ parameters for the metal using the fit values
442     for toluene. These are then used to construct reasonable mixing
443     parameters for the interactions between the gold and other atoms.
444     Table 1 in the supporting information summarizes the
445     ``metal/non-metal'' parameters utilized in our simulations.
446    
447     \section{Results}
448 skuang 3773 [L-J COMPARED TO RNEMD NIVS; WATER COMPARED TO RNEMD NIVS AND EMD;
449 skuang 3770 SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
450    
451 gezelter 3769 There are many factors contributing to the measured interfacial
452     conductance; some of these factors are physically motivated
453     (e.g. coverage of the surface by the capping agent coverage and
454     solvent identity), while some are governed by parameters of the
455     methodology (e.g. applied flux and the formulas used to obtain the
456     conductance). In this section we discuss the major physical and
457     calculational effects on the computed conductivity.
458    
459     \subsection{Effects due to capping agent coverage}
460    
461     A series of different initial conditions with a range of surface
462     coverages was prepared and solvated with various with both of the
463     solvent molecules. These systems were then equilibrated and their
464     interfacial thermal conductivity was measured with the NIVS
465     algorithm. Figure \ref{coverage} demonstrates the trend of conductance
466     with respect to surface coverage.
467    
468     \begin{figure}
469     \includegraphics[width=\linewidth]{coverage}
470     \caption{The interfacial thermal conductivity ($G$) has a
471     non-monotonic dependence on the degree of surface capping. This
472     data is for the Au(111) / butanethiol / solvent interface with
473     various UA force fields at $\langle T\rangle \sim $200K.}
474     \label{coverage}
475     \end{figure}
476    
477     In partially covered surfaces, the derivative definition for
478     $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
479     location of maximum change of $\lambda$ becomes washed out. The
480     discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
481     Gibbs dividing surface is still well-defined. Therefore, $G$ (not
482     $G^\prime$) was used in this section.
483    
484     From Figure \ref{coverage}, one can see the significance of the
485     presence of capping agents. When even a small fraction of the Au(111)
486     surface sites are covered with butanethiols, the conductivity exhibits
487     an enhancement by at least a factor of 3. Capping agents are clearly
488     playing a major role in thermal transport at metal / organic solvent
489     surfaces.
490    
491     We note a non-monotonic behavior in the interfacial conductance as a
492     function of surface coverage. The maximum conductance (largest $G$)
493     happens when the surfaces are about 75\% covered with butanethiol
494     caps. The reason for this behavior is not entirely clear. One
495     explanation is that incomplete butanethiol coverage allows small gaps
496     between butanethiols to form. These gaps can be filled by transient
497     solvent molecules. These solvent molecules couple very strongly with
498     the hot capping agent molecules near the surface, and can then carry
499     away (diffusively) the excess thermal energy from the surface.
500    
501     There appears to be a competition between the conduction of the
502     thermal energy away from the surface by the capping agents (enhanced
503     by greater coverage) and the coupling of the capping agents with the
504     solvent (enhanced by interdigitation at lower coverages). This
505     competition would lead to the non-monotonic coverage behavior observed
506     here.
507    
508     Results for rigid body toluene solvent, as well as the UA hexane, are
509     within the ranges expected from prior experimental
510     work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
511     that explicit hydrogen atoms might not be required for modeling
512     thermal transport in these systems. C-H vibrational modes do not see
513     significant excited state population at low temperatures, and are not
514     likely to carry lower frequency excitations from the solid layer into
515     the bulk liquid.
516    
517     The toluene solvent does not exhibit the same behavior as hexane in
518     that $G$ remains at approximately the same magnitude when the capping
519     coverage increases from 25\% to 75\%. Toluene, as a rigid planar
520     molecule, cannot occupy the relatively small gaps between the capping
521     agents as easily as the chain-like {\it n}-hexane. The effect of
522     solvent coupling to the capping agent is therefore weaker in toluene
523     except at the very lowest coverage levels. This effect counters the
524     coverage-dependent conduction of heat away from the metal surface,
525     leading to a much flatter $G$ vs. coverage trend than is observed in
526     {\it n}-hexane.
527    
528     \subsection{Effects due to Solvent \& Solvent Models}
529     In addition to UA solvent and capping agent models, AA models have
530     also been included in our simulations. In most of this work, the same
531     (UA or AA) model for solvent and capping agent was used, but it is
532     also possible to utilize different models for different components.
533     We have also included isotopic substitutions (Hydrogen to Deuterium)
534     to decrease the explicit vibrational overlap between solvent and
535     capping agent. Table \ref{modelTest} summarizes the results of these
536     studies.
537    
538     \begin{table*}
539     \begin{minipage}{\linewidth}
540     \begin{center}
541    
542     \caption{Computed interfacial thermal conductance ($G$ and
543     $G^\prime$) values for interfaces using various models for
544     solvent and capping agent (or without capping agent) at
545     $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
546     solvent or capping agent molecules. Error estimates are
547     indicated in parentheses.}
548    
549     \begin{tabular}{llccc}
550     \hline\hline
551     Butanethiol model & Solvent & $G$ & $G^\prime$ \\
552     (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
553     \hline
554     UA & UA hexane & 131(9) & 87(10) \\
555     & UA hexane(D) & 153(5) & 136(13) \\
556     & AA hexane & 131(6) & 122(10) \\
557     & UA toluene & 187(16) & 151(11) \\
558     & AA toluene & 200(36) & 149(53) \\
559     \hline
560     AA & UA hexane & 116(9) & 129(8) \\
561     & AA hexane & 442(14) & 356(31) \\
562     & AA hexane(D) & 222(12) & 234(54) \\
563     & UA toluene & 125(25) & 97(60) \\
564     & AA toluene & 487(56) & 290(42) \\
565     \hline
566     AA(D) & UA hexane & 158(25) & 172(4) \\
567     & AA hexane & 243(29) & 191(11) \\
568     & AA toluene & 364(36) & 322(67) \\
569     \hline
570     bare & UA hexane & 46.5(3.2) & 49.4(4.5) \\
571     & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
572     & AA hexane & 31.0(1.4) & 29.4(1.3) \\
573     & UA toluene & 70.1(1.3) & 65.8(0.5) \\
574     \hline\hline
575     \end{tabular}
576     \label{modelTest}
577     \end{center}
578     \end{minipage}
579     \end{table*}
580    
581     To facilitate direct comparison between force fields, systems with the
582     same capping agent and solvent were prepared with the same length
583     scales for the simulation cells.
584    
585     On bare metal / solvent surfaces, different force field models for
586     hexane yield similar results for both $G$ and $G^\prime$, and these
587     two definitions agree with each other very well. This is primarily an
588     indicator of weak interactions between the metal and the solvent.
589    
590     For the fully-covered surfaces, the choice of force field for the
591     capping agent and solvent has a large impact on the calculated values
592     of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
593     much larger than their UA to UA counterparts, and these values exceed
594     the experimental estimates by a large measure. The AA force field
595     allows significant energy to go into C-H (or C-D) stretching modes,
596     and since these modes are high frequency, this non-quantum behavior is
597     likely responsible for the overestimate of the conductivity. Compared
598     to the AA model, the UA model yields more reasonable conductivity
599     values with much higher computational efficiency.
600    
601     \subsubsection{Are electronic excitations in the metal important?}
602     Because they lack electronic excitations, the QSC and related embedded
603     atom method (EAM) models for gold are known to predict unreasonably
604     low values for bulk conductivity
605     ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
606     conductance between the phases ($G$) is governed primarily by phonon
607     excitation (and not electronic degrees of freedom), one would expect a
608     classical model to capture most of the interfacial thermal
609     conductance. Our results for $G$ and $G^\prime$ indicate that this is
610     indeed the case, and suggest that the modeling of interfacial thermal
611     transport depends primarily on the description of the interactions
612     between the various components at the interface. When the metal is
613     chemically capped, the primary barrier to thermal conductivity appears
614     to be the interface between the capping agent and the surrounding
615     solvent, so the excitations in the metal have little impact on the
616     value of $G$.
617    
618     \subsection{Effects due to methodology and simulation parameters}
619    
620     We have varied the parameters of the simulations in order to
621     investigate how these factors would affect the computation of $G$. Of
622     particular interest are: 1) the length scale for the applied thermal
623     gradient (modified by increasing the amount of solvent in the system),
624     2) the sign and magnitude of the applied thermal flux, 3) the average
625     temperature of the simulation (which alters the solvent density during
626     equilibration), and 4) the definition of the interfacial conductance
627     (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
628     calculation.
629    
630     Systems of different lengths were prepared by altering the number of
631     solvent molecules and extending the length of the box along the $z$
632     axis to accomodate the extra solvent. Equilibration at the same
633     temperature and pressure conditions led to nearly identical surface
634     areas ($L_x$ and $L_y$) available to the metal and capping agent,
635     while the extra solvent served mainly to lengthen the axis that was
636     used to apply the thermal flux. For a given value of the applied
637     flux, the different $z$ length scale has only a weak effect on the
638     computed conductivities.
639    
640     \subsubsection{Effects of applied flux}
641     The NIVS algorithm allows changes in both the sign and magnitude of
642     the applied flux. It is possible to reverse the direction of heat
643     flow simply by changing the sign of the flux, and thermal gradients
644     which would be difficult to obtain experimentally ($5$ K/\AA) can be
645     easily simulated. However, the magnitude of the applied flux is not
646     arbitrary if one aims to obtain a stable and reliable thermal gradient.
647     A temperature gradient can be lost in the noise if $|J_z|$ is too
648     small, and excessive $|J_z|$ values can cause phase transitions if the
649     extremes of the simulation cell become widely separated in
650     temperature. Also, if $|J_z|$ is too large for the bulk conductivity
651     of the materials, the thermal gradient will never reach a stable
652     state.
653    
654     Within a reasonable range of $J_z$ values, we were able to study how
655     $G$ changes as a function of this flux. In what follows, we use
656     positive $J_z$ values to denote the case where energy is being
657     transferred by the method from the metal phase and into the liquid.
658     The resulting gradient therefore has a higher temperature in the
659     liquid phase. Negative flux values reverse this transfer, and result
660     in higher temperature metal phases. The conductance measured under
661     different applied $J_z$ values is listed in Tables 2 and 3 in the
662     supporting information. These results do not indicate that $G$ depends
663     strongly on $J_z$ within this flux range. The linear response of flux
664     to thermal gradient simplifies our investigations in that we can rely
665     on $G$ measurement with only a small number $J_z$ values.
666    
667     The sign of $J_z$ is a different matter, however, as this can alter
668     the temperature on the two sides of the interface. The average
669     temperature values reported are for the entire system, and not for the
670     liquid phase, so at a given $\langle T \rangle$, the system with
671     positive $J_z$ has a warmer liquid phase. This means that if the
672     liquid carries thermal energy via diffusive transport, {\it positive}
673     $J_z$ values will result in increased molecular motion on the liquid
674     side of the interface, and this will increase the measured
675     conductivity.
676    
677     \subsubsection{Effects due to average temperature}
678    
679     We also studied the effect of average system temperature on the
680     interfacial conductance. The simulations are first equilibrated in
681     the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
682     predict a lower boiling point (and liquid state density) than
683     experiments. This lower-density liquid phase leads to reduced contact
684     between the hexane and butanethiol, and this accounts for our
685     observation of lower conductance at higher temperatures. In raising
686     the average temperature from 200K to 250K, the density drop of
687     $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
688     conductance.
689    
690     Similar behavior is observed in the TraPPE-UA model for toluene,
691     although this model has better agreement with the experimental
692     densities of toluene. The expansion of the toluene liquid phase is
693     not as significant as that of the hexane (8.3\% over 100K), and this
694     limits the effect to $\sim$20\% drop in thermal conductivity.
695    
696     Although we have not mapped out the behavior at a large number of
697     temperatures, is clear that there will be a strong temperature
698     dependence in the interfacial conductance when the physical properties
699     of one side of the interface (notably the density) change rapidly as a
700     function of temperature.
701    
702     Besides the lower interfacial thermal conductance, surfaces at
703     relatively high temperatures are susceptible to reconstructions,
704     particularly when butanethiols fully cover the Au(111) surface. These
705     reconstructions include surface Au atoms which migrate outward to the
706     S atom layer, and butanethiol molecules which embed into the surface
707     Au layer. The driving force for this behavior is the strong Au-S
708     interactions which are modeled here with a deep Lennard-Jones
709     potential. This phenomenon agrees with reconstructions that have been
710     experimentally
711     observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
712     {\it et al.} kept their Au(111) slab rigid so that their simulations
713     could reach 300K without surface
714     reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
715     blur the interface, the measurement of $G$ becomes more difficult to
716     conduct at higher temperatures. For this reason, most of our
717     measurements are undertaken at $\langle T\rangle\sim$200K where
718     reconstruction is minimized.
719    
720     However, when the surface is not completely covered by butanethiols,
721     the simulated system appears to be more resistent to the
722     reconstruction. Our Au / butanethiol / toluene system had the Au(111)
723     surfaces 90\% covered by butanethiols, but did not see this above
724     phenomena even at $\langle T\rangle\sim$300K. That said, we did
725     observe butanethiols migrating to neighboring three-fold sites during
726     a simulation. Since the interface persisted in these simulations, we
727     were able to obtain $G$'s for these interfaces even at a relatively
728     high temperature without being affected by surface reconstructions.
729    
730     \section{Discussion}
731 skuang 3770 [COMBINE W. RESULTS]
732 gezelter 3769 The primary result of this work is that the capping agent acts as an
733     efficient thermal coupler between solid and solvent phases. One of
734     the ways the capping agent can carry out this role is to down-shift
735     between the phonon vibrations in the solid (which carry the heat from
736     the gold) and the molecular vibrations in the liquid (which carry some
737     of the heat in the solvent).
738    
739     To investigate the mechanism of interfacial thermal conductance, the
740     vibrational power spectrum was computed. Power spectra were taken for
741     individual components in different simulations. To obtain these
742     spectra, simulations were run after equilibration in the
743     microcanonical (NVE) ensemble and without a thermal
744     gradient. Snapshots of configurations were collected at a frequency
745     that is higher than that of the fastest vibrations occurring in the
746     simulations. With these configurations, the velocity auto-correlation
747     functions can be computed:
748     \begin{equation}
749     C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
750     \label{vCorr}
751     \end{equation}
752     The power spectrum is constructed via a Fourier transform of the
753     symmetrized velocity autocorrelation function,
754     \begin{equation}
755     \hat{f}(\omega) =
756     \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
757     \label{fourier}
758     \end{equation}
759    
760     \subsection{The role of specific vibrations}
761     The vibrational spectra for gold slabs in different environments are
762     shown as in Figure \ref{specAu}. Regardless of the presence of
763     solvent, the gold surfaces which are covered by butanethiol molecules
764     exhibit an additional peak observed at a frequency of
765     $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
766     vibration. This vibration enables efficient thermal coupling of the
767     surface Au layer to the capping agents. Therefore, in our simulations,
768     the Au / S interfaces do not appear to be the primary barrier to
769     thermal transport when compared with the butanethiol / solvent
770     interfaces. This supports the results of Luo {\it et
771     al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
772     twice as large as what we have computed for the thiol-liquid
773     interfaces.
774    
775     \begin{figure}
776     \includegraphics[width=\linewidth]{vibration}
777     \caption{The vibrational power spectrum for thiol-capped gold has an
778     additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
779     surfaces (both with and without a solvent over-layer) are missing
780     this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
781     the vibrational power spectrum for the butanethiol capping agents.}
782     \label{specAu}
783     \end{figure}
784    
785     Also in this figure, we show the vibrational power spectrum for the
786     bound butanethiol molecules, which also exhibits the same
787     $\sim$165cm$^{-1}$ peak.
788    
789     \subsection{Overlap of power spectra}
790     A comparison of the results obtained from the two different organic
791     solvents can also provide useful information of the interfacial
792     thermal transport process. In particular, the vibrational overlap
793     between the butanethiol and the organic solvents suggests a highly
794     efficient thermal exchange between these components. Very high
795     thermal conductivity was observed when AA models were used and C-H
796     vibrations were treated classically. The presence of extra degrees of
797     freedom in the AA force field yields higher heat exchange rates
798     between the two phases and results in a much higher conductivity than
799     in the UA force field. The all-atom classical models include high
800     frequency modes which should be unpopulated at our relatively low
801     temperatures. This artifact is likely the cause of the high thermal
802     conductance in all-atom MD simulations.
803    
804     The similarity in the vibrational modes available to solvent and
805     capping agent can be reduced by deuterating one of the two components
806     (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
807     are deuterated, one can observe a significantly lower $G$ and
808     $G^\prime$ values (Table \ref{modelTest}).
809    
810     \begin{figure}
811     \includegraphics[width=\linewidth]{aahxntln}
812     \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
813     systems. When butanethiol is deuterated (lower left), its
814     vibrational overlap with hexane decreases significantly. Since
815     aromatic molecules and the butanethiol are vibrationally dissimilar,
816     the change is not as dramatic when toluene is the solvent (right).}
817     \label{aahxntln}
818     \end{figure}
819    
820     For the Au / butanethiol / toluene interfaces, having the AA
821     butanethiol deuterated did not yield a significant change in the
822     measured conductance. Compared to the C-H vibrational overlap between
823     hexane and butanethiol, both of which have alkyl chains, the overlap
824     between toluene and butanethiol is not as significant and thus does
825     not contribute as much to the heat exchange process.
826    
827     Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
828     that the {\it intra}molecular heat transport due to alkylthiols is
829     highly efficient. Combining our observations with those of Zhang {\it
830     et al.}, it appears that butanethiol acts as a channel to expedite
831     heat flow from the gold surface and into the alkyl chain. The
832     vibrational coupling between the metal and the liquid phase can
833     therefore be enhanced with the presence of suitable capping agents.
834    
835     Deuterated models in the UA force field did not decouple the thermal
836     transport as well as in the AA force field. The UA models, even
837     though they have eliminated the high frequency C-H vibrational
838     overlap, still have significant overlap in the lower-frequency
839     portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
840     the UA models did not decouple the low frequency region enough to
841     produce an observable difference for the results of $G$ (Table
842     \ref{modelTest}).
843    
844     \begin{figure}
845     \includegraphics[width=\linewidth]{uahxnua}
846     \caption{Vibrational power spectra for UA models for the butanethiol
847     and hexane solvent (upper panel) show the high degree of overlap
848     between these two molecules, particularly at lower frequencies.
849     Deuterating a UA model for the solvent (lower panel) does not
850     decouple the two spectra to the same degree as in the AA force
851     field (see Fig \ref{aahxntln}).}
852     \label{uahxnua}
853     \end{figure}
854    
855     \section{Conclusions}
856     The NIVS algorithm has been applied to simulations of
857     butanethiol-capped Au(111) surfaces in the presence of organic
858     solvents. This algorithm allows the application of unphysical thermal
859     flux to transfer heat between the metal and the liquid phase. With the
860     flux applied, we were able to measure the corresponding thermal
861     gradients and to obtain interfacial thermal conductivities. Under
862     steady states, 2-3 ns trajectory simulations are sufficient for
863     computation of this quantity.
864    
865     Our simulations have seen significant conductance enhancement in the
866     presence of capping agent, compared with the bare gold / liquid
867     interfaces. The vibrational coupling between the metal and the liquid
868     phase is enhanced by a chemically-bonded capping agent. Furthermore,
869     the coverage percentage of the capping agent plays an important role
870     in the interfacial thermal transport process. Moderately low coverages
871     allow higher contact between capping agent and solvent, and thus could
872     further enhance the heat transfer process, giving a non-monotonic
873     behavior of conductance with increasing coverage.
874    
875     Our results, particularly using the UA models, agree well with
876     available experimental data. The AA models tend to overestimate the
877     interfacial thermal conductance in that the classically treated C-H
878     vibrations become too easily populated. Compared to the AA models, the
879     UA models have higher computational efficiency with satisfactory
880     accuracy, and thus are preferable in modeling interfacial thermal
881     transport.
882    
883     Of the two definitions for $G$, the discrete form
884     (Eq. \ref{discreteG}) was easier to use and gives out relatively
885     consistent results, while the derivative form (Eq. \ref{derivativeG})
886     is not as versatile. Although $G^\prime$ gives out comparable results
887     and follows similar trend with $G$ when measuring close to fully
888     covered or bare surfaces, the spatial resolution of $T$ profile
889     required for the use of a derivative form is limited by the number of
890     bins and the sampling required to obtain thermal gradient information.
891    
892     Vlugt {\it et al.} have investigated the surface thiol structures for
893     nanocrystalline gold and pointed out that they differ from those of
894     the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
895     difference could also cause differences in the interfacial thermal
896     transport behavior. To investigate this problem, one would need an
897     effective method for applying thermal gradients in non-planar
898     (i.e. spherical) geometries.
899    
900     \section{Acknowledgments}
901     Support for this project was provided by the National Science
902     Foundation under grant CHE-0848243. Computational time was provided by
903     the Center for Research Computing (CRC) at the University of Notre
904     Dame.
905    
906     \newpage
907    
908 skuang 3770 \bibliography{stokes}
909 gezelter 3769
910     \end{doublespace}
911     \end{document}
912