ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3775
Committed: Thu Dec 8 01:34:21 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 43488 byte(s)
Log Message:
a rought draft of simulation details and a figure added.

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 skuang 3775 as in reference \cite{kuang:AuThl}.
266 skuang 3773
267 skuang 3774 As for the ice/liquid water interfaces, the basal surface of ice
268 skuang 3775 lattice was first constructed. Hirsch {\it et
269     al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice
270     lattices with different proton orders. We refer to their results and
271     choose the configuration of the lowest energy after geometry
272     optimization as the unit cells of our ice lattices. Although
273 skuang 3774 experimental solid/liquid coexistant temperature near normal pressure
274 skuang 3775 is 273K, Bryk and Haymet's simulations of ice/liquid water interfaces
275     with different models suggest that for SPC/E, the most stable
276     interface is observed at 225$\pm$5K. Therefore, all our ice/liquid
277     water simulations were carried out under 225K. To have extra
278     protection of the ice lattice during initial equilibration (when the
279     randomly generated liquid phase configuration could release large
280     amount of energy in relaxation), a constraint method (REF?) was
281     adopted until the high energy configuration was relaxed.
282     [MAY ADD A FIGURE HERE FOR BASAL PLANE, MAY INCLUDE PRISM IF POSSIBLE]
283 gezelter 3769
284     \subsection{Force Field Parameters}
285 skuang 3774 For comparison of our new method with previous work, we retain our
286     force field parameters consistent with the results we will compare
287 skuang 3775 with. The Lennard-Jones fluid used here for argon , and reduced unit
288     results are reported for direct comparison purpose.
289 gezelter 3769
290 skuang 3774 As for our water simulations, SPC/E model is used throughout this work
291     for consistency. Previous work for transport properties of SPC/E water
292 skuang 3775 model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
293     that unnecessary repetition of previous methods can be avoided.
294 gezelter 3769
295 skuang 3774 The Au-Au interaction parameters in all simulations are described by
296     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
297     QSC potentials include zero-point quantum corrections and are
298 gezelter 3769 reparametrized for accurate surface energies compared to the
299 skuang 3775 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
300     Spohr potential was adopted\cite{ISI:000167766600035} to depict
301     Au-H$_2$O interactions.
302 gezelter 3769
303 skuang 3774 The small organic molecules included in our simulations are the Au
304     surface capping agent butanethiol and liquid hexane and toluene. The
305     United-Atom
306     models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
307     for these components were used in this work for better computational
308     efficiency, while maintaining good accuracy. We refer readers to our
309     previous work\cite{kuang:AuThl} for further details of these models,
310     as well as the interactions between Au and the above organic molecule
311     components.
312 gezelter 3769
313 skuang 3774 \subsection{Thermal conductivities}
314 skuang 3775 When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
315     impose kinetic energy transfer, the method can be used for thermal
316     conductivity computations. Similar to previous RNEMD methods, we
317     assume linear response of the temperature gradient with respect to the
318     thermal flux in general case. And the thermal conductivity ($\lambda$)
319     can be obtained with the imposed kinetic energy flux and the measured
320     thermal gradient:
321     \begin{equation}
322     J_z = -\lambda \frac{\partial T}{\partial z}
323     \end{equation}
324     Like other imposed-flux methods, the energy flux was calculated using
325     the total non-physical energy transferred (${E_{total}}$) from slab
326     ``c'' to slab ``h'', which is recorded throughout a simulation, and
327     the time for data collection $t$:
328     \begin{equation}
329     J_z = \frac{E_{total}}{2 t L_x L_y}
330     \end{equation}
331     where $L_x$ and $L_y$ denotes the dimensions of the plane in a
332     simulation cell perpendicular to the thermal gradient, and a factor of
333     two in the denominator is present for the heat transport occurs in
334     both $+z$ and $-z$ directions. The temperature gradient
335     ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
336     regression of the temperature profile, which is recorded during a
337     simulation for each slab in a cell. For Lennard-Jones simulations,
338     thermal conductivities are reported in reduced units
339     (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
340    
341 skuang 3774 \subsection{Shear viscosities}
342 skuang 3775 Alternatively, the method can carry out shear viscosity calculations
343     by switching off $J_z$. One can specify the vector
344     $\vec{j}_z(\vec{p})$ by choosing the three components
345     respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
346     set to zero. Although for isotropic systems, the direction of
347     $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, the ability
348     of arbitarily specifying the vector direction in our method provides
349     convenience in anisotropic simulations.
350    
351     Similar to thermal conductivity computations, linear response of the
352     momentum gradient with respect to the shear stress is assumed, and the
353     shear viscosity ($\eta$) can be obtained with the imposed momentum
354     flux (e.g. in $x$ direction) and the measured gradient:
355     \begin{equation}
356     j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
357     \end{equation}
358     where the flux is similarly defined:
359     \begin{equation}
360     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
361     \end{equation}
362     with $P_x$ being the total non-physical momentum transferred within
363     the data collection time. Also, the velocity gradient
364     ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
365     regression of the $x$ component of the mean velocity, $\langle
366     v_x\rangle$, in each of the bins. For Lennard-Jones simulations, shear
367     viscosities are reported in reduced units
368     (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
369    
370 skuang 3774 \subsection{Interfacial friction and Slip length}
371 skuang 3775 While the shear stress results in a velocity gradient within bulk
372     fluid phase, its effect at a solid-liquid interface could vary due to
373     the interaction strength between the two phases. The interfacial
374     friction coefficient $\kappa$ is defined to relate the shear stress
375     (e.g. along $x$-axis) and the relative fluid velocity tangent to the
376     interface:
377     \begin{equation}
378     j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
379     \end{equation}
380     Under ``stick'' boundary condition, $\Delta v_x|_{interface}
381     \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
382     ``slip'' boundary condition at the solid-liquid interface, $\kappa$
383     becomes finite. To characterize the interfacial boundary conditions,
384     slip length ($\delta$) is defined using $\kappa$ and the shear
385     viscocity of liquid phase ($\eta$):
386     \begin{equation}
387     \delta = \frac{\eta}{\kappa}
388     \end{equation}
389     so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
390     and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
391     illustrates how this quantity is defined and computed for a
392     solid-liquid interface.
393 gezelter 3769
394 skuang 3775 \begin{figure}
395     \includegraphics[width=\linewidth]{defDelta}
396     \caption{The slip length $\delta$ can be obtained from a velocity
397     profile of a solid-liquid interface. An example of Au/hexane
398     interfaces is shown.}
399     \label{slipLength}
400     \end{figure}
401 gezelter 3769
402 skuang 3775 In our method, a shear stress can be applied similar to shear
403     viscosity computations by applying an unphysical momentum flux
404     (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
405     shown in Figure \ref{slipLength}, in which the velocity gradients
406     within liquid phase and velocity difference at the liquid-solid
407     interface can be measured respectively. Further calculations and
408     characterizations of the interface can be carried out using these
409     data.
410     [MENTION IN RESULTS THAT ETA OBTAINED HERE DOES NOT NECESSARILY EQUAL
411     TO BULK VALUES]
412    
413 gezelter 3769 \section{Results}
414 skuang 3773 [L-J COMPARED TO RNEMD NIVS; WATER COMPARED TO RNEMD NIVS AND EMD;
415 skuang 3770 SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
416    
417 gezelter 3769 There are many factors contributing to the measured interfacial
418     conductance; some of these factors are physically motivated
419     (e.g. coverage of the surface by the capping agent coverage and
420     solvent identity), while some are governed by parameters of the
421     methodology (e.g. applied flux and the formulas used to obtain the
422     conductance). In this section we discuss the major physical and
423     calculational effects on the computed conductivity.
424    
425     \subsection{Effects due to capping agent coverage}
426    
427     A series of different initial conditions with a range of surface
428     coverages was prepared and solvated with various with both of the
429     solvent molecules. These systems were then equilibrated and their
430     interfacial thermal conductivity was measured with the NIVS
431     algorithm. Figure \ref{coverage} demonstrates the trend of conductance
432     with respect to surface coverage.
433    
434     \begin{figure}
435     \includegraphics[width=\linewidth]{coverage}
436     \caption{The interfacial thermal conductivity ($G$) has a
437     non-monotonic dependence on the degree of surface capping. This
438     data is for the Au(111) / butanethiol / solvent interface with
439     various UA force fields at $\langle T\rangle \sim $200K.}
440     \label{coverage}
441     \end{figure}
442    
443     In partially covered surfaces, the derivative definition for
444     $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
445     location of maximum change of $\lambda$ becomes washed out. The
446     discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
447     Gibbs dividing surface is still well-defined. Therefore, $G$ (not
448     $G^\prime$) was used in this section.
449    
450     From Figure \ref{coverage}, one can see the significance of the
451     presence of capping agents. When even a small fraction of the Au(111)
452     surface sites are covered with butanethiols, the conductivity exhibits
453     an enhancement by at least a factor of 3. Capping agents are clearly
454     playing a major role in thermal transport at metal / organic solvent
455     surfaces.
456    
457     We note a non-monotonic behavior in the interfacial conductance as a
458     function of surface coverage. The maximum conductance (largest $G$)
459     happens when the surfaces are about 75\% covered with butanethiol
460     caps. The reason for this behavior is not entirely clear. One
461     explanation is that incomplete butanethiol coverage allows small gaps
462     between butanethiols to form. These gaps can be filled by transient
463     solvent molecules. These solvent molecules couple very strongly with
464     the hot capping agent molecules near the surface, and can then carry
465     away (diffusively) the excess thermal energy from the surface.
466    
467     There appears to be a competition between the conduction of the
468     thermal energy away from the surface by the capping agents (enhanced
469     by greater coverage) and the coupling of the capping agents with the
470     solvent (enhanced by interdigitation at lower coverages). This
471     competition would lead to the non-monotonic coverage behavior observed
472     here.
473    
474     Results for rigid body toluene solvent, as well as the UA hexane, are
475     within the ranges expected from prior experimental
476     work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
477     that explicit hydrogen atoms might not be required for modeling
478     thermal transport in these systems. C-H vibrational modes do not see
479     significant excited state population at low temperatures, and are not
480     likely to carry lower frequency excitations from the solid layer into
481     the bulk liquid.
482    
483     The toluene solvent does not exhibit the same behavior as hexane in
484     that $G$ remains at approximately the same magnitude when the capping
485     coverage increases from 25\% to 75\%. Toluene, as a rigid planar
486     molecule, cannot occupy the relatively small gaps between the capping
487     agents as easily as the chain-like {\it n}-hexane. The effect of
488     solvent coupling to the capping agent is therefore weaker in toluene
489     except at the very lowest coverage levels. This effect counters the
490     coverage-dependent conduction of heat away from the metal surface,
491     leading to a much flatter $G$ vs. coverage trend than is observed in
492     {\it n}-hexane.
493    
494     \subsection{Effects due to Solvent \& Solvent Models}
495     In addition to UA solvent and capping agent models, AA models have
496     also been included in our simulations. In most of this work, the same
497     (UA or AA) model for solvent and capping agent was used, but it is
498     also possible to utilize different models for different components.
499     We have also included isotopic substitutions (Hydrogen to Deuterium)
500     to decrease the explicit vibrational overlap between solvent and
501     capping agent. Table \ref{modelTest} summarizes the results of these
502     studies.
503    
504     \begin{table*}
505     \begin{minipage}{\linewidth}
506     \begin{center}
507    
508     \caption{Computed interfacial thermal conductance ($G$ and
509     $G^\prime$) values for interfaces using various models for
510     solvent and capping agent (or without capping agent) at
511     $\langle T\rangle\sim$200K. Here ``D'' stands for deuterated
512     solvent or capping agent molecules. Error estimates are
513     indicated in parentheses.}
514    
515     \begin{tabular}{llccc}
516     \hline\hline
517     Butanethiol model & Solvent & $G$ & $G^\prime$ \\
518     (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
519     \hline
520     UA & UA hexane & 131(9) & 87(10) \\
521     & UA hexane(D) & 153(5) & 136(13) \\
522     & AA hexane & 131(6) & 122(10) \\
523     & UA toluene & 187(16) & 151(11) \\
524     & AA toluene & 200(36) & 149(53) \\
525     \hline
526     AA & UA hexane & 116(9) & 129(8) \\
527     & AA hexane & 442(14) & 356(31) \\
528     & AA hexane(D) & 222(12) & 234(54) \\
529     & UA toluene & 125(25) & 97(60) \\
530     & AA toluene & 487(56) & 290(42) \\
531     \hline
532     AA(D) & UA hexane & 158(25) & 172(4) \\
533     & AA hexane & 243(29) & 191(11) \\
534     & AA toluene & 364(36) & 322(67) \\
535     \hline
536     bare & UA hexane & 46.5(3.2) & 49.4(4.5) \\
537     & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
538     & AA hexane & 31.0(1.4) & 29.4(1.3) \\
539     & UA toluene & 70.1(1.3) & 65.8(0.5) \\
540     \hline\hline
541     \end{tabular}
542     \label{modelTest}
543     \end{center}
544     \end{minipage}
545     \end{table*}
546    
547     To facilitate direct comparison between force fields, systems with the
548     same capping agent and solvent were prepared with the same length
549     scales for the simulation cells.
550    
551     On bare metal / solvent surfaces, different force field models for
552     hexane yield similar results for both $G$ and $G^\prime$, and these
553     two definitions agree with each other very well. This is primarily an
554     indicator of weak interactions between the metal and the solvent.
555    
556     For the fully-covered surfaces, the choice of force field for the
557     capping agent and solvent has a large impact on the calculated values
558     of $G$ and $G^\prime$. The AA thiol to AA solvent conductivities are
559     much larger than their UA to UA counterparts, and these values exceed
560     the experimental estimates by a large measure. The AA force field
561     allows significant energy to go into C-H (or C-D) stretching modes,
562     and since these modes are high frequency, this non-quantum behavior is
563     likely responsible for the overestimate of the conductivity. Compared
564     to the AA model, the UA model yields more reasonable conductivity
565     values with much higher computational efficiency.
566    
567     \subsubsection{Are electronic excitations in the metal important?}
568     Because they lack electronic excitations, the QSC and related embedded
569     atom method (EAM) models for gold are known to predict unreasonably
570     low values for bulk conductivity
571     ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
572     conductance between the phases ($G$) is governed primarily by phonon
573     excitation (and not electronic degrees of freedom), one would expect a
574     classical model to capture most of the interfacial thermal
575     conductance. Our results for $G$ and $G^\prime$ indicate that this is
576     indeed the case, and suggest that the modeling of interfacial thermal
577     transport depends primarily on the description of the interactions
578     between the various components at the interface. When the metal is
579     chemically capped, the primary barrier to thermal conductivity appears
580     to be the interface between the capping agent and the surrounding
581     solvent, so the excitations in the metal have little impact on the
582     value of $G$.
583    
584     \subsection{Effects due to methodology and simulation parameters}
585    
586     We have varied the parameters of the simulations in order to
587     investigate how these factors would affect the computation of $G$. Of
588     particular interest are: 1) the length scale for the applied thermal
589     gradient (modified by increasing the amount of solvent in the system),
590     2) the sign and magnitude of the applied thermal flux, 3) the average
591     temperature of the simulation (which alters the solvent density during
592     equilibration), and 4) the definition of the interfacial conductance
593     (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
594     calculation.
595    
596     Systems of different lengths were prepared by altering the number of
597     solvent molecules and extending the length of the box along the $z$
598     axis to accomodate the extra solvent. Equilibration at the same
599     temperature and pressure conditions led to nearly identical surface
600     areas ($L_x$ and $L_y$) available to the metal and capping agent,
601     while the extra solvent served mainly to lengthen the axis that was
602     used to apply the thermal flux. For a given value of the applied
603     flux, the different $z$ length scale has only a weak effect on the
604     computed conductivities.
605    
606     \subsubsection{Effects of applied flux}
607     The NIVS algorithm allows changes in both the sign and magnitude of
608     the applied flux. It is possible to reverse the direction of heat
609     flow simply by changing the sign of the flux, and thermal gradients
610     which would be difficult to obtain experimentally ($5$ K/\AA) can be
611     easily simulated. However, the magnitude of the applied flux is not
612     arbitrary if one aims to obtain a stable and reliable thermal gradient.
613     A temperature gradient can be lost in the noise if $|J_z|$ is too
614     small, and excessive $|J_z|$ values can cause phase transitions if the
615     extremes of the simulation cell become widely separated in
616     temperature. Also, if $|J_z|$ is too large for the bulk conductivity
617     of the materials, the thermal gradient will never reach a stable
618     state.
619    
620     Within a reasonable range of $J_z$ values, we were able to study how
621     $G$ changes as a function of this flux. In what follows, we use
622     positive $J_z$ values to denote the case where energy is being
623     transferred by the method from the metal phase and into the liquid.
624     The resulting gradient therefore has a higher temperature in the
625     liquid phase. Negative flux values reverse this transfer, and result
626     in higher temperature metal phases. The conductance measured under
627     different applied $J_z$ values is listed in Tables 2 and 3 in the
628     supporting information. These results do not indicate that $G$ depends
629     strongly on $J_z$ within this flux range. The linear response of flux
630     to thermal gradient simplifies our investigations in that we can rely
631     on $G$ measurement with only a small number $J_z$ values.
632    
633     The sign of $J_z$ is a different matter, however, as this can alter
634     the temperature on the two sides of the interface. The average
635     temperature values reported are for the entire system, and not for the
636     liquid phase, so at a given $\langle T \rangle$, the system with
637     positive $J_z$ has a warmer liquid phase. This means that if the
638     liquid carries thermal energy via diffusive transport, {\it positive}
639     $J_z$ values will result in increased molecular motion on the liquid
640     side of the interface, and this will increase the measured
641     conductivity.
642    
643     \subsubsection{Effects due to average temperature}
644    
645     We also studied the effect of average system temperature on the
646     interfacial conductance. The simulations are first equilibrated in
647     the NPT ensemble at 1 atm. The TraPPE-UA model for hexane tends to
648     predict a lower boiling point (and liquid state density) than
649     experiments. This lower-density liquid phase leads to reduced contact
650     between the hexane and butanethiol, and this accounts for our
651     observation of lower conductance at higher temperatures. In raising
652     the average temperature from 200K to 250K, the density drop of
653     $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
654     conductance.
655    
656     Similar behavior is observed in the TraPPE-UA model for toluene,
657     although this model has better agreement with the experimental
658     densities of toluene. The expansion of the toluene liquid phase is
659     not as significant as that of the hexane (8.3\% over 100K), and this
660     limits the effect to $\sim$20\% drop in thermal conductivity.
661    
662     Although we have not mapped out the behavior at a large number of
663     temperatures, is clear that there will be a strong temperature
664     dependence in the interfacial conductance when the physical properties
665     of one side of the interface (notably the density) change rapidly as a
666     function of temperature.
667    
668     Besides the lower interfacial thermal conductance, surfaces at
669     relatively high temperatures are susceptible to reconstructions,
670     particularly when butanethiols fully cover the Au(111) surface. These
671     reconstructions include surface Au atoms which migrate outward to the
672     S atom layer, and butanethiol molecules which embed into the surface
673     Au layer. The driving force for this behavior is the strong Au-S
674     interactions which are modeled here with a deep Lennard-Jones
675     potential. This phenomenon agrees with reconstructions that have been
676     experimentally
677     observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}. Vlugt
678     {\it et al.} kept their Au(111) slab rigid so that their simulations
679     could reach 300K without surface
680     reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
681     blur the interface, the measurement of $G$ becomes more difficult to
682     conduct at higher temperatures. For this reason, most of our
683     measurements are undertaken at $\langle T\rangle\sim$200K where
684     reconstruction is minimized.
685    
686     However, when the surface is not completely covered by butanethiols,
687     the simulated system appears to be more resistent to the
688     reconstruction. Our Au / butanethiol / toluene system had the Au(111)
689     surfaces 90\% covered by butanethiols, but did not see this above
690     phenomena even at $\langle T\rangle\sim$300K. That said, we did
691     observe butanethiols migrating to neighboring three-fold sites during
692     a simulation. Since the interface persisted in these simulations, we
693     were able to obtain $G$'s for these interfaces even at a relatively
694     high temperature without being affected by surface reconstructions.
695    
696     \section{Discussion}
697 skuang 3770 [COMBINE W. RESULTS]
698 gezelter 3769 The primary result of this work is that the capping agent acts as an
699     efficient thermal coupler between solid and solvent phases. One of
700     the ways the capping agent can carry out this role is to down-shift
701     between the phonon vibrations in the solid (which carry the heat from
702     the gold) and the molecular vibrations in the liquid (which carry some
703     of the heat in the solvent).
704    
705     To investigate the mechanism of interfacial thermal conductance, the
706     vibrational power spectrum was computed. Power spectra were taken for
707     individual components in different simulations. To obtain these
708     spectra, simulations were run after equilibration in the
709     microcanonical (NVE) ensemble and without a thermal
710     gradient. Snapshots of configurations were collected at a frequency
711     that is higher than that of the fastest vibrations occurring in the
712     simulations. With these configurations, the velocity auto-correlation
713     functions can be computed:
714     \begin{equation}
715     C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
716     \label{vCorr}
717     \end{equation}
718     The power spectrum is constructed via a Fourier transform of the
719     symmetrized velocity autocorrelation function,
720     \begin{equation}
721     \hat{f}(\omega) =
722     \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
723     \label{fourier}
724     \end{equation}
725    
726     \subsection{The role of specific vibrations}
727     The vibrational spectra for gold slabs in different environments are
728     shown as in Figure \ref{specAu}. Regardless of the presence of
729     solvent, the gold surfaces which are covered by butanethiol molecules
730     exhibit an additional peak observed at a frequency of
731     $\sim$165cm$^{-1}$. We attribute this peak to the S-Au bonding
732     vibration. This vibration enables efficient thermal coupling of the
733     surface Au layer to the capping agents. Therefore, in our simulations,
734     the Au / S interfaces do not appear to be the primary barrier to
735     thermal transport when compared with the butanethiol / solvent
736     interfaces. This supports the results of Luo {\it et
737     al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
738     twice as large as what we have computed for the thiol-liquid
739     interfaces.
740    
741     \begin{figure}
742     \includegraphics[width=\linewidth]{vibration}
743     \caption{The vibrational power spectrum for thiol-capped gold has an
744     additional vibrational peak at $\sim $165cm$^{-1}$. Bare gold
745     surfaces (both with and without a solvent over-layer) are missing
746     this peak. A similar peak at $\sim $165cm$^{-1}$ also appears in
747     the vibrational power spectrum for the butanethiol capping agents.}
748     \label{specAu}
749     \end{figure}
750    
751     Also in this figure, we show the vibrational power spectrum for the
752     bound butanethiol molecules, which also exhibits the same
753     $\sim$165cm$^{-1}$ peak.
754    
755     \subsection{Overlap of power spectra}
756     A comparison of the results obtained from the two different organic
757     solvents can also provide useful information of the interfacial
758     thermal transport process. In particular, the vibrational overlap
759     between the butanethiol and the organic solvents suggests a highly
760     efficient thermal exchange between these components. Very high
761     thermal conductivity was observed when AA models were used and C-H
762     vibrations were treated classically. The presence of extra degrees of
763     freedom in the AA force field yields higher heat exchange rates
764     between the two phases and results in a much higher conductivity than
765     in the UA force field. The all-atom classical models include high
766     frequency modes which should be unpopulated at our relatively low
767     temperatures. This artifact is likely the cause of the high thermal
768     conductance in all-atom MD simulations.
769    
770     The similarity in the vibrational modes available to solvent and
771     capping agent can be reduced by deuterating one of the two components
772     (Fig. \ref{aahxntln}). Once either the hexanes or the butanethiols
773     are deuterated, one can observe a significantly lower $G$ and
774     $G^\prime$ values (Table \ref{modelTest}).
775    
776     \begin{figure}
777     \includegraphics[width=\linewidth]{aahxntln}
778     \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
779     systems. When butanethiol is deuterated (lower left), its
780     vibrational overlap with hexane decreases significantly. Since
781     aromatic molecules and the butanethiol are vibrationally dissimilar,
782     the change is not as dramatic when toluene is the solvent (right).}
783     \label{aahxntln}
784     \end{figure}
785    
786     For the Au / butanethiol / toluene interfaces, having the AA
787     butanethiol deuterated did not yield a significant change in the
788     measured conductance. Compared to the C-H vibrational overlap between
789     hexane and butanethiol, both of which have alkyl chains, the overlap
790     between toluene and butanethiol is not as significant and thus does
791     not contribute as much to the heat exchange process.
792    
793     Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
794     that the {\it intra}molecular heat transport due to alkylthiols is
795     highly efficient. Combining our observations with those of Zhang {\it
796     et al.}, it appears that butanethiol acts as a channel to expedite
797     heat flow from the gold surface and into the alkyl chain. The
798     vibrational coupling between the metal and the liquid phase can
799     therefore be enhanced with the presence of suitable capping agents.
800    
801     Deuterated models in the UA force field did not decouple the thermal
802     transport as well as in the AA force field. The UA models, even
803     though they have eliminated the high frequency C-H vibrational
804     overlap, still have significant overlap in the lower-frequency
805     portions of the infrared spectra (Figure \ref{uahxnua}). Deuterating
806     the UA models did not decouple the low frequency region enough to
807     produce an observable difference for the results of $G$ (Table
808     \ref{modelTest}).
809    
810     \begin{figure}
811     \includegraphics[width=\linewidth]{uahxnua}
812     \caption{Vibrational power spectra for UA models for the butanethiol
813     and hexane solvent (upper panel) show the high degree of overlap
814     between these two molecules, particularly at lower frequencies.
815     Deuterating a UA model for the solvent (lower panel) does not
816     decouple the two spectra to the same degree as in the AA force
817     field (see Fig \ref{aahxntln}).}
818     \label{uahxnua}
819     \end{figure}
820    
821     \section{Conclusions}
822     The NIVS algorithm has been applied to simulations of
823     butanethiol-capped Au(111) surfaces in the presence of organic
824     solvents. This algorithm allows the application of unphysical thermal
825     flux to transfer heat between the metal and the liquid phase. With the
826     flux applied, we were able to measure the corresponding thermal
827     gradients and to obtain interfacial thermal conductivities. Under
828     steady states, 2-3 ns trajectory simulations are sufficient for
829     computation of this quantity.
830    
831     Our simulations have seen significant conductance enhancement in the
832     presence of capping agent, compared with the bare gold / liquid
833     interfaces. The vibrational coupling between the metal and the liquid
834     phase is enhanced by a chemically-bonded capping agent. Furthermore,
835     the coverage percentage of the capping agent plays an important role
836     in the interfacial thermal transport process. Moderately low coverages
837     allow higher contact between capping agent and solvent, and thus could
838     further enhance the heat transfer process, giving a non-monotonic
839     behavior of conductance with increasing coverage.
840    
841     Our results, particularly using the UA models, agree well with
842     available experimental data. The AA models tend to overestimate the
843     interfacial thermal conductance in that the classically treated C-H
844     vibrations become too easily populated. Compared to the AA models, the
845     UA models have higher computational efficiency with satisfactory
846     accuracy, and thus are preferable in modeling interfacial thermal
847     transport.
848    
849     Of the two definitions for $G$, the discrete form
850     (Eq. \ref{discreteG}) was easier to use and gives out relatively
851     consistent results, while the derivative form (Eq. \ref{derivativeG})
852     is not as versatile. Although $G^\prime$ gives out comparable results
853     and follows similar trend with $G$ when measuring close to fully
854     covered or bare surfaces, the spatial resolution of $T$ profile
855     required for the use of a derivative form is limited by the number of
856     bins and the sampling required to obtain thermal gradient information.
857    
858     Vlugt {\it et al.} have investigated the surface thiol structures for
859     nanocrystalline gold and pointed out that they differ from those of
860     the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
861     difference could also cause differences in the interfacial thermal
862     transport behavior. To investigate this problem, one would need an
863     effective method for applying thermal gradients in non-planar
864     (i.e. spherical) geometries.
865    
866     \section{Acknowledgments}
867     Support for this project was provided by the National Science
868     Foundation under grant CHE-0848243. Computational time was provided by
869     the Center for Research Computing (CRC) at the University of Notre
870     Dame.
871    
872     \newpage
873    
874 skuang 3770 \bibliography{stokes}
875 gezelter 3769
876     \end{doublespace}
877     \end{document}
878