ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
(Generate patch)

Comparing stokes/stokes.tex (file contents):
Revision 3770 by skuang, Fri Dec 2 20:14:03 2011 UTC vs.
Revision 3779 by skuang, Sat Dec 10 23:46:08 2011 UTC

# Line 43 | Line 43 | Notre Dame, Indiana 46556}
43   \begin{doublespace}
44  
45   \begin{abstract}
46 <  REPLACE ABSTRACT HERE
47 <  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.
46 >  We present a new method for introducing stable nonequilibrium
47 >  velocity and temperature gradients in molecular dynamics simulations
48 >  of heterogeneous systems. This method conserves the linear momentum
49 >  and total energy of the system and improves previous Reverse
50 >  Non-Equilibrium Molecular Dynamics (RNEMD) methods and maintains
51 >  thermal velocity distributions. It also avoid thermal anisotropy
52 >  occured in NIVS simulations by using isotropic velocity scaling on
53 >  the molecules in specific regions of a system. To test the method,
54 >  we have computed the thermal conductivity and shear viscosity of
55 >  model liquid systems as well as the interfacial frictions of a
56 >  series of  metal/liquid interfaces.
57  
58   \end{abstract}
59  
# Line 74 | Line 66 | Notre Dame, Indiana 46556}
66   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67  
68   \section{Introduction}
69 < [DO THIS LATER]
69 > [REFINE LATER, ADD MORE REF.S]
70 > Imposed-flux methods in Molecular Dynamics (MD)
71 > simulations\cite{MullerPlathe:1997xw} can establish steady state
72 > systems with a set applied flux vs a corresponding gradient that can
73 > be measured. These methods does not need many trajectories to provide
74 > information of transport properties of a given system. Thus, they are
75 > utilized in computing thermal and mechanical transfer of homogeneous
76 > or bulk systems as well as heterogeneous systems such as liquid-solid
77 > interfaces.\cite{kuang:AuThl}
78  
79 < [IMPORTANCE OF NANOSCALE TRANSPORT PROPERTIES STUDIES]
79 > The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that
80 > satisfy linear momentum and total energy conservation of a system when
81 > imposing fluxes in a simulation. Thus they are compatible with various
82 > ensembles, including the micro-canonical (NVE) ensemble, without the
83 > need of an external thermostat. The original approaches by
84 > M\"{u}ller-Plathe {\it et
85 >  al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
86 > momentum swapping for generating energy/momentum fluxes, which is also
87 > compatible with particles of different identities. Although simple to
88 > implement in a simulation, this approach can create nonthermal
89 > velocity distributions, as discovered by Tenney and
90 > Maginn\cite{Maginn:2010}. Furthermore, this approach to kinetic energy
91 > transfer between particles of different identities is less efficient
92 > when the mass difference between the particles becomes significant,
93 > which also limits its application on heterogeneous interfacial
94 > systems.
95  
96 < Due to the importance of heat flow (and heat removal) in
97 < nanotechnology, interfacial thermal conductance has been studied
98 < extensively both experimentally and computationally.\cite{cahill:793}
99 < Nanoscale materials have a significant fraction of their atoms at
100 < interfaces, and the chemical details of these interfaces govern the
101 < thermal transport properties.  Furthermore, the interfaces are often
102 < heterogeneous (e.g. solid - liquid), which provides a challenge to
103 < computational methods which have been developed for homogeneous or
104 < bulk systems.
96 > Recently, we developed a different approach, using Non-Isotropic
97 > Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose
98 > fluxes. Compared to the momentum swapping move, it scales the velocity
99 > vectors in two separate regions of a simulated system with respective
100 > diagonal scaling matrices. These matrices are determined by solving a
101 > set of equations including linear momentum and kinetic energy
102 > conservation constraints and target flux satisfaction. This method is
103 > able to effectively impose a wide range of kinetic energy fluxes
104 > without obvious perturbation to the velocity distributions of the
105 > simulated systems, regardless of the presence of heterogeneous
106 > interfaces. We have successfully applied this approach in studying the
107 > interfacial thermal conductance at metal-solvent
108 > interfaces.\cite{kuang:AuThl}
109  
110 < Experimentally, the thermal properties of a number of interfaces have
111 < been investigated.  Cahill and coworkers studied nanoscale thermal
112 < transport from metal nanoparticle/fluid interfaces, to epitaxial
113 < TiN/single crystal oxides interfaces, and hydrophilic and hydrophobic
114 < interfaces between water and solids with different self-assembled
115 < monolayers.\cite{Wilson:2002uq,PhysRevB.67.054302,doi:10.1021/jp048375k,PhysRevLett.96.186101}
116 < Wang {\it et al.} studied heat transport through long-chain
117 < hydrocarbon monolayers on gold substrate at individual molecular
99 < level,\cite{Wang10082007} Schmidt {\it et al.} studied the role of
100 < cetyltrimethylammonium bromide (CTAB) on the thermal transport between
101 < gold nanorods and solvent,\cite{doi:10.1021/jp8051888} and Juv\'e {\it
102 <  et al.} studied the cooling dynamics, which is controlled by thermal
103 < interface resistance of glass-embedded metal
104 < nanoparticles.\cite{PhysRevB.80.195406} Although interfaces are
105 < normally considered barriers for heat transport, Alper {\it et al.}
106 < suggested that specific ligands (capping agents) could completely
107 < eliminate this barrier
108 < ($G\rightarrow\infty$).\cite{doi:10.1021/la904855s}
110 > However, the NIVS approach limits its application in imposing momentum
111 > fluxes. Temperature anisotropy can happen under high momentum fluxes,
112 > due to the nature of the algorithm. Thus, combining thermal and
113 > momentum flux is also difficult to implement with this
114 > approach. However, such combination may provide a means to simulate
115 > thermal/momentum gradient coupled processes such as freeze
116 > desalination. Therefore, developing novel approaches to extend the
117 > application of imposed-flux method is desired.
118  
119 < The acoustic mismatch model for interfacial conductance utilizes the
120 < acoustic impedance ($Z_a = \rho_a v^s_a$) on both sides of the
121 < interface.\cite{swartz1989} Here, $\rho_a$ and $v^s_a$ are the density
122 < and speed of sound in material $a$.  The phonon transmission
123 < probability at the $a-b$ interface is
124 < \begin{equation}
125 < t_{ab} = \frac{4 Z_a Z_b}{\left(Z_a + Z_b \right)^2},
126 < \end{equation}
127 < and the interfacial conductance can then be approximated as
128 < \begin{equation}
120 < G_{ab} \approx \frac{1}{4} C_D v_D t_{ab}
121 < \end{equation}
122 < where $C_D$ is the Debye heat capacity of the hot side, and $v_D$ is
123 < the Debye phonon velocity ($1/v_D^3 = 1/3v_L^3 + 2/3 v_T^3$) where
124 < $v_L$ and $v_T$ are the longitudinal and transverse speeds of sound,
125 < respectively.  For the Au/hexane and Au/toluene interfaces, the
126 < acoustic mismatch model predicts room-temperature $G \approx 87 \mbox{
127 <  and } 129$ MW m$^{-2}$ K$^{-1}$, respectively.  However, it is not
128 < clear how to apply the acoustic mismatch model to a
129 < chemically-modified surface, particularly when the acoustic properties
130 < of a monolayer film may not be well characterized.
119 > In this paper, we improve the NIVS method and propose a novel approach
120 > to impose fluxes. This approach separate the means of applying
121 > momentum and thermal flux with operations in one time step and thus is
122 > able to simutaneously impose thermal and momentum flux. Furthermore,
123 > the approach retains desirable features of previous RNEMD approaches
124 > and is simpler to implement compared to the NIVS method. In what
125 > follows, we first present the method to implement the method in a
126 > simulation. Then we compare the method on bulk fluids to previous
127 > methods. Also, interfacial frictions are computed for a series of
128 > interfaces.
129  
132 [PREVIOUS METHODS INCLUDING NIVS AND THEIR LIMITATIONS]
133 [DIFFICULTY TO GENERATE JZKE AND JZP SIMUTANEOUSLY]
134
135 More precise computational models have also been used to study the
136 interfacial thermal transport in order to gain an understanding of
137 this phenomena at the molecular level. Recently, Hase and coworkers
138 employed Non-Equilibrium Molecular Dynamics (NEMD) simulations to
139 study thermal transport from hot Au(111) substrate to a self-assembled
140 monolayer of alkylthiol with relatively long chain (8-20 carbon
141 atoms).\cite{hase:2010,hase:2011} However, ensemble averaged
142 measurements for heat conductance of interfaces between the capping
143 monolayer on Au and a solvent phase have yet to be studied with their
144 approach. The comparatively low thermal flux through interfaces is
145 difficult to measure with Equilibrium
146 MD\cite{doi:10.1080/0026897031000068578} or forward NEMD simulation
147 methods. Therefore, the Reverse NEMD (RNEMD)
148 methods\cite{MullerPlathe:1997xw,kuang:164101} would be advantageous
149 in that they {\it apply} the difficult to measure quantity (flux),
150 while {\it measuring} the easily-computed quantity (the thermal
151 gradient).  This is particularly true for inhomogeneous interfaces
152 where it would not be clear how to apply a gradient {\it a priori}.
153 Garde and coworkers\cite{garde:nl2005,garde:PhysRevLett2009} applied
154 this approach to various liquid interfaces and studied how thermal
155 conductance (or resistance) is dependent on chemical details of a
156 number of hydrophobic and hydrophilic aqueous interfaces. And
157 recently, Luo {\it et al.} studied the thermal conductance of
158 Au-SAM-Au junctions using the same approach, comparing to a constant
159 temperature difference method.\cite{Luo20101} While this latter
160 approach establishes more ideal Maxwell-Boltzmann distributions than
161 previous RNEMD methods, it does not guarantee momentum or kinetic
162 energy conservation.
163
164 Recently, we have developed a Non-Isotropic Velocity Scaling (NIVS)
165 algorithm for RNEMD simulations\cite{kuang:164101}. This algorithm
166 retains the desirable features of RNEMD (conservation of linear
167 momentum and total energy, compatibility with periodic boundary
168 conditions) while establishing true thermal distributions in each of
169 the two slabs. Furthermore, it allows effective thermal exchange
170 between particles of different identities, and thus makes the study of
171 interfacial conductance much simpler.
172
173 [WHAT IS COVERED IN THIS MANUSCRIPT]
174 [MAY PUT FIGURE 1 HERE]
175 The work presented here deals with the Au(111) surface covered to
176 varying degrees by butanethiol, a capping agent with short carbon
177 chain, and solvated with organic solvents of different molecular
178 properties. Different models were used for both the capping agent and
179 the solvent force field parameters. Using the NIVS algorithm, the
180 thermal transport across these interfaces was studied and the
181 underlying mechanism for the phenomena was investigated.
182
130   \section{Methodology}
131   Similar to the NIVS methodology,\cite{kuang:164101} we consider a
132   periodic system divided into a series of slabs along a certain axis
# Line 218 | Line 165 | thermal flux $J_z$, one would have
165   The above operations conserve the linear momentum of a periodic
166   system. To satisfy total energy conservation as well as to impose a
167   thermal flux $J_z$, one would have
168 < %SUPPORT INFO MIGHT BE NECESSARY TO PUT EXTRA MATH IN
168 > [SUPPORT INFO MIGHT BE NECESSARY TO PUT EXTRA MATH IN]
169   \begin{eqnarray}
170   K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
171   \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
# Line 236 | Line 183 | the positive roots (which are closer to 1) are chosen.
183   scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
184   $\vec{a}_h$. Note that two roots of $c$ and $h$ exist
185   respectively. However, to avoid dramatic perturbations to a system,
186 < the positive roots (which are closer to 1) are chosen.
187 <
186 > the positive roots (which are closer to 1) are chosen. Figure
187 > \ref{method} illustrates the implementation of this algorithm in an
188 > individual step.
189 >
190 > \begin{figure}
191 > \includegraphics[width=\linewidth]{method}
192 > \caption{Illustration of the implementation of the algorithm in a
193 >  single step. Starting from an ideal velocity distribution, the
194 >  transformation is used to apply both thermal and momentum flux from
195 >  the ``c'' slab to the ``h'' slab. As the figure shows, the thermal
196 >  distributions preserve after this operation.}
197 > \label{method}
198 > \end{figure}
199 >
200   By implementing these operations at a certain frequency, a steady
201   thermal and/or momentum flux can be applied and the corresponding
202   temperature and/or momentum gradients can be established.
244 [REFER TO NIVS PAPER]
245 [ADVANTAGES]
203  
204 < Steady state MD simulations have an advantage in that not many
205 < trajectories are needed to study the relationship between thermal flux
206 < and thermal gradients. For systems with low interfacial conductance,
207 < one must have a method capable of generating or measuring relatively
208 < small fluxes, compared to those required for bulk conductivity. This
209 < requirement makes the calculation even more difficult for
210 < slowly-converging equilibrium methods.\cite{Viscardy:2007lq} Forward
211 < NEMD methods impose a gradient (and measure a flux), but at interfaces
212 < it is not clear what behavior should be imposed at the boundaries
256 < between materials.  Imposed-flux reverse non-equilibrium
257 < methods\cite{MullerPlathe:1997xw} set the flux {\it a priori} and
258 < the thermal response becomes an easy-to-measure quantity.  Although
259 < M\"{u}ller-Plathe's original momentum swapping approach can be used
260 < for exchanging energy between particles of different identity, the
261 < kinetic energy transfer efficiency is affected by the mass difference
262 < between the particles, which limits its application on heterogeneous
263 < interfacial systems.
204 > This approach is more computationaly efficient compared to the
205 > previous NIVS method, in that only quadratic equations are involved,
206 > while the NIVS method needs to solve a quartic equations. Furthermore,
207 > the method implements isotropic scaling of velocities in respective
208 > slabs, unlike the NIVS, where an extra criteria function is necessary
209 > to choose a set of coefficients that performs the most isotropic
210 > scaling. More importantly, separating the momentum flux imposing from
211 > velocity scaling avoids the underlying cause that NIVS produced
212 > thermal anisotropy when applying a momentum flux.
213  
214 < The non-isotropic velocity scaling (NIVS) \cite{kuang:164101} approach
215 < to non-equilibrium MD simulations is able to impose a wide range of
216 < kinetic energy fluxes without obvious perturbation to the velocity
217 < distributions of the simulated systems. Furthermore, this approach has
218 < the advantage in heterogeneous interfaces in that kinetic energy flux
219 < can be applied between regions of particles of arbitrary identity, and
220 < the flux will not be restricted by difference in particle mass.
214 > The advantages of the approach over the original momentum swapping
215 > approach lies in its nature to preserve a Gaussian
216 > distribution. Because the momentum swapping tends to render a
217 > nonthermal distribution, when the imposed flux is relatively large,
218 > diffusion of the neighboring slabs could no longer remedy this effect,
219 > and nonthermal distributions would be observed. Results in later
220 > section will illustrate this effect.
221  
222 < The NIVS algorithm scales the velocity vectors in two separate regions
223 < of a simulation system with respective diagonal scaling matrices. To
224 < determine these scaling factors in the matrices, a set of equations
225 < including linear momentum conservation and kinetic energy conservation
226 < constraints and target energy flux satisfaction is solved. With the
227 < scaling operation applied to the system in a set frequency, bulk
228 < temperature gradients can be easily established, and these can be used
229 < for computing thermal conductivities. The NIVS algorithm conserves
281 < momenta and energy and does not depend on an external thermostat.
222 > \section{Computational Details}
223 > The algorithm has been implemented in our MD simulation code,
224 > OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with
225 > previous RNEMD methods or equilibrium MD methods in homogeneous fluids
226 > (Lennard-Jones and SPC/E water). And taking advantage of the method,
227 > we simulate the interfacial friction of different heterogeneous
228 > interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid
229 > water).
230  
231 < \subsection{Defining Interfacial Thermal Conductivity ($G$)}
231 > \subsection{Simulation Protocols}
232 > The systems to be investigated are set up in a orthorhombic simulation
233 > cell with periodic boundary conditions in all three dimensions. The
234 > $z$ axis of these cells were longer and was set as the gradient axis
235 > of temperature and/or momentum. Thus the cells were divided into $N$
236 > slabs along this axis, with various $N$ depending on individual
237 > system. The $x$ and $y$ axis were usually of the same length in
238 > homogeneous systems or close to each other where interfaces
239 > presents. In all cases, before introducing a nonequilibrium method to
240 > establish steady thermal and/or momentum gradients for further
241 > measurements and calculations, canonical ensemble with a Nos\'e-Hoover
242 > thermostat\cite{hoover85} and microcanonical ensemble equilibrations
243 > were used to prepare systems ready for data
244 > collections. Isobaric-isothermal equilibrations are performed before
245 > this for SPC/E water systems to reach normal pressure (1 bar), while
246 > similar equilibrations are used for interfacial systems to relax the
247 > surface tensions.
248  
249 < For an interface with relatively low interfacial conductance, and a
250 < thermal flux between two distinct bulk regions, the regions on either
251 < side of the interface rapidly come to a state in which the two phases
252 < have relatively homogeneous (but distinct) temperatures. The
253 < interfacial thermal conductivity $G$ can therefore be approximated as:
254 < \begin{equation}
291 <  G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
292 <    \langle T_\mathrm{cold}\rangle \right)}
293 < \label{lowG}
294 < \end{equation}
295 < where ${E_{total}}$ is the total imposed non-physical kinetic energy
296 < transfer during the simulation and ${\langle T_\mathrm{hot}\rangle}$
297 < and ${\langle T_\mathrm{cold}\rangle}$ are the average observed
298 < temperature of the two separated phases.  For an applied flux $J_z$
299 < operating over a simulation time $t$ on a periodically-replicated slab
300 < of dimensions $L_x \times L_y$, $E_{total} = 2 J_z t L_x L_y$.
249 > While homogeneous fluid systems can be set up with random
250 > configurations, our interfacial systems needs extra steps to ensure
251 > the interfaces be established properly for computations. The
252 > preparation and equilibration of butanethiol covered gold (111)
253 > surface and further solvation and equilibration process is described
254 > as in reference \cite{kuang:AuThl}.
255  
256 < When the interfacial conductance is {\it not} small, there are two
257 < ways to define $G$. One common way is to assume the temperature is
258 < discrete on the two sides of the interface. $G$ can be calculated
259 < using the applied thermal flux $J$ and the maximum temperature
260 < difference measured along the thermal gradient max($\Delta T$), which
261 < occurs at the Gibbs dividing surface (Figure \ref{demoPic}). This is
262 < known as the Kapitza conductance, which is the inverse of the Kapitza
263 < resistance.
264 < \begin{equation}
265 <  G=\frac{J}{\Delta T}
266 < \label{discreteG}
267 < \end{equation}
256 > As for the ice/liquid water interfaces, the basal surface of ice
257 > lattice was first constructed. Hirsch {\it et
258 >  al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice
259 > lattices with different proton orders. We refer to their results and
260 > choose the configuration of the lowest energy after geometry
261 > optimization as the unit cells of our ice lattices. Although
262 > experimental solid/liquid coexistant temperature near normal pressure
263 > is 273K, Bryk and Haymet's simulations of ice/liquid water interfaces
264 > with different models suggest that for SPC/E, the most stable
265 > interface is observed at 225$\pm$5K. Therefore, all our ice/liquid
266 > water simulations were carried out under 225K. To have extra
267 > protection of the ice lattice during initial equilibration (when the
268 > randomly generated liquid phase configuration could release large
269 > amount of energy in relaxation), a constraint method (REF?) was
270 > adopted until the high energy configuration was relaxed.
271 > [MAY ADD A FIGURE HERE FOR BASAL PLANE, MAY INCLUDE PRISM IF POSSIBLE]
272  
273 < \begin{figure}
274 < \includegraphics[width=\linewidth]{method}
275 < \caption{Interfacial conductance can be calculated by applying an
276 <  (unphysical) kinetic energy flux between two slabs, one located
277 <  within the metal and another on the edge of the periodic box.  The
320 <  system responds by forming a thermal gradient.  In bulk liquids,
321 <  this gradient typically has a single slope, but in interfacial
322 <  systems, there are distinct thermal conductivity domains.  The
323 <  interfacial conductance, $G$ is found by measuring the temperature
324 <  gap at the Gibbs dividing surface, or by using second derivatives of
325 <  the thermal profile.}
326 < \label{demoPic}
327 < \end{figure}
273 > \subsection{Force Field Parameters}
274 > For comparison of our new method with previous work, we retain our
275 > force field parameters consistent with the results we will compare
276 > with. The Lennard-Jones fluid used here for argon , and reduced unit
277 > results are reported for direct comparison purpose.
278  
279 < Another approach is to assume that the temperature is continuous and
280 < differentiable throughout the space. Given that $\lambda$ is also
281 < differentiable, $G$ can be defined as its gradient ($\nabla\lambda$)
282 < projected along a vector normal to the interface ($\mathbf{\hat{n}}$)
333 < and evaluated at the interface location ($z_0$). This quantity,
334 < \begin{align}
335 < G^\prime &= \left(\nabla\lambda \cdot \mathbf{\hat{n}}\right)_{z_0} \\
336 <         &= \frac{\partial}{\partial z}\left(-\frac{J_z}{
337 <           \left(\frac{\partial T}{\partial z}\right)}\right)_{z_0} \\
338 <         &= J_z\left(\frac{\partial^2 T}{\partial z^2}\right)_{z_0} \Big/
339 <         \left(\frac{\partial T}{\partial z}\right)_{z_0}^2 \label{derivativeG}
340 < \end{align}
341 < has the same units as the common definition for $G$, and the maximum
342 < of its magnitude denotes where thermal conductivity has the largest
343 < change, i.e. the interface.  In the geometry used in this study, the
344 < vector normal to the interface points along the $z$ axis, as do
345 < $\vec{J}$ and the thermal gradient.  This yields the simplified
346 < expressions in Eq. \ref{derivativeG}.
279 > As for our water simulations, SPC/E model is used throughout this work
280 > for consistency. Previous work for transport properties of SPC/E water
281 > model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
282 > that unnecessary repetition of previous methods can be avoided.
283  
284 < With temperature profiles obtained from simulation, one is able to
285 < approximate the first and second derivatives of $T$ with finite
286 < difference methods and calculate $G^\prime$. In what follows, both
287 < definitions have been used, and are compared in the results.
284 > The Au-Au interaction parameters in all simulations are described by
285 > the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
286 > QSC potentials include zero-point quantum corrections and are
287 > reparametrized for accurate surface energies compared to the
288 > Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
289 > Spohr potential was adopted\cite{ISI:000167766600035} to depict
290 > Au-H$_2$O interactions.
291  
292 < To investigate the interfacial conductivity at metal / solvent
293 < interfaces, we have modeled a metal slab with its (111) surfaces
294 < perpendicular to the $z$-axis of our simulation cells. The metal slab
295 < has been prepared both with and without capping agents on the exposed
296 < surface, and has been solvated with simple organic solvents, as
297 < illustrated in Figure \ref{gradT}.
292 > The small organic molecules included in our simulations are the Au
293 > surface capping agent butanethiol and liquid hexane and toluene. The
294 > United-Atom
295 > models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
296 > for these components were used in this work for better computational
297 > efficiency, while maintaining good accuracy. We refer readers to our
298 > previous work\cite{kuang:AuThl} for further details of these models,
299 > as well as the interactions between Au and the above organic molecule
300 > components.
301  
302 < With the simulation cell described above, we are able to equilibrate
303 < the system and impose an unphysical thermal flux between the liquid
304 < and the metal phase using the NIVS algorithm. By periodically applying
305 < the unphysical flux, we obtained a temperature profile and its spatial
306 < derivatives. Figure \ref{gradT} shows how an applied thermal flux can
307 < be used to obtain the 1st and 2nd derivatives of the temperature
308 < profile.
302 > \subsection{Thermal conductivities}
303 > When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
304 > impose kinetic energy transfer, the method can be used for thermal
305 > conductivity computations. Similar to previous RNEMD methods, we
306 > assume linear response of the temperature gradient with respect to the
307 > thermal flux in general case. And the thermal conductivity ($\lambda$)
308 > can be obtained with the imposed kinetic energy flux and the measured
309 > thermal gradient:
310 > \begin{equation}
311 > J_z = -\lambda \frac{\partial T}{\partial z}
312 > \end{equation}
313 > Like other imposed-flux methods, the energy flux was calculated using
314 > the total non-physical energy transferred (${E_{total}}$) from slab
315 > ``c'' to slab ``h'', which is recorded throughout a simulation, and
316 > the time for data collection $t$:
317 > \begin{equation}
318 > J_z = \frac{E_{total}}{2 t L_x L_y}
319 > \end{equation}
320 > where $L_x$ and $L_y$ denotes the dimensions of the plane in a
321 > simulation cell perpendicular to the thermal gradient, and a factor of
322 > two in the denominator is present for the heat transport occurs in
323 > both $+z$ and $-z$ directions. The temperature gradient
324 > ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
325 > regression of the temperature profile, which is recorded during a
326 > simulation for each slab in a cell. For Lennard-Jones simulations,
327 > thermal conductivities are reported in reduced units
328 > (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
329  
330 < \begin{figure}
331 < \includegraphics[width=\linewidth]{gradT}
332 < \caption{A sample of Au (111) / butanethiol / hexane interfacial
333 <  system with the temperature profile after a kinetic energy flux has
334 <  been imposed.  Note that the largest temperature jump in the thermal
335 <  profile (corresponding to the lowest interfacial conductance) is at
336 <  the interface between the butanethiol molecules (blue) and the
337 <  solvent (grey).  First and second derivatives of the temperature
338 <  profile are obtained using a finite difference approximation (lower
377 <  panel).}
378 < \label{gradT}
379 < \end{figure}
330 > \subsection{Shear viscosities}
331 > Alternatively, the method can carry out shear viscosity calculations
332 > by switching off $J_z$. One can specify the vector
333 > $\vec{j}_z(\vec{p})$ by choosing the three components
334 > respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
335 > set to zero. Although for isotropic systems, the direction of
336 > $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, the ability
337 > of arbitarily specifying the vector direction in our method provides
338 > convenience in anisotropic simulations.
339  
340 < \section{Computational Details}
341 < \subsection{Simulation Protocol}
342 < The NIVS algorithm has been implemented in our MD simulation code,
343 < OpenMD\cite{Meineke:2005gd,openmd}, and was used for our simulations.
344 < Metal slabs of 6 or 11 layers of Au atoms were first equilibrated
345 < under atmospheric pressure (1 atm) and 200K. After equilibration,
346 < butanethiol capping agents were placed at three-fold hollow sites on
347 < the Au(111) surfaces. These sites are either {\it fcc} or {\it
348 <  hcp} sites, although Hase {\it et al.} found that they are
349 < equivalent in a heat transfer process,\cite{hase:2010} so we did not
350 < distinguish between these sites in our study. The maximum butanethiol
351 < capacity on Au surface is $1/3$ of the total number of surface Au
352 < atoms, and the packing forms a $(\sqrt{3}\times\sqrt{3})R30^\circ$
353 < structure\cite{doi:10.1021/ja00008a001,doi:10.1021/cr9801317}. A
354 < series of lower coverages was also prepared by eliminating
355 < butanethiols from the higher coverage surface in a regular manner. The
356 < lower coverages were prepared in order to study the relation between
357 < coverage and interfacial conductance.
340 > Similar to thermal conductivity computations, linear response of the
341 > momentum gradient with respect to the shear stress is assumed, and the
342 > shear viscosity ($\eta$) can be obtained with the imposed momentum
343 > flux (e.g. in $x$ direction) and the measured gradient:
344 > \begin{equation}
345 > j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
346 > \end{equation}
347 > where the flux is similarly defined:
348 > \begin{equation}
349 > j_z(p_x) = \frac{P_x}{2 t L_x L_y}
350 > \end{equation}
351 > with $P_x$ being the total non-physical momentum transferred within
352 > the data collection time. Also, the velocity gradient
353 > ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
354 > regression of the $x$ component of the mean velocity, $\langle
355 > v_x\rangle$, in each of the bins. For Lennard-Jones simulations, shear
356 > viscosities are reported in reduced units
357 > (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
358  
359 < The capping agent molecules were allowed to migrate during the
360 < simulations. They distributed themselves uniformly and sampled a
361 < number of three-fold sites throughout out study. Therefore, the
362 < initial configuration does not noticeably affect the sampling of a
363 < variety of configurations of the same coverage, and the final
364 < conductance measurement would be an average effect of these
365 < configurations explored in the simulations.
366 <
367 < After the modified Au-butanethiol surface systems were equilibrated in
409 < the canonical (NVT) ensemble, organic solvent molecules were packed in
410 < the previously empty part of the simulation cells.\cite{packmol} Two
411 < solvents were investigated, one which has little vibrational overlap
412 < with the alkanethiol and which has a planar shape (toluene), and one
413 < which has similar vibrational frequencies to the capping agent and
414 < chain-like shape ({\it n}-hexane).
415 <
416 < The simulation cells were not particularly extensive along the
417 < $z$-axis, as a very long length scale for the thermal gradient may
418 < cause excessively hot or cold temperatures in the middle of the
419 < solvent region and lead to undesired phenomena such as solvent boiling
420 < or freezing when a thermal flux is applied. Conversely, too few
421 < solvent molecules would change the normal behavior of the liquid
422 < phase. Therefore, our $N_{solvent}$ values were chosen to ensure that
423 < these extreme cases did not happen to our simulations. The spacing
424 < between periodic images of the gold interfaces is $45 \sim 75$\AA in
425 < our simulations.
426 <
427 < The initial configurations generated are further equilibrated with the
428 < $x$ and $y$ dimensions fixed, only allowing the $z$-length scale to
429 < change. This is to ensure that the equilibration of liquid phase does
430 < not affect the metal's crystalline structure. Comparisons were made
431 < with simulations that allowed changes of $L_x$ and $L_y$ during NPT
432 < equilibration. No substantial changes in the box geometry were noticed
433 < in these simulations. After ensuring the liquid phase reaches
434 < equilibrium at atmospheric pressure (1 atm), further equilibration was
435 < carried out under canonical (NVT) and microcanonical (NVE) ensembles.
436 <
437 < After the systems reach equilibrium, NIVS was used to impose an
438 < unphysical thermal flux between the metal and the liquid phases. Most
439 < of our simulations were done under an average temperature of
440 < $\sim$200K. Therefore, thermal flux usually came from the metal to the
441 < liquid so that the liquid has a higher temperature and would not
442 < freeze due to lowered temperatures. After this induced temperature
443 < gradient had stabilized, the temperature profile of the simulation cell
444 < was recorded. To do this, the simulation cell is divided evenly into
445 < $N$ slabs along the $z$-axis. The average temperatures of each slab
446 < are recorded for 1$\sim$2 ns. When the slab width $d$ of each slab is
447 < the same, the derivatives of $T$ with respect to slab number $n$ can
448 < be directly used for $G^\prime$ calculations: \begin{equation}
449 <  G^\prime = |J_z|\Big|\frac{\partial^2 T}{\partial z^2}\Big|
450 <         \Big/\left(\frac{\partial T}{\partial z}\right)^2
451 <         = |J_z|\Big|\frac{1}{d^2}\frac{\partial^2 T}{\partial n^2}\Big|
452 <         \Big/\left(\frac{1}{d}\frac{\partial T}{\partial n}\right)^2
453 <         = |J_z|\Big|\frac{\partial^2 T}{\partial n^2}\Big|
454 <         \Big/\left(\frac{\partial T}{\partial n}\right)^2
455 < \label{derivativeG2}
359 > \subsection{Interfacial friction and Slip length}
360 > While the shear stress results in a velocity gradient within bulk
361 > fluid phase, its effect at a solid-liquid interface could vary due to
362 > the interaction strength between the two phases. The interfacial
363 > friction coefficient $\kappa$ is defined to relate the shear stress
364 > (e.g. along $x$-axis) and the relative fluid velocity tangent to the
365 > interface:
366 > \begin{equation}
367 > j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
368   \end{equation}
369 < The absolute values in Eq. \ref{derivativeG2} appear because the
370 < direction of the flux $\vec{J}$ is in an opposing direction on either
371 < side of the metal slab.
369 > Under ``stick'' boundary condition, $\Delta v_x|_{interface}
370 > \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
371 > ``slip'' boundary condition at the solid-liquid interface, $\kappa$
372 > becomes finite. To characterize the interfacial boundary conditions,
373 > slip length ($\delta$) is defined using $\kappa$ and the shear
374 > viscocity of liquid phase ($\eta$):
375 > \begin{equation}
376 > \delta = \frac{\eta}{\kappa}
377 > \end{equation}
378 > so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
379 > and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
380 > illustrates how this quantity is defined and computed for a
381 > solid-liquid interface. [MAY INCLUDE A SNAPSHOT IN FIGURE]
382  
461 All of the above simulation procedures use a time step of 1 fs. Each
462 equilibration stage took a minimum of 100 ps, although in some cases,
463 longer equilibration stages were utilized.
464
465 \subsection{Force Field Parameters}
466 Our simulations include a number of chemically distinct components.
467 Figure \ref{demoMol} demonstrates the sites defined for both
468 United-Atom and All-Atom models of the organic solvent and capping
469 agents in our simulations. Force field parameters are needed for
470 interactions both between the same type of particles and between
471 particles of different species.
472
383   \begin{figure}
384 < \includegraphics[width=\linewidth]{structures}
385 < \caption{Structures of the capping agent and solvents utilized in
386 <  these simulations. The chemically-distinct sites (a-e) are expanded
387 <  in terms of constituent atoms for both United Atom (UA) and All Atom
388 <  (AA) force fields.  Most parameters are from References
479 <  \protect\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes,TraPPE-UA.thiols}
480 <  (UA) and \protect\cite{OPLSAA} (AA). Cross-interactions with the Au
481 <  atoms are given in Table 1 in the supporting information.}
482 < \label{demoMol}
384 > \includegraphics[width=\linewidth]{defDelta}
385 > \caption{The slip length $\delta$ can be obtained from a velocity
386 >  profile of a solid-liquid interface. An example of Au/hexane
387 >  interfaces is shown.}
388 > \label{slipLength}
389   \end{figure}
390  
391 < The Au-Au interactions in metal lattice slab is described by the
392 < quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The QSC
393 < potentials include zero-point quantum corrections and are
394 < reparametrized for accurate surface energies compared to the
395 < Sutton-Chen potentials.\cite{Chen90}
396 <
397 < For the two solvent molecules, {\it n}-hexane and toluene, two
398 < different atomistic models were utilized. Both solvents were modeled
493 < using United-Atom (UA) and All-Atom (AA) models. The TraPPE-UA
494 < parameters\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} are used
495 < for our UA solvent molecules. In these models, sites are located at
496 < the carbon centers for alkyl groups. Bonding interactions, including
497 < bond stretches and bends and torsions, were used for intra-molecular
498 < sites closer than 3 bonds. For non-bonded interactions, Lennard-Jones
499 < potentials are used.
500 <
501 < By eliminating explicit hydrogen atoms, the TraPPE-UA models are
502 < simple and computationally efficient, while maintaining good accuracy.
503 < However, the TraPPE-UA model for alkanes is known to predict a slightly
504 < lower boiling point than experimental values. This is one of the
505 < reasons we used a lower average temperature (200K) for our
506 < simulations. If heat is transferred to the liquid phase during the
507 < NIVS simulation, the liquid in the hot slab can actually be
508 < substantially warmer than the mean temperature in the simulation. The
509 < lower mean temperatures therefore prevent solvent boiling.
510 <
511 < For UA-toluene, the non-bonded potentials between intermolecular sites
512 < have a similar Lennard-Jones formulation. The toluene molecules were
513 < treated as a single rigid body, so there was no need for
514 < intramolecular interactions (including bonds, bends, or torsions) in
515 < this solvent model.
516 <
517 < Besides the TraPPE-UA models, AA models for both organic solvents are
518 < included in our studies as well. The OPLS-AA\cite{OPLSAA} force fields
519 < were used. For hexane, additional explicit hydrogen sites were
520 < included. Besides bonding and non-bonded site-site interactions,
521 < partial charges and the electrostatic interactions were added to each
522 < CT and HC site. For toluene, a flexible model for the toluene molecule
523 < was utilized which included bond, bend, torsion, and inversion
524 < potentials to enforce ring planarity.
525 <
526 < The butanethiol capping agent in our simulations, were also modeled
527 < with both UA and AA model. The TraPPE-UA force field includes
528 < parameters for thiol molecules\cite{TraPPE-UA.thiols} and are used for
529 < UA butanethiol model in our simulations. The OPLS-AA also provides
530 < parameters for alkyl thiols. However, alkyl thiols adsorbed on Au(111)
531 < surfaces do not have the hydrogen atom bonded to sulfur. To derive
532 < suitable parameters for butanethiol adsorbed on Au(111) surfaces, we
533 < adopt the S parameters from Luedtke and Landman\cite{landman:1998} and
534 < modify the parameters for the CTS atom to maintain charge neutrality
535 < in the molecule.  Note that the model choice (UA or AA) for the capping
536 < agent can be different from the solvent. Regardless of model choice,
537 < the force field parameters for interactions between capping agent and
538 < solvent can be derived using Lorentz-Berthelot Mixing Rule:
539 < \begin{eqnarray}
540 <  \sigma_{ij} & = & \frac{1}{2} \left(\sigma_{ii} + \sigma_{jj}\right) \\
541 <  \epsilon_{ij} & = & \sqrt{\epsilon_{ii}\epsilon_{jj}}
542 < \end{eqnarray}
543 <
544 < To describe the interactions between metal (Au) and non-metal atoms,
545 < we refer to an adsorption study of alkyl thiols on gold surfaces by
546 < Vlugt {\it et al.}\cite{vlugt:cpc2007154} They fitted an effective
547 < Lennard-Jones form of potential parameters for the interaction between
548 < Au and pseudo-atoms CH$_x$ and S based on a well-established and
549 < widely-used effective potential of Hautman and Klein for the Au(111)
550 < surface.\cite{hautman:4994} As our simulations require the gold slab
551 < to be flexible to accommodate thermal excitation, the pair-wise form
552 < of potentials they developed was used for our study.
553 <
554 < The potentials developed from {\it ab initio} calculations by Leng
555 < {\it et al.}\cite{doi:10.1021/jp034405s} are adopted for the
556 < interactions between Au and aromatic C/H atoms in toluene. However,
557 < the Lennard-Jones parameters between Au and other types of particles,
558 < (e.g. AA alkanes) have not yet been established. For these
559 < interactions, the Lorentz-Berthelot mixing rule can be used to derive
560 < effective single-atom LJ parameters for the metal using the fit values
561 < for toluene. These are then used to construct reasonable mixing
562 < parameters for the interactions between the gold and other atoms.
563 < Table 1 in the supporting information summarizes the
564 < ``metal/non-metal'' parameters utilized in our simulations.
565 <
566 < \section{Results}
567 < [L-J COMPARED TO RENMD NIVS; WATER COMPARED TO RNEMD NIVS;
568 < SLIP BOUNDARY VS STICK BOUNDARY; ICE-WATER INTERFACES]
569 <
570 < There are many factors contributing to the measured interfacial
571 < conductance; some of these factors are physically motivated
572 < (e.g. coverage of the surface by the capping agent coverage and
573 < solvent identity), while some are governed by parameters of the
574 < methodology (e.g. applied flux and the formulas used to obtain the
575 < conductance). In this section we discuss the major physical and
576 < calculational effects on the computed conductivity.
577 <
578 < \subsection{Effects due to capping agent coverage}
579 <
580 < A series of different initial conditions with a range of surface
581 < coverages was prepared and solvated with various with both of the
582 < solvent molecules. These systems were then equilibrated and their
583 < interfacial thermal conductivity was measured with the NIVS
584 < algorithm. Figure \ref{coverage} demonstrates the trend of conductance
585 < with respect to surface coverage.
586 <
587 < \begin{figure}
588 < \includegraphics[width=\linewidth]{coverage}
589 < \caption{The interfacial thermal conductivity ($G$) has a
590 <  non-monotonic dependence on the degree of surface capping.  This
591 <  data is for the Au(111) / butanethiol / solvent interface with
592 <  various UA force fields at $\langle T\rangle \sim $200K.}
593 < \label{coverage}
594 < \end{figure}
595 <
596 < In partially covered surfaces, the derivative definition for
597 < $G^\prime$ (Eq. \ref{derivativeG}) becomes difficult to apply, as the
598 < location of maximum change of $\lambda$ becomes washed out.  The
599 < discrete definition (Eq. \ref{discreteG}) is easier to apply, as the
600 < Gibbs dividing surface is still well-defined. Therefore, $G$ (not
601 < $G^\prime$) was used in this section.
602 <
603 < From Figure \ref{coverage}, one can see the significance of the
604 < presence of capping agents. When even a small fraction of the Au(111)
605 < surface sites are covered with butanethiols, the conductivity exhibits
606 < an enhancement by at least a factor of 3.  Capping agents are clearly
607 < playing a major role in thermal transport at metal / organic solvent
608 < surfaces.
391 > In our method, a shear stress can be applied similar to shear
392 > viscosity computations by applying an unphysical momentum flux
393 > (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
394 > shown in Figure \ref{slipLength}, in which the velocity gradients
395 > within liquid phase and velocity difference at the liquid-solid
396 > interface can be measured respectively. Further calculations and
397 > characterizations of the interface can be carried out using these
398 > data.
399  
400 < We note a non-monotonic behavior in the interfacial conductance as a
401 < function of surface coverage. The maximum conductance (largest $G$)
402 < happens when the surfaces are about 75\% covered with butanethiol
403 < caps.  The reason for this behavior is not entirely clear.  One
404 < explanation is that incomplete butanethiol coverage allows small gaps
405 < between butanethiols to form. These gaps can be filled by transient
406 < solvent molecules.  These solvent molecules couple very strongly with
407 < the hot capping agent molecules near the surface, and can then carry
408 < away (diffusively) the excess thermal energy from the surface.
400 > \section{Results and Discussions}
401 > \subsection{Lennard-Jones fluid}
402 > Our orthorhombic simulation cell of Lennard-Jones fluid has identical
403 > parameters to our previous work\cite{kuang:164101} to facilitate
404 > comparison. Thermal conductivitis and shear viscosities were computed
405 > with the algorithm applied to the simulations. The results of thermal
406 > conductivity are compared with our previous NIVS algorithm. However,
407 > since the NIVS algorithm could produce temperature anisotropy for
408 > shear viscocity computations, these results are instead compared to
409 > the momentum swapping approaches. Table \ref{LJ} lists these
410 > calculations with various fluxes in reduced units.
411  
620 There appears to be a competition between the conduction of the
621 thermal energy away from the surface by the capping agents (enhanced
622 by greater coverage) and the coupling of the capping agents with the
623 solvent (enhanced by interdigitation at lower coverages).  This
624 competition would lead to the non-monotonic coverage behavior observed
625 here.
626
627 Results for rigid body toluene solvent, as well as the UA hexane, are
628 within the ranges expected from prior experimental
629 work.\cite{Wilson:2002uq,cahill:793,PhysRevB.80.195406} This suggests
630 that explicit hydrogen atoms might not be required for modeling
631 thermal transport in these systems.  C-H vibrational modes do not see
632 significant excited state population at low temperatures, and are not
633 likely to carry lower frequency excitations from the solid layer into
634 the bulk liquid.
635
636 The toluene solvent does not exhibit the same behavior as hexane in
637 that $G$ remains at approximately the same magnitude when the capping
638 coverage increases from 25\% to 75\%.  Toluene, as a rigid planar
639 molecule, cannot occupy the relatively small gaps between the capping
640 agents as easily as the chain-like {\it n}-hexane.  The effect of
641 solvent coupling to the capping agent is therefore weaker in toluene
642 except at the very lowest coverage levels.  This effect counters the
643 coverage-dependent conduction of heat away from the metal surface,
644 leading to a much flatter $G$ vs. coverage trend than is observed in
645 {\it n}-hexane.
646
647 \subsection{Effects due to Solvent \& Solvent Models}
648 In addition to UA solvent and capping agent models, AA models have
649 also been included in our simulations.  In most of this work, the same
650 (UA or AA) model for solvent and capping agent was used, but it is
651 also possible to utilize different models for different components.
652 We have also included isotopic substitutions (Hydrogen to Deuterium)
653 to decrease the explicit vibrational overlap between solvent and
654 capping agent. Table \ref{modelTest} summarizes the results of these
655 studies.
656
412   \begin{table*}
413    \begin{minipage}{\linewidth}
414      \begin{center}
415 +
416 +      \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
417 +        ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
418 +        ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
419 +        at various momentum fluxes.  The new method yields similar
420 +        results to previous RNEMD methods. All results are reported in
421 +        reduced unit. Uncertainties are indicated in parentheses.}
422        
423 <      \caption{Computed interfacial thermal conductance ($G$ and
662 <        $G^\prime$) values for interfaces using various models for
663 <        solvent and capping agent (or without capping agent) at
664 <        $\langle T\rangle\sim$200K.  Here ``D'' stands for deuterated
665 <        solvent or capping agent molecules. Error estimates are
666 <        indicated in parentheses.}
667 <      
668 <      \begin{tabular}{llccc}
423 >      \begin{tabular}{cccccc}
424          \hline\hline
425 <        Butanethiol model & Solvent & $G$ & $G^\prime$ \\
426 <        (or bare surface) & model & \multicolumn{2}{c}{(MW/m$^2$/K)} \\
425 >        \multicolumn{2}{c}{Momentum Exchange} &
426 >        \multicolumn{2}{c}{$\lambda^*$} &
427 >        \multicolumn{2}{c}{$\eta^*$} \\
428          \hline
429 <        UA    & UA hexane    & 131(9)    & 87(10)    \\
430 <              & UA hexane(D) & 153(5)    & 136(13)   \\
675 <              & AA hexane    & 131(6)    & 122(10)   \\
676 <              & UA toluene   & 187(16)   & 151(11)   \\
677 <              & AA toluene   & 200(36)   & 149(53)   \\
429 >        Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
430 >        NIVS & This work & Swapping & This work \\
431          \hline
432 <        AA    & UA hexane    & 116(9)    & 129(8)    \\
433 <              & AA hexane    & 442(14)   & 356(31)   \\
434 <              & AA hexane(D) & 222(12)   & 234(54)   \\
435 <              & UA toluene   & 125(25)   & 97(60)    \\
436 <              & AA toluene   & 487(56)   & 290(42)   \\
684 <        \hline
685 <        AA(D) & UA hexane    & 158(25)   & 172(4)    \\
686 <              & AA hexane    & 243(29)   & 191(11)   \\
687 <              & AA toluene   & 364(36)   & 322(67)   \\
688 <        \hline
689 <        bare  & UA hexane    & 46.5(3.2) & 49.4(4.5) \\
690 <              & UA hexane(D) & 43.9(4.6) & 43.0(2.0) \\
691 <              & AA hexane    & 31.0(1.4) & 29.4(1.3) \\
692 <              & UA toluene   & 70.1(1.3) & 65.8(0.5) \\
432 >        0.116 & 0.16  & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
433 >        0.232 & 0.09  & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
434 >        0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
435 >        0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
436 >        1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
437          \hline\hline
438        \end{tabular}
439 <      \label{modelTest}
439 >      \label{LJ}
440      \end{center}
441    \end{minipage}
442   \end{table*}
443  
444 < To facilitate direct comparison between force fields, systems with the
445 < same capping agent and solvent were prepared with the same length
446 < scales for the simulation cells.
444 > \subsubsection{Thermal conductivity}
445 > Our thermal conductivity calculations with this method yields
446 > comparable results to the previous NIVS algorithm. This indicates that
447 > the thermal gradients rendered using this method are also close to
448 > previous RNEMD methods. Simulations with moderately higher thermal
449 > fluxes tend to yield more reliable thermal gradients and thus avoid
450 > large errors, while overly high thermal fluxes could introduce side
451 > effects such as non-linear temperature gradient response or
452 > inadvertent phase transitions.
453  
454 < On bare metal / solvent surfaces, different force field models for
455 < hexane yield similar results for both $G$ and $G^\prime$, and these
456 < two definitions agree with each other very well. This is primarily an
457 < indicator of weak interactions between the metal and the solvent.
454 > Since the scaling operation is isotropic in this method, one does not
455 > need extra care to ensure temperature isotropy between the $x$, $y$
456 > and $z$ axes, while thermal anisotropy might happen if the criteria
457 > function for choosing scaling coefficients does not perform as
458 > expected. Furthermore, this method avoids inadvertent concomitant
459 > momentum flux when only thermal flux is imposed, which could not be
460 > achieved with swapping or NIVS approaches. The thermal energy exchange
461 > in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
462 > or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
463 > P^\alpha$) would not obtain this result unless thermal flux vanishes
464 > (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which are trivial to apply a
465 > thermal flux). In this sense, this method contributes to having
466 > minimal perturbation to a simulation while imposing thermal flux.
467  
468 < For the fully-covered surfaces, the choice of force field for the
469 < capping agent and solvent has a large impact on the calculated values
470 < of $G$ and $G^\prime$.  The AA thiol to AA solvent conductivities are
471 < much larger than their UA to UA counterparts, and these values exceed
472 < the experimental estimates by a large measure.  The AA force field
473 < allows significant energy to go into C-H (or C-D) stretching modes,
474 < and since these modes are high frequency, this non-quantum behavior is
475 < likely responsible for the overestimate of the conductivity.  Compared
476 < to the AA model, the UA model yields more reasonable conductivity
718 < values with much higher computational efficiency.
719 <
720 < \subsubsection{Are electronic excitations in the metal important?}
721 < Because they lack electronic excitations, the QSC and related embedded
722 < atom method (EAM) models for gold are known to predict unreasonably
723 < low values for bulk conductivity
724 < ($\lambda$).\cite{kuang:164101,ISI:000207079300006,Clancy:1992} If the
725 < conductance between the phases ($G$) is governed primarily by phonon
726 < excitation (and not electronic degrees of freedom), one would expect a
727 < classical model to capture most of the interfacial thermal
728 < conductance.  Our results for $G$ and $G^\prime$ indicate that this is
729 < indeed the case, and suggest that the modeling of interfacial thermal
730 < transport depends primarily on the description of the interactions
731 < between the various components at the interface.  When the metal is
732 < chemically capped, the primary barrier to thermal conductivity appears
733 < to be the interface between the capping agent and the surrounding
734 < solvent, so the excitations in the metal have little impact on the
735 < value of $G$.
736 <
737 < \subsection{Effects due to methodology and simulation parameters}
738 <
739 < We have varied the parameters of the simulations in order to
740 < investigate how these factors would affect the computation of $G$.  Of
741 < particular interest are: 1) the length scale for the applied thermal
742 < gradient (modified by increasing the amount of solvent in the system),
743 < 2) the sign and magnitude of the applied thermal flux, 3) the average
744 < temperature of the simulation (which alters the solvent density during
745 < equilibration), and 4) the definition of the interfacial conductance
746 < (Eqs. (\ref{discreteG}) or (\ref{derivativeG})) used in the
747 < calculation.
748 <
749 < Systems of different lengths were prepared by altering the number of
750 < solvent molecules and extending the length of the box along the $z$
751 < axis to accomodate the extra solvent.  Equilibration at the same
752 < temperature and pressure conditions led to nearly identical surface
753 < areas ($L_x$ and $L_y$) available to the metal and capping agent,
754 < while the extra solvent served mainly to lengthen the axis that was
755 < used to apply the thermal flux.  For a given value of the applied
756 < flux, the different $z$ length scale has only a weak effect on the
757 < computed conductivities.
758 <
759 < \subsubsection{Effects of applied flux}
760 < The NIVS algorithm allows changes in both the sign and magnitude of
761 < the applied flux.  It is possible to reverse the direction of heat
762 < flow simply by changing the sign of the flux, and thermal gradients
763 < which would be difficult to obtain experimentally ($5$ K/\AA) can be
764 < easily simulated.  However, the magnitude of the applied flux is not
765 < arbitrary if one aims to obtain a stable and reliable thermal gradient.
766 < A temperature gradient can be lost in the noise if $|J_z|$ is too
767 < small, and excessive $|J_z|$ values can cause phase transitions if the
768 < extremes of the simulation cell become widely separated in
769 < temperature.  Also, if $|J_z|$ is too large for the bulk conductivity
770 < of the materials, the thermal gradient will never reach a stable
771 < state.  
772 <
773 < Within a reasonable range of $J_z$ values, we were able to study how
774 < $G$ changes as a function of this flux.  In what follows, we use
775 < positive $J_z$ values to denote the case where energy is being
776 < transferred by the method from the metal phase and into the liquid.
777 < The resulting gradient therefore has a higher temperature in the
778 < liquid phase.  Negative flux values reverse this transfer, and result
779 < in higher temperature metal phases.  The conductance measured under
780 < different applied $J_z$ values is listed in Tables 2 and 3 in the
781 < supporting information. These results do not indicate that $G$ depends
782 < strongly on $J_z$ within this flux range. The linear response of flux
783 < to thermal gradient simplifies our investigations in that we can rely
784 < on $G$ measurement with only a small number $J_z$ values.
785 <
786 < The sign of $J_z$ is a different matter, however, as this can alter
787 < the temperature on the two sides of the interface. The average
788 < temperature values reported are for the entire system, and not for the
789 < liquid phase, so at a given $\langle T \rangle$, the system with
790 < positive $J_z$ has a warmer liquid phase.  This means that if the
791 < liquid carries thermal energy via diffusive transport, {\it positive}
792 < $J_z$ values will result in increased molecular motion on the liquid
793 < side of the interface, and this will increase the measured
794 < conductivity.
795 <
796 < \subsubsection{Effects due to average temperature}
797 <
798 < We also studied the effect of average system temperature on the
799 < interfacial conductance.  The simulations are first equilibrated in
800 < the NPT ensemble at 1 atm.  The TraPPE-UA model for hexane tends to
801 < predict a lower boiling point (and liquid state density) than
802 < experiments.  This lower-density liquid phase leads to reduced contact
803 < between the hexane and butanethiol, and this accounts for our
804 < observation of lower conductance at higher temperatures.  In raising
805 < the average temperature from 200K to 250K, the density drop of
806 < $\sim$20\% in the solvent phase leads to a $\sim$40\% drop in the
807 < conductance.
808 <
809 < Similar behavior is observed in the TraPPE-UA model for toluene,
810 < although this model has better agreement with the experimental
811 < densities of toluene.  The expansion of the toluene liquid phase is
812 < not as significant as that of the hexane (8.3\% over 100K), and this
813 < limits the effect to $\sim$20\% drop in thermal conductivity.
814 <
815 < Although we have not mapped out the behavior at a large number of
816 < temperatures, is clear that there will be a strong temperature
817 < dependence in the interfacial conductance when the physical properties
818 < of one side of the interface (notably the density) change rapidly as a
819 < function of temperature.
820 <
821 < Besides the lower interfacial thermal conductance, surfaces at
822 < relatively high temperatures are susceptible to reconstructions,
823 < particularly when butanethiols fully cover the Au(111) surface. These
824 < reconstructions include surface Au atoms which migrate outward to the
825 < S atom layer, and butanethiol molecules which embed into the surface
826 < Au layer. The driving force for this behavior is the strong Au-S
827 < interactions which are modeled here with a deep Lennard-Jones
828 < potential. This phenomenon agrees with reconstructions that have been
829 < experimentally
830 < observed.\cite{doi:10.1021/j100035a033,doi:10.1021/la026493y}.  Vlugt
831 < {\it et al.} kept their Au(111) slab rigid so that their simulations
832 < could reach 300K without surface
833 < reconstructions.\cite{vlugt:cpc2007154} Since surface reconstructions
834 < blur the interface, the measurement of $G$ becomes more difficult to
835 < conduct at higher temperatures.  For this reason, most of our
836 < measurements are undertaken at $\langle T\rangle\sim$200K where
837 < reconstruction is minimized.
838 <
839 < However, when the surface is not completely covered by butanethiols,
840 < the simulated system appears to be more resistent to the
841 < reconstruction. Our Au / butanethiol / toluene system had the Au(111)
842 < surfaces 90\% covered by butanethiols, but did not see this above
843 < phenomena even at $\langle T\rangle\sim$300K.  That said, we did
844 < observe butanethiols migrating to neighboring three-fold sites during
845 < a simulation.  Since the interface persisted in these simulations, we
846 < were able to obtain $G$'s for these interfaces even at a relatively
847 < high temperature without being affected by surface reconstructions.
848 <
849 < \section{Discussion}
850 < [COMBINE W. RESULTS]
851 < The primary result of this work is that the capping agent acts as an
852 < efficient thermal coupler between solid and solvent phases.  One of
853 < the ways the capping agent can carry out this role is to down-shift
854 < between the phonon vibrations in the solid (which carry the heat from
855 < the gold) and the molecular vibrations in the liquid (which carry some
856 < of the heat in the solvent).
857 <
858 < To investigate the mechanism of interfacial thermal conductance, the
859 < vibrational power spectrum was computed. Power spectra were taken for
860 < individual components in different simulations. To obtain these
861 < spectra, simulations were run after equilibration in the
862 < microcanonical (NVE) ensemble and without a thermal
863 < gradient. Snapshots of configurations were collected at a frequency
864 < that is higher than that of the fastest vibrations occurring in the
865 < simulations. With these configurations, the velocity auto-correlation
866 < functions can be computed:
867 < \begin{equation}
868 < C_A (t) = \langle\vec{v}_A (t)\cdot\vec{v}_A (0)\rangle
869 < \label{vCorr}
870 < \end{equation}
871 < The power spectrum is constructed via a Fourier transform of the
872 < symmetrized velocity autocorrelation function,
873 < \begin{equation}
874 <  \hat{f}(\omega) =
875 <  \int_{-\infty}^{\infty} C_A (t) e^{-2\pi it\omega}\,dt
876 < \label{fourier}
877 < \end{equation}
878 <
879 < \subsection{The role of specific vibrations}
880 < The vibrational spectra for gold slabs in different environments are
881 < shown as in Figure \ref{specAu}. Regardless of the presence of
882 < solvent, the gold surfaces which are covered by butanethiol molecules
883 < exhibit an additional peak observed at a frequency of
884 < $\sim$165cm$^{-1}$.  We attribute this peak to the S-Au bonding
885 < vibration. This vibration enables efficient thermal coupling of the
886 < surface Au layer to the capping agents. Therefore, in our simulations,
887 < the Au / S interfaces do not appear to be the primary barrier to
888 < thermal transport when compared with the butanethiol / solvent
889 < interfaces.  This supports the results of Luo {\it et
890 <  al.}\cite{Luo20101}, who reported $G$ for Au-SAM junctions roughly
891 < twice as large as what we have computed for the thiol-liquid
892 < interfaces.
893 <
894 < \begin{figure}
895 < \includegraphics[width=\linewidth]{vibration}
896 < \caption{The vibrational power spectrum for thiol-capped gold has an
897 <  additional vibrational peak at $\sim $165cm$^{-1}$.  Bare gold
898 <  surfaces (both with and without a solvent over-layer) are missing
899 <  this peak.   A similar peak at  $\sim $165cm$^{-1}$ also appears in
900 <  the vibrational power spectrum for the butanethiol capping agents.}
901 < \label{specAu}
902 < \end{figure}
903 <
904 < Also in this figure, we show the vibrational power spectrum for the
905 < bound butanethiol molecules, which also exhibits the same
906 < $\sim$165cm$^{-1}$ peak.
907 <
908 < \subsection{Overlap of power spectra}
909 < A comparison of the results obtained from the two different organic
910 < solvents can also provide useful information of the interfacial
911 < thermal transport process.  In particular, the vibrational overlap
912 < between the butanethiol and the organic solvents suggests a highly
913 < efficient thermal exchange between these components.  Very high
914 < thermal conductivity was observed when AA models were used and C-H
915 < vibrations were treated classically. The presence of extra degrees of
916 < freedom in the AA force field yields higher heat exchange rates
917 < between the two phases and results in a much higher conductivity than
918 < in the UA force field. The all-atom classical models include high
919 < frequency modes which should be unpopulated at our relatively low
920 < temperatures.  This artifact is likely the cause of the high thermal
921 < conductance in all-atom MD simulations.
468 > \subsubsection{Shear viscosity}
469 > Table \ref{LJ} also compares our shear viscosity results with momentum
470 > swapping approach. Our calculations show that our method predicted
471 > similar values for shear viscosities to the momentum swapping
472 > approach, as well as the velocity gradient profiles. Moderately larger
473 > momentum fluxes are helpful to reduce the errors of measured velocity
474 > gradients and thus the final result. However, it is pointed out that
475 > the momentum swapping approach tends to produce nonthermal velocity
476 > distributions.\cite{Maginn:2010}
477  
478 < The similarity in the vibrational modes available to solvent and
479 < capping agent can be reduced by deuterating one of the two components
480 < (Fig. \ref{aahxntln}).  Once either the hexanes or the butanethiols
481 < are deuterated, one can observe a significantly lower $G$ and
482 < $G^\prime$ values (Table \ref{modelTest}).
478 > To examine that temperature isotropy holds in simulations using our
479 > method, we measured the three one-dimensional temperatures in each of
480 > the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
481 > temperatures were calculated after subtracting the effects from bulk
482 > velocities of the slabs. The one-dimensional temperature profiles
483 > showed no observable difference between the three dimensions. This
484 > ensures that isotropic scaling automatically preserves temperature
485 > isotropy and that our method is useful in shear viscosity
486 > computations.
487  
488   \begin{figure}
489 < \includegraphics[width=\linewidth]{aahxntln}
490 < \caption{Spectra obtained for all-atom (AA) Au / butanethiol / solvent
491 <  systems. When butanethiol is deuterated (lower left), its
492 <  vibrational overlap with hexane decreases significantly.  Since
493 <  aromatic molecules and the butanethiol are vibrationally dissimilar,
494 <  the change is not as dramatic when toluene is the solvent (right).}
495 < \label{aahxntln}
489 > \includegraphics[width=\linewidth]{tempXyz}
490 > \caption{Unlike the previous NIVS algorithm, the new method does not
491 >  produce a thermal anisotropy. No temperature difference between
492 >  different dimensions were observed beyond the magnitude of the error
493 >  bars. Note that the two ``hotter'' regions are caused by the shear
494 >  stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
495 >  an effect that only observed in our methods.}
496 > \label{tempXyz}
497   \end{figure}
498  
499 < For the Au / butanethiol / toluene interfaces, having the AA
500 < butanethiol deuterated did not yield a significant change in the
501 < measured conductance. Compared to the C-H vibrational overlap between
502 < hexane and butanethiol, both of which have alkyl chains, the overlap
503 < between toluene and butanethiol is not as significant and thus does
504 < not contribute as much to the heat exchange process.
499 > Furthermore, the velocity distribution profiles are tested by imposing
500 > a large shear stress into the simulations. Figure \ref{vDist}
501 > demonstrates how our method is able to maintain thermal velocity
502 > distributions against the momentum swapping approach even under large
503 > imposed fluxes. Previous swapping methods tend to deplete particles of
504 > positive velocities in the negative velocity slab (``c'') and vice
505 > versa in slab ``h'', where the distributions leave a notch. This
506 > problematic profiles become significant when the imposed-flux becomes
507 > larger and diffusions from neighboring slabs could not offset the
508 > depletion. Simutaneously, abnormal peaks appear corresponding to
509 > excessive velocity swapped from the other slab. This nonthermal
510 > distributions limit applications of the swapping approach in shear
511 > stress simulations. Our method avoids the above problematic
512 > distributions by altering the means of applying momentum
513 > fluxes. Comparatively, velocity distributions recorded from
514 > simulations with our method is so close to the ideal thermal
515 > prediction that no observable difference is shown in Figure
516 > \ref{vDist}. Conclusively, our method avoids problems happened in
517 > previous RNEMD methods and provides a useful means for shear viscosity
518 > computations.
519  
946 Previous observations of Zhang {\it et al.}\cite{hase:2010} indicate
947 that the {\it intra}molecular heat transport due to alkylthiols is
948 highly efficient.  Combining our observations with those of Zhang {\it
949  et al.}, it appears that butanethiol acts as a channel to expedite
950 heat flow from the gold surface and into the alkyl chain.  The
951 vibrational coupling between the metal and the liquid phase can
952 therefore be enhanced with the presence of suitable capping agents.
953
954 Deuterated models in the UA force field did not decouple the thermal
955 transport as well as in the AA force field.  The UA models, even
956 though they have eliminated the high frequency C-H vibrational
957 overlap, still have significant overlap in the lower-frequency
958 portions of the infrared spectra (Figure \ref{uahxnua}).  Deuterating
959 the UA models did not decouple the low frequency region enough to
960 produce an observable difference for the results of $G$ (Table
961 \ref{modelTest}).
962
520   \begin{figure}
521 < \includegraphics[width=\linewidth]{uahxnua}
522 < \caption{Vibrational power spectra for UA models for the butanethiol
523 <  and hexane solvent (upper panel) show the high degree of overlap
524 <  between these two molecules, particularly at lower frequencies.
525 <  Deuterating a UA model for the solvent (lower panel) does not
526 <  decouple the two spectra to the same degree as in the AA force
527 <  field (see Fig \ref{aahxntln}).}
528 < \label{uahxnua}
521 > \includegraphics[width=\linewidth]{velDist}
522 > \caption{Velocity distributions that develop under the swapping and
523 >  our methods at high flux. These distributions were obtained from
524 >  Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
525 >  swapping interval of 20 time steps). This is a relatively large flux
526 >  to demonstrate the nonthermal distributions that develop under the
527 >  swapping method. Distributions produced by our method are very close
528 >  to the ideal thermal situations.}
529 > \label{vDist}
530   \end{figure}
531  
532 < \section{Conclusions}
533 < The NIVS algorithm has been applied to simulations of
534 < butanethiol-capped Au(111) surfaces in the presence of organic
535 < solvents. This algorithm allows the application of unphysical thermal
536 < flux to transfer heat between the metal and the liquid phase. With the
537 < flux applied, we were able to measure the corresponding thermal
538 < gradients and to obtain interfacial thermal conductivities. Under
539 < steady states, 2-3 ns trajectory simulations are sufficient for
540 < computation of this quantity.
532 > \subsection{Bulk SPC/E water}
533 > Since our method was in good performance of thermal conductivity and
534 > shear viscosity computations for simple Lennard-Jones fluid, we extend
535 > our applications of these simulations to complex fluid like SPC/E
536 > water model. A simulation cell with 1000 molecules was set up in the
537 > same manner as in \cite{kuang:164101}. For thermal conductivity
538 > simulations, measurements were taken to compare with previous RNEMD
539 > methods; for shear viscosity computations, simulations were run under
540 > a series of temperatures (with corresponding pressure relaxation using
541 > the isobaric-isothermal ensemble[CITE NIVS REF 32]), and results were
542 > compared to available data from Equilibrium MD methods[CITATIONS].
543  
544 < Our simulations have seen significant conductance enhancement in the
545 < presence of capping agent, compared with the bare gold / liquid
546 < interfaces. The vibrational coupling between the metal and the liquid
547 < phase is enhanced by a chemically-bonded capping agent. Furthermore,
548 < the coverage percentage of the capping agent plays an important role
549 < in the interfacial thermal transport process. Moderately low coverages
550 < allow higher contact between capping agent and solvent, and thus could
551 < further enhance the heat transfer process, giving a non-monotonic
992 < behavior of conductance with increasing coverage.
544 > \subsubsection{Thermal conductivity}
545 > Table \ref{spceThermal} summarizes our thermal conductivity
546 > computations under different temperatures and thermal gradients, in
547 > comparison to the previous NIVS results\cite{kuang:164101} and
548 > experimental measurements\cite{WagnerKruse}. Note that no appreciable
549 > drift of total system energy or temperature was observed when our
550 > method is applied, which indicates that our algorithm conserves total
551 > energy even for systems involving electrostatic interactions.
552  
553 < Our results, particularly using the UA models, agree well with
554 < available experimental data.  The AA models tend to overestimate the
555 < interfacial thermal conductance in that the classically treated C-H
556 < vibrations become too easily populated. Compared to the AA models, the
557 < UA models have higher computational efficiency with satisfactory
558 < accuracy, and thus are preferable in modeling interfacial thermal
559 < transport.
553 > Measurements using our method established similar temperature
554 > gradients to the previous NIVS method. Our simulation results are in
555 > good agreement with those from previous simulations. And both methods
556 > yield values in reasonable agreement with experimental
557 > values. Simulations using moderately higher thermal gradient or those
558 > with longer gradient axis ($z$) for measurement seem to have better
559 > accuracy, from our results.
560  
561 < Of the two definitions for $G$, the discrete form
562 < (Eq. \ref{discreteG}) was easier to use and gives out relatively
563 < consistent results, while the derivative form (Eq. \ref{derivativeG})
564 < is not as versatile. Although $G^\prime$ gives out comparable results
565 < and follows similar trend with $G$ when measuring close to fully
566 < covered or bare surfaces, the spatial resolution of $T$ profile
567 < required for the use of a derivative form is limited by the number of
568 < bins and the sampling required to obtain thermal gradient information.
561 > \begin{table*}
562 >  \begin{minipage}{\linewidth}
563 >    \begin{center}
564 >      
565 >      \caption{Thermal conductivity of SPC/E water under various
566 >        imposed thermal gradients. Uncertainties are indicated in
567 >        parentheses.}
568 >      
569 >      \begin{tabular}{ccccc}
570 >        \hline\hline
571 >        $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
572 >        {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
573 >        (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
574 >        Experiment\cite{WagnerKruse} \\
575 >        \hline
576 >        300 & 0.8 & 0.815(0.027)     & 0.770(0.008) & 0.61 \\
577 >        318 & 0.8 & 0.801(0.024)     & 0.750(0.032) & 0.64 \\
578 >            & 1.6 & 0.766(0.007)     & 0.778(0.019) &      \\
579 >            & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
580 >                     twice as long.} &              &      \\
581 >        \hline\hline
582 >      \end{tabular}
583 >      \label{spceThermal}
584 >    \end{center}
585 >  \end{minipage}
586 > \end{table*}
587  
588 < Vlugt {\it et al.} have investigated the surface thiol structures for
589 < nanocrystalline gold and pointed out that they differ from those of
590 < the Au(111) surface.\cite{landman:1998,vlugt:cpc2007154} This
591 < difference could also cause differences in the interfacial thermal
592 < transport behavior. To investigate this problem, one would need an
593 < effective method for applying thermal gradients in non-planar
594 < (i.e. spherical) geometries.
588 > \subsubsection{Shear viscosity}
589 > The improvement our method achieves for shear viscosity computations
590 > enables us to apply it on SPC/E water models. The series of
591 > temperatures under which our shear viscosity calculations were carried
592 > out covers the liquid range under normal pressure. Our simulations
593 > predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
594 > (Table \ref{spceShear}). Considering subtlties such as temperature or
595 > pressure/density errors in these two series of measurements, our
596 > results show no significant difference from those with EMD
597 > methods. Since each value reported using our method takes only one
598 > single trajectory of simulation, instead of average from many
599 > trajectories when using EMD, our method provides an effective means
600 > for shear viscosity computations.
601  
602 + \begin{table*}
603 +  \begin{minipage}{\linewidth}
604 +    \begin{center}
605 +      
606 +      \caption{Computed shear viscosity of SPC/E water under different
607 +        temperatures. Results are compared to those obtained with EMD
608 +        method[CITATION]. Uncertainties are indicated in parentheses.}
609 +      
610 +      \begin{tabular}{cccc}
611 +        \hline\hline
612 +        $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
613 +        {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
614 +        (K) & (10$^{10}$s$^{-1}$) & This work & Previous simulations[CITATION]\\
615 +        \hline
616 +        273 &  & 1.218(0.004) &  \\
617 +            &  & 1.140(0.012) &  \\
618 +        303 &  & 0.646(0.008) &  \\
619 +        318 &  & 0.536(0.007) &  \\
620 +            &  & 0.510(0.007) &  \\
621 +            &  &  &  \\
622 +        333 &  & 0.428(0.002) &  \\
623 +        363 &  & 0.279(0.014) &  \\
624 +            &  & 0.306(0.001) &  \\
625 +        \hline\hline
626 +      \end{tabular}
627 +      \label{spceShear}
628 +    \end{center}
629 +  \end{minipage}
630 + \end{table*}
631 +
632 + [MAY COMBINE JZPX AND JZKE TO RUN ONE JOB BUT GET ETA=F(T)]
633 + [PUT RESULTS AND FIGURE HERE IF IT WORKS]
634 + \subsection{Interfacial frictions and slip lengths}
635 + An attractive aspect of our method is the ability to apply momentum
636 + and/or thermal flux in nonhomogeneous systems, where molecules of
637 + different identities (or phases) are segregated in different
638 + regions. We have previously studied the interfacial thermal transport
639 + of a series of metal gold-liquid
640 + surfaces\cite{kuang:164101,kuang:AuThl}, and attemptions have been
641 + made to investigate the relationship between this phenomenon and the
642 + interfacial frictions.
643 +
644 + Table \ref{etaKappaDelta} includes these computations and previous
645 + calculations of corresponding interfacial thermal conductance. For
646 + bare Au(111) surfaces, slip boundary conditions were observed for both
647 + organic and aqueous liquid phases, corresponding to previously
648 + computed low interfacial thermal conductance. Instead, the butanethiol
649 + covered Au(111) surface appeared to be sticky to the organic liquid
650 + molecules in our simulations. We have reported conductance enhancement
651 + effect for this surface capping agent,\cite{kuang:AuThl} and these
652 + observations have a qualitative agreement with the thermal conductance
653 + results. This agreement also supports discussions on the relationship
654 + between surface wetting and slip effect and thermal conductance of the
655 + interface.[CITE BARRAT, GARDE]
656 +
657 + \begin{table*}
658 +  \begin{minipage}{\linewidth}
659 +    \begin{center}
660 +      
661 +      \caption{Computed interfacial friction coefficient values for
662 +        interfaces with various components for liquid and solid
663 +        phase. Error estimates are indicated in parentheses.}
664 +      
665 +      \begin{tabular}{llcccccc}
666 +        \hline\hline
667 +        Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
668 +        & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
669 +          \cite{kuang:164101}.} \\
670 +        surface & model & K & MPa & mPa$\cdot$s & Pa$\cdot$s/m & nm &
671 +        MW/m$^2$/K \\
672 +        \hline
673 +        Au(111) & hexane & 200 & 1.08 & 0.20() & 5.3$\times$10$^4$() &
674 +        3.7 & 46.5 \\
675 +                &        &     & 2.15 & 0.14() & 5.3$\times$10$^4$() &
676 +        2.7 &      \\
677 +        Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.29() & $\infty$ & 0 &
678 +        131 \\
679 +                       &        &     & 5.39 & 0.32() & $\infty$ & 0 &
680 +            \\
681 +        \hline
682 +        Au(111) & toluene & 200 & 1.08 & 0.72() & 1.?$\times$10$^5$() &
683 +        4.6 & 70.1 \\
684 +                &         &     & 2.16 & 0.54() & 1.?$\times$10$^5$() &
685 +        4.9 &      \\
686 +        Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.98() & $\infty$ & 0
687 +        & 187 \\
688 +                       &         &     & 10.8 & 0.99() & $\infty$ & 0
689 +        &     \\
690 +        \hline
691 +        Au(111) & water & 300 & 1.08 & 0.40() & 1.9$\times$10$^4$() &
692 +        20.7 & 1.65 \\
693 +                &       &     & 2.16 & 0.79() & 1.9$\times$10$^4$() &
694 +        41.9 &      \\
695 +        \hline
696 +        ice(basal) & water & 225 & 19.4 & 15.8() & $\infty$ & 0 & \\
697 +        \hline\hline
698 +      \end{tabular}
699 +      \label{etaKappaDelta}
700 +    \end{center}
701 +  \end{minipage}
702 + \end{table*}
703 +
704 + An interesting effect alongside the surface friction change is
705 + observed on the shear viscosity of liquids in the regions close to the
706 + solid surface. Note that $\eta$ measured near a ``slip'' surface tends
707 + to be smaller than that near a ``stick'' surface. This suggests that
708 + an interface could affect the dynamic properties on its neighbor
709 + regions. It is known that diffusions of solid particles in liquid
710 + phase is affected by their surface conditions (stick or slip
711 + boundary).[CITE SCHMIDT AND SKINNER] Our observations could provide
712 + support to this phenomenon.
713 +
714 + In addition to these previously studied interfaces, we attempt to
715 + construct ice-water interfaces and the basal plane of ice lattice was
716 + first studied. In contrast to the Au(111)/water interface, where the
717 + friction coefficient is relatively small and large slip effect
718 + presents, the ice/liquid water interface demonstrates strong
719 + interactions and appears to be sticky. The supercooled liquid phase is
720 + an order of magnitude viscous than measurements in previous
721 + section. It would be of interst to investigate the effect of different
722 + ice lattice planes (such as prism surface) on interfacial friction and
723 + corresponding liquid viscosity.
724 +
725 + \section{Conclusions}
726 + Our simulations demonstrate the validity of our method in RNEMD
727 + computations of thermal conductivity and shear viscosity in atomic and
728 + molecular liquids. Our method maintains thermal velocity distributions
729 + and avoids thermal anisotropy in previous NIVS shear stress
730 + simulations, as well as retains attractive features of previous RNEMD
731 + methods. There is no {\it a priori} restrictions to the method to be
732 + applied in various ensembles, so prospective applications to
733 + extended-system methods are possible.
734 +
735 + Furthermore, using this method, investigations can be carried out to
736 + characterize interfacial interactions. Our method is capable of
737 + effectively imposing both thermal and momentum flux accross an
738 + interface and thus facilitates studies that relates dynamic property
739 + measurements to the chemical details of an interface.
740 +
741 + Another attractive feature of our method is the ability of
742 + simultaneously imposing thermal and momentum flux in a
743 + system. potential researches that might be benefit include complex
744 + systems that involve thermal and momentum gradients. For example, the
745 + Soret effects under a velocity gradient would be of interest to
746 + purification and separation researches.
747 +
748   \section{Acknowledgments}
749   Support for this project was provided by the National Science
750   Foundation under grant CHE-0848243. Computational time was provided by
# Line 1028 | Line 757 | Dame.
757  
758   \end{doublespace}
759   \end{document}
1031

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines