ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3790
Committed: Thu Dec 15 18:32:10 2011 UTC (12 years, 9 months ago) by skuang
Content type: application/x-tex
File size: 43903 byte(s)
Log Message:
some edits

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 gezelter 3786 \title{Velocity Shearing and Scaling RNEMD: a minimally perturbing
32 gezelter 3788 method for simulating temperature and momentum gradients}
33 gezelter 3769
34     \author{Shenyu Kuang and J. Daniel
35     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
36     Department of Chemistry and Biochemistry,\\
37     University of Notre Dame\\
38     Notre Dame, Indiana 46556}
39    
40     \date{\today}
41    
42     \maketitle
43    
44     \begin{doublespace}
45    
46     \begin{abstract}
47 skuang 3779 We present a new method for introducing stable nonequilibrium
48     velocity and temperature gradients in molecular dynamics simulations
49 gezelter 3788 of heterogeneous and interfacial systems. This method conserves both
50     the linear momentum and total energy of the system and improves
51     previous reverse non-equilibrium molecular dynamics (RNEMD) methods
52     while retaining equilibrium thermal velocity distributions in each
53     region of the system. The new method avoids the thermal anisotropy
54 gezelter 3786 produced by previous methods by using isotropic velocity scaling and
55 gezelter 3788 shearing (VSS) on all of the molecules in specific regions. To test
56     the method, we have computed the thermal conductivity and shear
57 skuang 3785 viscosity of model liquid systems as well as the interfacial
58 gezelter 3786 friction coeefficients of a series of solid / liquid interfaces. The
59     method's ability to combine the thermal and momentum gradients
60     allows us to obtain shear viscosity data for a range of temperatures
61     from a single trajectory.
62 gezelter 3769 \end{abstract}
63    
64     \newpage
65    
66     %\narrowtext
67    
68     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69     % BODY OF TEXT
70     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71    
72     \section{Introduction}
73 gezelter 3786 One of the standard ways to compute transport coefficients such as the
74     viscosities and thermal conductivities of liquids is to use
75     imposed-flux non-equilibrium molecular dynamics
76     methods.\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
77     These methods establish stable non-equilibrium conditions in a
78     simulation box using an applied momentum or thermal flux between
79     different regions of the box. The corresponding temperature or
80     velocity gradients which develop in response to the applied flux is
81     related (via linear response theory) to the transport coefficient of
82     interest. These methods are quite efficient, in that they do not need
83     many trajectories to provide information about transport properties. To
84     date, they have been utilized in computing thermal and mechanical
85     transfer of both homogeneous liquids as well as heterogeneous systems
86     such as solid-liquid
87 skuang 3785 interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
88 skuang 3770
89 gezelter 3786 The reverse non-equilibrium molecular dynamics (RNEMD) methods utilize
90     additional constraints that ensure conservation of linear momentum and
91     total energy of the system while imposing the desired flux. The RNEMD
92     methods are therefore capable of sampling various thermodynamically
93     relevent ensembles, including the micro-canonical (NVE) ensemble,
94     without resorting to an external thermostat. The original approaches
95     proposed by M\"{u}ller-Plathe {\it et
96 skuang 3771 al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
97 gezelter 3786 momentum swapping moves for generating energy/momentum fluxes. The
98     swapping moves can also be made compatible with particles of different
99     identities. Although the swapping moves are simple to implement in
100     molecular simulations, Tenney and Maginn have shown that they create
101     nonthermal velocity distributions.\cite{Maginn:2010} Furthermore, this
102     approach is not particularly efficient for kinetic energy transfer
103     between particles of different identities, particularly when the mass
104     difference between the particles becomes significant. This problem
105     makes applying swapping-move RNEMD methods on heterogeneous
106     interfacial systems somewhat difficult.
107 skuang 3770
108 gezelter 3786 Recently, we developed a somewhat different approach to applying
109 gezelter 3788 thermal fluxes in RNEMD simulation using a non-isotropic velocity
110     scaling (NIVS) algorithm.\cite{kuang:164101} This algorithm scales all
111 gezelter 3786 atomic velocity vectors in two separate regions of a simulated system
112     using two diagonal scaling matrices. The scaling matrices are
113     determined by solving single quartic equation which includes linear
114     momentum and kinetic energy conservation constraints as well as the
115     target thermal flux between the regions. The NIVS method is able to
116     effectively impose a wide range of kinetic energy fluxes without
117     significant perturbation to the velocity distributions away from ideal
118     Maxwell-Boltzmann distributions, even in the presence of heterogeneous
119     interfaces. We successfully applied this approach in studying the
120     interfacial thermal conductance at chemically-capped metal-solvent
121 skuang 3771 interfaces.\cite{kuang:AuThl}
122 gezelter 3769
123 gezelter 3786 The NIVS approach works very well for preparing stable {\it thermal}
124     gradients. However, as we pointed out in the original paper, it has
125     limited application in imposing {\it linear} momentum fluxes (which
126     are required for measuring shear viscosities). The reason for this is
127     that linear momentum flux was being imposed by scaling random
128     fluctuations of the center of the velocity distributions. Repeated
129     application of the original NIVS approach resulted in temperature
130     anisotropy, i.e. the width of the velocity distributions depended on
131     coordinates perpendicular to the desired gradient direction. For this
132     reason, combining thermal and momentum fluxes was difficult with the
133     original NIVS algorithm. However, combinations of thermal and
134     velocity gradients would provide a means to simulate thermal-linear
135     coupled processes such as Soret effect in liquid flows. Therefore,
136     developing improved approaches to the scaling imposed-flux methods
137     would be useful.
138 gezelter 3769
139 gezelter 3786 In this paper, we improve the RNEMD methods by introducing a novel
140     approach to creating imposed fluxes. This approach separates the
141     shearing and scaling of the velocity distributions in different
142     spatial regions and can apply both transformations within a single
143     time step. The approach retains desirable features of previous RNEMD
144     approaches and is simpler to implement compared to the earlier NIVS
145 gezelter 3788 method. In what follows, we first present the velocity shearing and
146     scaling (VSS) RNEMD method and its implementation in a
147     simulation. Then we compare the VSS-RNEMD method in bulk fluids to
148     previous methods. We also compute interfacial frictions are computed
149     for a series of heterogeneous interfaces.
150 gezelter 3769
151     \section{Methodology}
152 gezelter 3786 In an approach similar to the earlier NIVS method,\cite{kuang:164101}
153     we consider a periodic system which has been divided into a series of
154     slabs along a single axis (e.g. $z$). The unphysical thermal and
155     momentum fluxes are applied from one of the end slabs to the center
156     slab, and thus the thermal flux produces a higher temperature in the
157     center slab than in the end slab, and the momentum flux results in a
158     center slab moving along the positive $y$ axis (see Fig. \ref{cartoon}).
159     The applied fluxes can be set to negative values if the reverse
160     gradients are desired. For convenience the center slab is denoted as
161     the {\it hot} or {\it ``H''} slab, while the end slab is denoted {\it
162     ``C''} (or {\it cold}).
163 skuang 3770
164 gezelter 3786 \begin{figure}
165     \includegraphics[width=\linewidth]{cartoon}
166 gezelter 3788 \caption{The VSS-RNEMD approach imposes unphysical transfer of both
167     linear momentum and kinetic energy between a ``hot'' slab and a
168     ``cold'' slab in the simulation box. The system responds to this
169     imposed flux by generating both momentum and temperature gradients.
170     The slope of the gradients can then be used to compute transport
171     properties (e.g. shear viscosity and thermal conductivity).}
172 gezelter 3786 \label{cartoon}
173     \end{figure}
174    
175     To impose these fluxes, we periodically apply a set of operations on
176     the velocities of particles {$i$} within the cold slab and a separate
177     operation on particles {$j$} within the hot slab.
178 skuang 3770 \begin{eqnarray}
179     \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
180 gezelter 3788 \rangle\right) + \left(\langle\vec{v}_c\rangle +
181     \vec{a}_c\right) \label{eq:xformc} \\
182 skuang 3770 \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
183 gezelter 3788 \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right) \label{eq:xformh}
184 skuang 3770 \end{eqnarray}
185 gezelter 3786 where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denote
186     the instantaneous average velocity of the molecules within slabs $C$
187     and $H$ respectively. When a momentum flux $\vec{j}_z(\vec{p})$ is
188     present, these slab-averaged velocities also get corresponding
189     incremental changes ($\vec{a}_c$ and $\vec{a}_h$ respectively) that
190     are applied to all particles within each slab. The incremental
191     changes are obtained using Newton's second law:
192 skuang 3770 \begin{eqnarray}
193 gezelter 3786 M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \label{eq:newton1} \\
194 skuang 3770 M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
195 gezelter 3786 \label{eq:newton2}
196 skuang 3770 \end{eqnarray}
197 skuang 3781 where $M$ denotes total mass of particles within a slab:
198 skuang 3770 \begin{eqnarray}
199     M_c & = & \sum_{i = 1}^{N_c} m_i \\
200     M_h & = & \sum_{j = 1}^{N_h} m_j
201     \end{eqnarray}
202 skuang 3781 and $\Delta t$ is the interval between two separate operations.
203 skuang 3770
204 gezelter 3786 The operations in Eqs. \ref{eq:newton1} and \ref{eq:newton2} already
205     conserve the linear momentum of a periodic system. To further satisfy
206     total energy conservation as well as to impose the thermal flux $J_z$,
207     the following constraint equations must be solved for the two scaling
208     variables $c$ and $h$:
209 skuang 3770 \begin{eqnarray}
210     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
211     \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
212     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
213 gezelter 3786 \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2.
214     \label{constraint}
215 skuang 3770 \end{eqnarray}
216 gezelter 3786 Here $K_c$ and $K_h$ denote the translational kinetic energy of slabs
217     $C$ and $H$ respectively. These conservation equations are sufficient
218     to ensure total energy conservation, as the operations applied in our
219     method do not change the kinetic energy related to orientational
220     degrees of freedom or the potential energy of a system (as long as the
221     potential energy is independent of particle velocity).
222 skuang 3770
223 gezelter 3786 Equations \ref{eq:newton1}-\ref{constraint} are sufficient to
224     determine the velocity scaling coefficients ($c$ and $h$) as well as
225     $\vec{a}_c$ and $\vec{a}_h$. Note that there are usually two roots
226     respectively for $c$ and $h$. However, the positive roots (which are
227     closer to 1) are chosen so that the perturbations to the system are
228     minimal. Figure \ref{method} illustrates the implementation of this
229     algorithm in an individual step.
230 skuang 3770
231 skuang 3772 \begin{figure}
232 gezelter 3786 \centering
233     \includegraphics[width=5in]{method}
234 gezelter 3788 \caption{A single step implementation of the VSS algorithm starts with
235     the velocity distributions for the two slabs in a shearing fluid
236     (solid lines). Equations \ref{eq:xformc} and \ref{eq:xformh} are
237     used to scale and shear velocities in both slabs to mimic the effect
238     of both a thermal and a momentum flux. Gaussian distributions are
239 gezelter 3786 preserved by both the scaling and shearing operations.}
240 skuang 3772 \label{method}
241     \end{figure}
242    
243 gezelter 3786 By implementing these operations at a fixed frequency, stable thermal
244     and momentum fluxes can both be applied and the corresponding
245     temperature and momentum gradients can be established.
246 skuang 3770
247 gezelter 3788 Compared to the previous NIVS method, the VSS-RNEMD approach is
248 gezelter 3786 computationally simpler in that only quadratic equations are involved
249     to determine a set of scaling coefficients, while the NIVS method
250     required solution of quartic equations. Furthermore, this method
251     implements {\it isotropic} scaling of velocities in respective slabs,
252     unlike NIVS, which required extra flexibility in the choice of scaling
253     coefficients to allow for the energy and momentum constraints. Most
254     importantly, separating the linear momentum flux from velocity scaling
255     avoids the underlying cause of the thermal anisotropy in NIVS. In
256     later sections, we demonstrate that this can improve the calculated
257     shear viscosities from RNEMD simulations.
258 gezelter 3769
259 gezelter 3788 The VSS-RNEMD approach has advantages over the original momentum
260 gezelter 3786 swapping in many respects. In the swapping method, the velocity
261     vectors involved are usually very different (or the generated flux
262     would be quite small), thus the swapping tends to create strong
263     perturbations in the neighborhood of the particles involved. In
264 gezelter 3788 comparison, the VSS approach distributes the flux widely to all
265 gezelter 3786 particles in a slab so that perturbations in the flux generating
266     region are minimized. Additionally, momentum swapping steps tend to
267     result in nonthermal velocity distributions when the imposed flux is
268     large and diffusion from the neighboring slabs cannot carry momentum
269     away quickly enough. In comparison, the scaling and shearing moves
270 gezelter 3788 made in the VSS approach preserve the shapes of the equilibrium
271 gezelter 3786 velocity disributions (e.g. Maxwell-Boltzmann). The results presented
272     in later sections will illustrate that this is quite helpful in
273     retaining reasonable thermal distributions in a simulation.
274 gezelter 3769
275 skuang 3772 \section{Computational Details}
276 skuang 3773 The algorithm has been implemented in our MD simulation code,
277 gezelter 3786 OpenMD.\cite{Meineke:2005gd,openmd} We will first compare the method
278     with previous RNEMD methods and equilibrium MD in homogeneous
279     fluids (Lennard-Jones and SPC/E water). We have also used the new
280     method to simulate the interfacial friction of different heterogeneous
281     interfaces (Au (111) with organic solvents, Au(111) with SPC/E water,
282 gezelter 3788 and the ice Ih - liquid water interface).
283 gezelter 3769
284 skuang 3773 \subsection{Simulation Protocols}
285 gezelter 3786 The systems we investigated were set up in orthorhombic simulation
286 skuang 3784 cells with periodic boundary conditions in all three dimensions. The
287 gezelter 3786 $z$-axes of these cells were typically quite long and served as the
288     temperature and/or momentum gradient axes. The cells were evenly
289     divided into $N$ slabs along this axis, with $N$ varying for the
290     individual system. The $x$ and $y$ axes were of similar lengths in all
291     simulations. In all cases, before introducing a nonequilibrium method
292     to establish steady thermal and/or momentum gradients, equilibration
293     simulations were run under the canonical ensemble with a Nos\'e-Hoover
294     thermostat\cite{hoover85} followed by further equilibration using
295     standard constant energy (NVE) conditions. For SPC/E water
296     simulations, isobaric-isothermal equilibrations\cite{melchionna93}
297     were performed before equilibration to reach standard densities at
298     atmospheric pressure (1 bar); for interfacial systems, similar
299     equilibrations with anisotropic box relaxations are used to relax the
300     surface tension in the $xy$ plane.
301 skuang 3772
302 gezelter 3786 While homogeneous fluid systems can be set up with random
303     configurations, interfacial systems are typically prepared with a
304     single crystal face presented perpendicular to the $z$-axis. This
305     crystal face is aligned in the x-y plane of the periodic cell, and the
306     solvent occupies the region directly above and below a crystalline
307     slab. The preparation and equilibration of butanethiol covered
308     Au(111) surfaces, as well as the solvation and equilibration processes
309     used for these interfaces are described in detail in reference
310     \cite{kuang:AuThl}.
311 skuang 3773
312 gezelter 3786 For the ice / liquid water interfaces, the basal surface of ice
313     lattice was first constructed. Hirsch and Ojam\"{a}e
314     \cite{doi:10.1021/jp048434u} explored the energetics of ice lattices
315     with all possible proton ordered configurations. We utilized Hirsch
316     and Ojam\"{a}e's structure 6 ($P2_12_12_1$) which is an orthorhombic
317 gezelter 3788 cell that produces a proton-ordered version of ice Ih. The basal face
318     of ice in this unit cell geometry is the $\{0~0~1\}$ face. Although
319 skuang 3784 experimental solid/liquid coexistant temperature under normal pressure
320 gezelter 3786 are close to 273K, Bryk and Haymet's simulations of ice/liquid water
321     interfaces with different models suggest that for SPC/E, the most
322     stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
323     Therefore, our ice/liquid water simulations were carried out at an
324     average temperature of 225K. Molecular translation and orientational
325     restraints were applied in the early stages of equilibration to
326     prevent melting of the ice slab. These restraints were removed during
327     NVT equilibration, well before data collection was carried out.
328 gezelter 3769
329     \subsection{Force Field Parameters}
330 gezelter 3788 For comparing the VSS-RNEMD method with previous work, we retained
331 skuang 3784 force field parameters consistent with previous simulations. Argon is
332     the Lennard-Jones fluid used here, and its results are reported in
333 gezelter 3786 reduced units for purposes of direct comparison with previous
334     calculations.
335 gezelter 3769
336 gezelter 3786 For our water simulations, we utilized the SPC/E model throughout this
337     work. Previous work for transport properties of SPC/E water model is
338     available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct
339     comparison with previous calculation methods is possible.
340 gezelter 3769
341 skuang 3774 The Au-Au interaction parameters in all simulations are described by
342     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
343     QSC potentials include zero-point quantum corrections and are
344 gezelter 3769 reparametrized for accurate surface energies compared to the
345 skuang 3775 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
346     Spohr potential was adopted\cite{ISI:000167766600035} to depict
347     Au-H$_2$O interactions.
348 gezelter 3769
349 skuang 3784 For our gold/organic liquid interfaces, the small organic molecules
350     included in our simulations are the Au surface capping agent
351 gezelter 3786 butanethiol as well as liquid hexane and liquid toluene. The
352     United-Atom
353 skuang 3774 models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
354 gezelter 3786 for these components were used in this work for computational
355 skuang 3774 efficiency, while maintaining good accuracy. We refer readers to our
356     previous work\cite{kuang:AuThl} for further details of these models,
357     as well as the interactions between Au and the above organic molecule
358     components.
359 gezelter 3769
360 skuang 3774 \subsection{Thermal conductivities}
361 gezelter 3786 When the linear momentum flux $\vec{j}_z(\vec{p})$ is set to zero and
362 gezelter 3788 the target $J_z$ is non-zero, VSS-RNEMD imposes kinetic energy
363     transfer between the slabs, which can be used for computation of
364     thermal conductivities. As with other RNEMD methods, we assume that we
365 gezelter 3786 are in the linear response regime of the temperature gradient with
366     respect to the thermal flux. The thermal conductivity ($\lambda$) can
367 gezelter 3788 then be calculated using the imposed kinetic energy flux and the
368     measured thermal gradient:
369 skuang 3775 \begin{equation}
370     J_z = -\lambda \frac{\partial T}{\partial z}
371     \end{equation}
372     Like other imposed-flux methods, the energy flux was calculated using
373     the total non-physical energy transferred (${E_{total}}$) from slab
374 gezelter 3788 {\it C} to slab {\it H}, which was recorded throughout the simulation, and
375 gezelter 3786 the total data collection time $t$:
376 skuang 3775 \begin{equation}
377 gezelter 3786 J_z = \frac{E_{total}}{2 t L_x L_y}.
378 skuang 3775 \end{equation}
379 gezelter 3788 Here, $L_x$ and $L_y$ denote the dimensions of the plane in the
380     simulation cell that is perpendicular to the thermal gradient, and a
381     factor of two in the denominator is necessary as the heat transport
382     occurs in both the $+z$ and $-z$ directions. The average temperature
383     gradient ${\langle\partial T/\partial z\rangle}$ can be obtained by a
384     linear regression of the temperature profile, which is recorded during
385     a simulation for each slab in a cell. For Lennard-Jones simulations,
386 skuang 3775 thermal conductivities are reported in reduced units
387     (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
388    
389 skuang 3774 \subsection{Shear viscosities}
390 gezelter 3786 Alternatively, when the linear momentum flux $\vec{j}_z(\vec{p})$ is
391     non-zero in either the $x$ or $y$ directions, the method can be used
392     to compute the shear viscosity. For isotropic systems, the direction
393     of $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
394     ability to arbitarily specify the vector direction in our method could
395 gezelter 3788 be convenient when working with anisotropic interfaces.
396 skuang 3775
397 gezelter 3786 In a manner similar to the thermal conductivity calculations, a linear
398     response of the momentum gradient with respect to the shear stress is
399     assumed, and the shear viscosity ($\eta$) can be obtained with the
400     imposed momentum flux (e.g. in $x$ direction) and the measured
401     velocity gradient:
402 skuang 3775 \begin{equation}
403     j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
404     \end{equation}
405     where the flux is similarly defined:
406     \begin{equation}
407     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
408     \end{equation}
409     with $P_x$ being the total non-physical momentum transferred within
410 skuang 3784 the data collection time. Also, the averaged velocity gradient
411 skuang 3775 ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
412 skuang 3784 regression of the $x$ component of the mean velocity ($\langle
413     v_x\rangle$) in each of the bins. For Lennard-Jones simulations, shear
414     viscosities are also reported in reduced units
415 skuang 3775 (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
416    
417 gezelter 3786 Although $J_z$ may be switched off for shear viscosity simulations,
418 gezelter 3788 the VSS-RNEMD method allows the user the ability to {\it
419     simultaneously} impose both a thermal and a momentum flux during a
420     single simulation. This creates configurations with coincident
421     temperature and a velocity gradients. Since the viscosity is
422     generally a function of temperature, the local viscosity depends on
423     the local temperature in the fluid. Therefore, in a single simulation,
424     viscosity at $z$ (corresponding to a certain $T$) can be computed with
425     the applied shear flux and the local velocity gradient (which can be
426     obtained by finite difference approximation). This means that the
427     temperature dependence of the viscosity can be mapped out in only one
428 gezelter 3786 trajectory. Results for shear viscosity computations of SPC/E water
429 gezelter 3788 will demonstrate VSS-RNEMD's efficiency in this respect.
430 skuang 3784
431 gezelter 3786 \subsection{Interfacial friction and slip length}
432     Shear stress creates a velocity gradient within bulk fluid phases, but
433     at solid-liquid interfaces, the effects of a shear stress depend on
434     the molecular details of the interface. The interfacial friction
435     coefficient, $\kappa$, relates the shear stress (e.g. along the
436 gezelter 3788 $x$-axis) with the relative velocity of the fluid tangent to the
437     interface:
438 skuang 3775 \begin{equation}
439 gezelter 3786 j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right]
440 skuang 3775 \end{equation}
441 gezelter 3786 where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured
442     directly adjacent to the interface. Under ``stick'' boundary
443     condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads
444     to $\kappa\rightarrow\infty$. However, for ``slip'' boundary
445     conditions at a solid-liquid interface, $\kappa$ becomes finite. To
446     characterize the interfacial boundary conditions, the slip length,
447     $\delta$, is defined by the ratio of the fluid-phase viscosity to the
448     friction coefficient of the interface:
449 skuang 3775 \begin{equation}
450     \delta = \frac{\eta}{\kappa}
451     \end{equation}
452 gezelter 3788 In ``stick'' boundary conditions, $\delta\rightarrow 0$, so $\delta$
453     is a measure of how ``slippery'' an interface is. Figure
454     \ref{slipLength} illustrates how this quantity is defined and computed
455     for a solid-liquid interface.
456 gezelter 3769
457 skuang 3775 \begin{figure}
458     \includegraphics[width=\linewidth]{defDelta}
459 gezelter 3788 \caption{The slip length ($\delta$) can be obtained from a velocity
460     profile along an axis perpendicular to the interface. The data shown
461     is for an Au / hexane interface -- the crystalline region (Au) is
462     moving as a block (lower dots), while the measured velocity gradient
463     in the hexane phase is discontinuous at the interface.}
464 skuang 3775 \label{slipLength}
465     \end{figure}
466 gezelter 3769
467 gezelter 3788 Since the VSS-RNEMD method can be applied for interfaces as well as
468     for bulk materials, the shear stress is applied in an identical manner
469     to the shear viscosity computations, e.g. by applying an unphysical
470     momentum flux, $j_z(\vec{p})$. With the correct choice of $\vec{p}$
471     in the $x-y$ plane, one can compute friction coefficients and slip
472     lengths for a number of different dragging vectors on a given slab.
473     The corresponding velocity profiles can be obtained as shown in Figure
474 gezelter 3786 \ref{slipLength}, in which the velocity gradients within the liquid
475     phase and the velocity difference at the liquid-solid interface can be
476     easily measured from saved simulation data.
477 skuang 3775
478 skuang 3776 \section{Results and Discussions}
479     \subsection{Lennard-Jones fluid}
480 gezelter 3786 Our orthorhombic simulation cell for the Lennard-Jones fluid has
481 gezelter 3788 identical parameters to previous work to facilitate
482     comparison.\cite{kuang:164101} Thermal conductivities and shear
483 gezelter 3789 viscosities were computed with the new VSS algorithm, and the results
484     were compared with the previous NIVS algorithm. However, since the
485     NIVS algorithm produces temperature anisotropy under shear stress,
486     these results are also compared to the previous momentum swapping
487 gezelter 3788 approaches. Table \ref{LJ} lists these values with various fluxes in
488     reduced units.
489 skuang 3770
490 skuang 3776 \begin{table*}
491     \begin{minipage}{\linewidth}
492     \begin{center}
493 gezelter 3769
494 skuang 3776 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
495     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
496     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
497 gezelter 3789 at various momentum fluxes. The new VSS method yields similar
498 skuang 3776 results to previous RNEMD methods. All results are reported in
499     reduced unit. Uncertainties are indicated in parentheses.}
500    
501 gezelter 3789 \begin{tabular}{cc|ccc|cc}
502 skuang 3776 \hline\hline
503     \multicolumn{2}{c}{Momentum Exchange} &
504 gezelter 3789 \multicolumn{3}{c}{$\lambda^*$} &
505 skuang 3776 \multicolumn{2}{c}{$\eta^*$} \\
506     \hline
507 gezelter 3789 Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ & Swapping &
508 skuang 3785 NIVS\cite{kuang:164101} & This work & Swapping & This work \\
509 skuang 3776 \hline
510 gezelter 3789 0.116 & 0.16 & 7.03(0.34) & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
511     0.232 & 0.09 & 7.03(0.14) & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
512     0.463 & 0.047 & 6.91(0.42) & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
513     0.926 & 0.024 & 7.52(0.15) & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
514     1.158 & 0.019 & 7.41(0.29) & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
515 skuang 3776 \hline\hline
516     \end{tabular}
517     \label{LJ}
518     \end{center}
519     \end{minipage}
520     \end{table*}
521 gezelter 3769
522 skuang 3776 \subsubsection{Thermal conductivity}
523 gezelter 3789 Our thermal conductivity calculations yield comparable results to the
524     previous NIVS algorithm. This indicates that the thermal gradients
525     produced using this method are also quite close to previous RNEMD
526     methods. Simulations with moderately large thermal fluxes tend to
527     yield more reliable thermal gradients than under NIVS and thus avoid
528     large errors. Extreme values for the applied thermal flux can
529     introduce side effects such as non-linear temperature gradients and
530     inadvertent phase transitions, so these are avoided.
531 gezelter 3769
532 gezelter 3789 Since the scaling operation is isotropic in VSS, the temperature
533     anisotropy that was observed under the earlier NIVS approach should be
534     absent. Furthermore, the VSS method avoids unintended momentum flux
535     when only thermal flux is being imposed. This was not always possible
536     with swapping or NIVS approaches. The thermal energy exchange in
537     swapping ($\vec{p}_i$ in slab {\it C} swapped with $\vec{p}_j$ in slab
538     {\it H}) or NIVS (total slab momentum components $P^\alpha$ scaled to
539     $\alpha P^\alpha$) do not achieve this unless thermal flux vanishes
540     (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$). In this sense, the VSS
541     method achieves minimal perturbation to a simulation while imposing
542     thermal flux.
543 skuang 3776
544     \subsubsection{Shear viscosity}
545 skuang 3785 Table \ref{LJ} also compares our shear viscosity results with the
546 gezelter 3789 momentum swapping approach. Our calculations show that the VSS method
547     predicted similar values for shear viscosities to the momentum
548     swapping approach, as well as providing similar velocity gradient
549     profiles. Moderately large momentum fluxes are helpful in reducing the
550     errors in measured velocity gradients and thus the shear viscosity
551     values. However, it should be noted that the momentum swapping
552     approach tends to produce non-thermal velocity distributions in the
553     two slabs.\cite{Maginn:2010}
554 skuang 3776
555 gezelter 3789 To show that the temperature is isotropic within each slab under VSS,
556     we measured the three one-dimensional temperatures in each of the
557     slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
558 skuang 3785 temperatures were calculated after subtracting the contribution from
559 gezelter 3789 bulk (shearing) velocities of the slabs. The one-dimensional
560     temperature profiles show no observable differences between the three
561     box dimensions. This ensures that VSS automatically preserves
562     temperature isotropy during shear viscosity calculations.
563 skuang 3776
564 gezelter 3769 \begin{figure}
565 skuang 3776 \includegraphics[width=\linewidth]{tempXyz}
566 gezelter 3788 \caption{Unlike the previous NIVS algorithm, the VSS method does not
567     produce thermal anisotropy under shearing stress. Temperature
568     differences along the different box axes were significantly smaller
569     than the error bars. Note that the two regions of elevated
570     temperature are caused by the shear stress. This is the same
571     frictional heating reported previously by Tenney and
572     Maginn.\cite{Maginn:2010}}
573 skuang 3776 \label{tempXyz}
574 gezelter 3769 \end{figure}
575    
576 gezelter 3789 The velocity distribution profiles were further tested by imposing a
577     large shear stress. Figure \ref{vDist} demonstrates how the VSS method
578     is able to maintain nearly ideal Maxwell-Boltzmann velocity
579     distributions even under large imposed fluxes. Previous swapping
580     methods tend to deplete particles with positive velocities in the
581     negative velocity slab ({\it C}) and deplete particles with negative
582     velocities in the {\it H} slab, leaving significant `notches' in the
583     velocity distributions. The problematic velocity distributions become
584     significant when the imposed-flux becomes so large that diffusion from
585     neighboring slabs cannot offset the depletion. Simultaneously,
586     abnormal peaks appear in the other regions of the velocity
587     distributions as high (or low) momentum particles are introduced in to
588     the corresponding slab. These nonthermal distributions limit
589     application of the swapping approach in shear stress simulations. The
590     VSS method avoids the above problematic distributions by altering the
591     means of applying momentum flux. Comparatively, velocity distributions
592     recorded from simulations with the VSS method is so close to the ideal
593     Maxwell-Boltzmann distributions that no obvious difference is visible
594     in Figure \ref{vDist}.
595 gezelter 3769
596 skuang 3776 \begin{figure}
597 gezelter 3788 \centering
598     \includegraphics[width=5.5in]{velDist}
599     \caption{Velocity distributions that develop under VSS more closely
600     resemble ideal Maxwell-Boltzmann distributions than earlier
601     momentum-swapping RNEMD approaches. This data was obtained from
602 skuang 3779 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
603 skuang 3777 swapping interval of 20 time steps). This is a relatively large flux
604 gezelter 3788 which demonstrates some of the non-thermal velocity distributions
605     that can develop under swapping RNEMD.}
606 skuang 3776 \label{vDist}
607     \end{figure}
608 gezelter 3769
609 skuang 3776 \subsection{Bulk SPC/E water}
610 gezelter 3789 We also computed the thermal conductivity and shear viscosity of SPC/E
611     water. A simulation cell with 1000 molecules was set up in a similar
612     manner to Ref. \cite{kuang:164101}. For thermal conductivity
613     calculations, measurements were taken to compare with previous RNEMD
614     methods; for shear viscosity computations, simulations were run under
615     a series of temperatures (with corresponding pressure relaxation using
616     the isobaric-isothermal ensemble\cite{melchionna93}), and results were
617     compared to available data from equilibrium molecular dynamics
618     calculations.\cite{10.1063/1.3330544,Medina2011} Additionally, a
619     simulation with {\it both} applied thermal and momentum gradients was
620     carried out to map out shear viscosity as a function of temperature.
621 skuang 3777
622 skuang 3776 \subsubsection{Thermal conductivity}
623 gezelter 3789 Table \ref{spceThermal} summarizes the thermal conductivity of SPC/E
624     under different temperatures and thermal gradients compared with the
625     previous NIVS results\cite{kuang:164101} and experimental
626     measurements.\cite{WagnerKruse} Note that no appreciable drift of
627     total system energy or temperature was observed when the VSS method
628     was applied, which indicates that our algorithm conserves total energy
629     well for systems involving electrostatic interactions.
630 gezelter 3769
631 gezelter 3789 Measurements using the VSS method established similar temperature
632 skuang 3778 gradients to the previous NIVS method. Our simulation results are in
633 gezelter 3789 fairly good agreement with those from previous simulations. Both
634     methods yield values in reasonable agreement with experimental
635     values. Simulations using larger thermal gradients and those with
636     longer gradient axes ($z$) have less measurable noise.
637 skuang 3778
638     \begin{table*}
639     \begin{minipage}{\linewidth}
640     \begin{center}
641    
642     \caption{Thermal conductivity of SPC/E water under various
643     imposed thermal gradients. Uncertainties are indicated in
644     parentheses.}
645    
646 gezelter 3789 \begin{tabular}{cc|ccc}
647 skuang 3778 \hline\hline
648     $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
649     {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
650 gezelter 3789 (K) & (K/\AA) & This work & NIVS\cite{kuang:164101} &
651 skuang 3778 Experiment\cite{WagnerKruse} \\
652     \hline
653 gezelter 3789 300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\ \hline
654 skuang 3778 318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
655 gezelter 3789 & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
656     & 0.8 & 0.786(0.009)\footnote{Simulation with 2000 SPC/E molecules} & & \\
657 skuang 3778 \hline\hline
658     \end{tabular}
659     \label{spceThermal}
660     \end{center}
661     \end{minipage}
662     \end{table*}
663    
664 skuang 3776 \subsubsection{Shear viscosity}
665 gezelter 3789 Shear viscosity was computed for SPC/E water at a range of
666     temperatures that span the liquidus range for water under atmospheric
667     pressure. VSS-RNEMD simulations predict a similar trend of $\eta$
668     vs. $T$ to equilibrium molecular dynamics (EMD) results and are
669     presented in table \ref{spceShear}. Our results show no significant
670     differences from the earlier EMD calculations. Since each value
671     reported using our method takes a single trajectory of simulation,
672     instead of average from many trajectories when using EMD, the VSS
673     method provides an efficient means for computing the shear viscosity.
674 gezelter 3769
675 skuang 3778 \begin{table*}
676     \begin{minipage}{\linewidth}
677     \begin{center}
678    
679     \caption{Computed shear viscosity of SPC/E water under different
680 gezelter 3789 temperatures. Results are compared to those obtained with
681     equilibrium molecular dynamics.\cite{Medina2011} Uncertainties
682     are indicated in parentheses.}
683 skuang 3778
684 gezelter 3789 \begin{tabular}{cc|cc}
685 skuang 3778 \hline\hline
686     $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
687     {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
688 skuang 3785 (K) & (10$^{10}$s$^{-1}$) & This work & Previous
689     simulations\cite{Medina2011} \\
690 skuang 3778 \hline
691 skuang 3785 273 & 1.12 & 1.218(0.004) & 1.282(0.048) \\
692     & 1.79 & 1.140(0.012) & \\
693 skuang 3790 \hline
694 skuang 3785 303 & 2.09 & 0.646(0.008) & 0.643(0.019) \\
695 skuang 3790 \hline
696 skuang 3785 318 & 2.50 & 0.536(0.007) & \\
697     & 5.25 & 0.510(0.007) & \\
698 skuang 3790 & 2.82 & 0.474(0.003)\footnote{Simulation with 2000 SPC/E
699     molecules.} & \\
700     \hline
701 skuang 3785 333 & 3.10 & 0.428(0.002) & 0.421(0.008) \\
702 skuang 3790 \hline
703 skuang 3785 363 & 2.34 & 0.279(0.014) & 0.291(0.005) \\
704     & 4.26 & 0.306(0.001) & \\
705 skuang 3778 \hline\hline
706     \end{tabular}
707     \label{spceShear}
708     \end{center}
709     \end{minipage}
710     \end{table*}
711 gezelter 3769
712 skuang 3785 A more effective way to map out $\eta$ vs $T$ is to combine a momentum
713     flux with a thermal flux. Figure \ref{Tvxdvdz} shows the thermal and
714     velocity gradient in one such simulation. At different positions with
715     different temperatures, the velocity gradient is not a constant but
716     can be computed locally. With the data provided in Figure
717     \ref{Tvxdvdz}, a series of $\eta$ is calculated as in Figure
718     \ref{etaT} and a linear fit was performed to $\partial v_x/\partial z$
719     vs. $z$ so that the resulted $\eta$ can be present as a curve as
720     well. For comparison, other results are also mapped in the figure.
721    
722     \begin{figure}
723 gezelter 3788 \centering
724     \includegraphics[width=5.5in]{tvxdvdz}
725 skuang 3785 \caption{With a combination of a thermal and a momentum flux, a
726 gezelter 3788 simulation can exhibit both a temperature (top) and a velocity
727     (middle) gradient. $\partial v_x/\partial z$ depends on the local
728     temperature, so it varies along the gradient axis. This derivative
729     can be computed using finite difference approximations (lower) and
730     can be used to calculate $\eta$ vs $T$ (Figure \ref{etaT}).}
731 skuang 3785 \label{Tvxdvdz}
732     \end{figure}
733    
734     From Figure \ref{etaT}, one can see that the generated curve agrees
735     well with the above RNEMD simulations at different temperatures, as
736     well as results reported using EMD
737     methods\cite{10.1063/1.3330544,Medina2011} in much of the temperature
738     range simulated. However, this curve has relatively large error in
739     lower temperature regions and has some difference in predicting $\eta$
740     near 273K. Provided that this curve only takes one trajectory to
741     generate, these results are of satisfactory efficiency and
742     accuracy. Since previous work already pointed out that the SPC/E model
743     tends to predict lower viscosity compared to experimental
744     data,\cite{Medina2011} experimental comparison are not given here.
745    
746     \begin{figure}
747     \includegraphics[width=\linewidth]{etaT}
748 gezelter 3788 \caption{The temperature dependence of the shear viscosity generated
749     by a {\it single} VSS-RNEMD simulation. Applying both thermal and
750     momentum gradient predicts reasonable values in much of the
751     temperature range being tested.}
752 skuang 3785 \label{etaT}
753     \end{figure}
754    
755 skuang 3779 \subsection{Interfacial frictions and slip lengths}
756 skuang 3785 Another attractive aspect of our method is the ability to apply
757     momentum and/or thermal flux in nonhomogeneous systems, where
758     molecules of different identities (or phases) are segregated in
759     different regions. We have previously studied the interfacial thermal
760     transport of a series of metal gold-liquid
761     surfaces\cite{kuang:164101,kuang:AuThl}, and would like to further
762     investigate the relationship between this phenomenon and the
763 skuang 3779 interfacial frictions.
764 gezelter 3769
765 skuang 3779 Table \ref{etaKappaDelta} includes these computations and previous
766     calculations of corresponding interfacial thermal conductance. For
767     bare Au(111) surfaces, slip boundary conditions were observed for both
768     organic and aqueous liquid phases, corresponding to previously
769 skuang 3785 computed low interfacial thermal conductance. In comparison, the
770     butanethiol covered Au(111) surface appeared to be sticky to the
771     organic liquid layers in our simulations. We have reported conductance
772     enhancement effect for this surface capping agent,\cite{kuang:AuThl}
773     and these observations have a qualitative agreement with the thermal
774     conductance results. This agreement also supports discussions on the
775     relationship between surface wetting and slip effect and thermal
776     conductance of the
777     interface.\cite{PhysRevLett.82.4671,doi:10.1080/0026897031000068578,garde:PhysRevLett2009}
778 skuang 3776
779 gezelter 3769 \begin{table*}
780     \begin{minipage}{\linewidth}
781     \begin{center}
782    
783 skuang 3779 \caption{Computed interfacial friction coefficient values for
784     interfaces with various components for liquid and solid
785     phase. Error estimates are indicated in parentheses.}
786 gezelter 3769
787 skuang 3779 \begin{tabular}{llcccccc}
788 gezelter 3769 \hline\hline
789 skuang 3779 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
790     & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
791     \cite{kuang:164101}.} \\
792 skuang 3785 surface & molecules & K & MPa & mPa$\cdot$s &
793     10$^4$Pa$\cdot$s/m & nm & MW/m$^2$/K \\
794 gezelter 3769 \hline
795 skuang 3785 Au(111) & hexane & 200 & 1.08 & 0.197(0.009) & 5.30(0.36) &
796 skuang 3790 3.72(0.25) & 46.5 \\
797 skuang 3785 & & & 2.15 & 0.141(0.002) & 5.31(0.26) &
798 skuang 3790 2.76(0.14) & \\
799 skuang 3785 Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.286(0.019) & $\infty$
800     & 0 & 131 \\
801     & & & 5.39 & 0.320(0.006) & $\infty$
802     & 0 & \\
803 gezelter 3769 \hline
804 skuang 3785 Au(111) & toluene & 200 & 1.08 & 0.722(0.035) & 15.7(0.7) &
805 skuang 3790 4.60(0.22) & 70.1 \\
806 skuang 3785 & & & 2.16 & 0.544(0.030) & 11.2(0.5) &
807 skuang 3790 4.86(0.27) & \\
808 skuang 3785 Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.980(0.057) &
809     $\infty$ & 0 & 187 \\
810     & & & 10.8 & 0.995(0.005) &
811     $\infty$ & 0 & \\
812 skuang 3779 \hline
813 skuang 3785 Au(111) & water & 300 & 1.08 & 0.399(0.050) & 1.928(0.022) &
814 skuang 3790 20.7(2.6) & 1.65 \\
815 skuang 3785 & & & 2.16 & 0.794(0.255) & 1.895(0.003) &
816 skuang 3790 41.9(13.5) & \\
817 skuang 3779 \hline
818 skuang 3785 ice(basal) & water & 225 & 19.4 & 15.8(0.2) & $\infty$ & 0 & \\
819 gezelter 3769 \hline\hline
820     \end{tabular}
821 skuang 3779 \label{etaKappaDelta}
822 gezelter 3769 \end{center}
823     \end{minipage}
824     \end{table*}
825    
826 skuang 3779 An interesting effect alongside the surface friction change is
827     observed on the shear viscosity of liquids in the regions close to the
828 skuang 3785 solid surface. In our results, $\eta$ measured near a ``slip'' surface
829     tends to be smaller than that near a ``stick'' surface. This may
830     suggest the influence from an interface on the dynamic properties of
831     liquid within its neighbor regions. It is known that diffusions of
832     solid particles in liquid phase is affected by their surface
833     conditions (stick or slip boundary).\cite{10.1063/1.1610442} Our
834     observations could provide a support to this phenomenon.
835 gezelter 3769
836 skuang 3779 In addition to these previously studied interfaces, we attempt to
837     construct ice-water interfaces and the basal plane of ice lattice was
838 skuang 3785 studied here. In contrast to the Au(111)/water interface, where the
839     friction coefficient is substantially small and large slip effect
840 skuang 3779 presents, the ice/liquid water interface demonstrates strong
841 skuang 3785 solid-liquid interactions and appears to be sticky. The supercooled
842     liquid phase is an order of magnitude more viscous than measurements
843     in previous section. It would be of interst to investigate the effect
844     of different ice lattice planes (such as prism and other surfaces) on
845     interfacial friction and the corresponding liquid viscosity.
846 gezelter 3769
847 skuang 3776 \section{Conclusions}
848 skuang 3779 Our simulations demonstrate the validity of our method in RNEMD
849     computations of thermal conductivity and shear viscosity in atomic and
850     molecular liquids. Our method maintains thermal velocity distributions
851     and avoids thermal anisotropy in previous NIVS shear stress
852     simulations, as well as retains attractive features of previous RNEMD
853     methods. There is no {\it a priori} restrictions to the method to be
854     applied in various ensembles, so prospective applications to
855     extended-system methods are possible.
856 gezelter 3769
857 skuang 3785 Our method is capable of effectively imposing thermal and/or momentum
858     flux accross an interface. This facilitates studies that relates
859     dynamic property measurements to the chemical details of an
860     interface. Therefore, investigations can be carried out to
861     characterize interfacial interactions using the method.
862 gezelter 3769
863 skuang 3779 Another attractive feature of our method is the ability of
864 skuang 3785 simultaneously introducing thermal and momentum gradients in a
865     system. This facilitates us to effectively map out the shear viscosity
866     with respect to a range of temperature in single trajectory of
867     simulation with satisafactory accuracy. Complex systems that involve
868     thermal and momentum gradients might potentially benefit from
869     this. For example, the Soret effects under a velocity gradient might
870     be models of interest to purification and separation researches.
871 gezelter 3769
872     \section{Acknowledgments}
873     Support for this project was provided by the National Science
874     Foundation under grant CHE-0848243. Computational time was provided by
875     the Center for Research Computing (CRC) at the University of Notre
876     Dame.
877    
878     \newpage
879    
880 skuang 3770 \bibliography{stokes}
881 gezelter 3769
882     \end{doublespace}
883     \end{document}