ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3791
Committed: Wed Mar 21 17:34:37 2012 UTC (12 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 42647 byte(s)
Log Message:
submitted version

File Contents

# User Rev Content
1 gezelter 3791 \documentclass{tMPH2e}
2 gezelter 3769 \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5 gezelter 3791 %\usepackage{endfloat}
6 gezelter 3769 \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 gezelter 3791 %\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 gezelter 3769
20     % double space list of tables and figures
21 gezelter 3791 %\AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22     %\setlength{\abovecaptionskip}{20 pt}
23     %\setlength{\belowcaptionskip}{30 pt}
24 gezelter 3769
25     %\renewcommand\citemid{\ } % no comma in optional reference note
26     \bibpunct{[}{]}{,}{n}{}{;}
27 gezelter 3791 \bibliographystyle{tMPH}
28     \markboth{Kuang \& Gezelter}{Molecular Physics}
29 gezelter 3769 \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 gezelter 3791 \author{Shenyu Kuang and J. Daniel Gezelter$^{\ast}$\thanks{
35     $^{\ast}$ Corresponding author. Email: gezelter@nd.edu} \\ \vspace{6pt}
36     {\em Department of Chemistry and Biochemistry,\\
37 gezelter 3769 University of Notre Dame\\
38 gezelter 3791 Notre Dame, Indiana 46556} \\ \vspace{6pt}
39     \received{15 December 2011}
40     }
41 gezelter 3769
42     \date{\today}
43    
44     \maketitle
45    
46 gezelter 3791 %\begin{doublespace}
47 gezelter 3769
48     \begin{abstract}
49 skuang 3779 We present a new method for introducing stable nonequilibrium
50     velocity and temperature gradients in molecular dynamics simulations
51 gezelter 3788 of heterogeneous and interfacial systems. This method conserves both
52     the linear momentum and total energy of the system and improves
53     previous reverse non-equilibrium molecular dynamics (RNEMD) methods
54     while retaining equilibrium thermal velocity distributions in each
55     region of the system. The new method avoids the thermal anisotropy
56 gezelter 3786 produced by previous methods by using isotropic velocity scaling and
57 gezelter 3788 shearing (VSS) on all of the molecules in specific regions. To test
58     the method, we have computed the thermal conductivity and shear
59 skuang 3785 viscosity of model liquid systems as well as the interfacial
60 gezelter 3786 friction coeefficients of a series of solid / liquid interfaces. The
61     method's ability to combine the thermal and momentum gradients
62     allows us to obtain shear viscosity data for a range of temperatures
63     from a single trajectory.
64 gezelter 3769 \end{abstract}
65    
66     \newpage
67    
68     %\narrowtext
69    
70     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71     % BODY OF TEXT
72     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73    
74     \section{Introduction}
75 gezelter 3786 One of the standard ways to compute transport coefficients such as the
76     viscosities and thermal conductivities of liquids is to use
77     imposed-flux non-equilibrium molecular dynamics
78     methods.\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
79     These methods establish stable non-equilibrium conditions in a
80     simulation box using an applied momentum or thermal flux between
81     different regions of the box. The corresponding temperature or
82     velocity gradients which develop in response to the applied flux is
83     related (via linear response theory) to the transport coefficient of
84     interest. These methods are quite efficient, in that they do not need
85     many trajectories to provide information about transport properties. To
86     date, they have been utilized in computing thermal and mechanical
87     transfer of both homogeneous liquids as well as heterogeneous systems
88     such as solid-liquid
89 skuang 3785 interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
90 skuang 3770
91 gezelter 3786 The reverse non-equilibrium molecular dynamics (RNEMD) methods utilize
92     additional constraints that ensure conservation of linear momentum and
93     total energy of the system while imposing the desired flux. The RNEMD
94     methods are therefore capable of sampling various thermodynamically
95     relevent ensembles, including the micro-canonical (NVE) ensemble,
96     without resorting to an external thermostat. The original approaches
97     proposed by M\"{u}ller-Plathe {\it et
98 skuang 3771 al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
99 gezelter 3786 momentum swapping moves for generating energy/momentum fluxes. The
100     swapping moves can also be made compatible with particles of different
101     identities. Although the swapping moves are simple to implement in
102     molecular simulations, Tenney and Maginn have shown that they create
103     nonthermal velocity distributions.\cite{Maginn:2010} Furthermore, this
104     approach is not particularly efficient for kinetic energy transfer
105     between particles of different identities, particularly when the mass
106     difference between the particles becomes significant. This problem
107     makes applying swapping-move RNEMD methods on heterogeneous
108     interfacial systems somewhat difficult.
109 skuang 3770
110 gezelter 3786 Recently, we developed a somewhat different approach to applying
111 gezelter 3788 thermal fluxes in RNEMD simulation using a non-isotropic velocity
112     scaling (NIVS) algorithm.\cite{kuang:164101} This algorithm scales all
113 gezelter 3786 atomic velocity vectors in two separate regions of a simulated system
114     using two diagonal scaling matrices. The scaling matrices are
115     determined by solving single quartic equation which includes linear
116     momentum and kinetic energy conservation constraints as well as the
117     target thermal flux between the regions. The NIVS method is able to
118     effectively impose a wide range of kinetic energy fluxes without
119     significant perturbation to the velocity distributions away from ideal
120     Maxwell-Boltzmann distributions, even in the presence of heterogeneous
121     interfaces. We successfully applied this approach in studying the
122     interfacial thermal conductance at chemically-capped metal-solvent
123 skuang 3771 interfaces.\cite{kuang:AuThl}
124 gezelter 3769
125 gezelter 3786 The NIVS approach works very well for preparing stable {\it thermal}
126     gradients. However, as we pointed out in the original paper, it has
127     limited application in imposing {\it linear} momentum fluxes (which
128     are required for measuring shear viscosities). The reason for this is
129     that linear momentum flux was being imposed by scaling random
130     fluctuations of the center of the velocity distributions. Repeated
131     application of the original NIVS approach resulted in temperature
132     anisotropy, i.e. the width of the velocity distributions depended on
133     coordinates perpendicular to the desired gradient direction. For this
134     reason, combining thermal and momentum fluxes was difficult with the
135     original NIVS algorithm. However, combinations of thermal and
136     velocity gradients would provide a means to simulate thermal-linear
137     coupled processes such as Soret effect in liquid flows. Therefore,
138     developing improved approaches to the scaling imposed-flux methods
139     would be useful.
140 gezelter 3769
141 gezelter 3786 In this paper, we improve the RNEMD methods by introducing a novel
142     approach to creating imposed fluxes. This approach separates the
143     shearing and scaling of the velocity distributions in different
144     spatial regions and can apply both transformations within a single
145     time step. The approach retains desirable features of previous RNEMD
146     approaches and is simpler to implement compared to the earlier NIVS
147 gezelter 3788 method. In what follows, we first present the velocity shearing and
148     scaling (VSS) RNEMD method and its implementation in a
149     simulation. Then we compare the VSS-RNEMD method in bulk fluids to
150     previous methods. We also compute interfacial frictions are computed
151     for a series of heterogeneous interfaces.
152 gezelter 3769
153     \section{Methodology}
154 gezelter 3786 In an approach similar to the earlier NIVS method,\cite{kuang:164101}
155     we consider a periodic system which has been divided into a series of
156     slabs along a single axis (e.g. $z$). The unphysical thermal and
157     momentum fluxes are applied from one of the end slabs to the center
158     slab, and thus the thermal flux produces a higher temperature in the
159     center slab than in the end slab, and the momentum flux results in a
160     center slab moving along the positive $y$ axis (see Fig. \ref{cartoon}).
161     The applied fluxes can be set to negative values if the reverse
162     gradients are desired. For convenience the center slab is denoted as
163     the {\it hot} or {\it ``H''} slab, while the end slab is denoted {\it
164     ``C''} (or {\it cold}).
165 skuang 3770
166 gezelter 3786 \begin{figure}
167     \includegraphics[width=\linewidth]{cartoon}
168 gezelter 3788 \caption{The VSS-RNEMD approach imposes unphysical transfer of both
169     linear momentum and kinetic energy between a ``hot'' slab and a
170     ``cold'' slab in the simulation box. The system responds to this
171     imposed flux by generating both momentum and temperature gradients.
172     The slope of the gradients can then be used to compute transport
173     properties (e.g. shear viscosity and thermal conductivity).}
174 gezelter 3786 \label{cartoon}
175     \end{figure}
176    
177     To impose these fluxes, we periodically apply a set of operations on
178     the velocities of particles {$i$} within the cold slab and a separate
179     operation on particles {$j$} within the hot slab.
180 skuang 3770 \begin{eqnarray}
181     \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
182 gezelter 3788 \rangle\right) + \left(\langle\vec{v}_c\rangle +
183     \vec{a}_c\right) \label{eq:xformc} \\
184 skuang 3770 \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
185 gezelter 3788 \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right) \label{eq:xformh}
186 skuang 3770 \end{eqnarray}
187 gezelter 3786 where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denote
188     the instantaneous average velocity of the molecules within slabs $C$
189     and $H$ respectively. When a momentum flux $\vec{j}_z(\vec{p})$ is
190     present, these slab-averaged velocities also get corresponding
191     incremental changes ($\vec{a}_c$ and $\vec{a}_h$ respectively) that
192     are applied to all particles within each slab. The incremental
193     changes are obtained using Newton's second law:
194 skuang 3770 \begin{eqnarray}
195 gezelter 3786 M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \label{eq:newton1} \\
196 skuang 3770 M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
197 gezelter 3786 \label{eq:newton2}
198 skuang 3770 \end{eqnarray}
199 skuang 3781 where $M$ denotes total mass of particles within a slab:
200 skuang 3770 \begin{eqnarray}
201     M_c & = & \sum_{i = 1}^{N_c} m_i \\
202     M_h & = & \sum_{j = 1}^{N_h} m_j
203     \end{eqnarray}
204 skuang 3781 and $\Delta t$ is the interval between two separate operations.
205 skuang 3770
206 gezelter 3786 The operations in Eqs. \ref{eq:newton1} and \ref{eq:newton2} already
207     conserve the linear momentum of a periodic system. To further satisfy
208     total energy conservation as well as to impose the thermal flux $J_z$,
209     the following constraint equations must be solved for the two scaling
210     variables $c$ and $h$:
211 skuang 3770 \begin{eqnarray}
212     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
213     \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
214     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
215 gezelter 3786 \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2.
216     \label{constraint}
217 skuang 3770 \end{eqnarray}
218 gezelter 3786 Here $K_c$ and $K_h$ denote the translational kinetic energy of slabs
219     $C$ and $H$ respectively. These conservation equations are sufficient
220     to ensure total energy conservation, as the operations applied in our
221     method do not change the kinetic energy related to orientational
222     degrees of freedom or the potential energy of a system (as long as the
223     potential energy is independent of particle velocity).
224 skuang 3770
225 gezelter 3786 Equations \ref{eq:newton1}-\ref{constraint} are sufficient to
226     determine the velocity scaling coefficients ($c$ and $h$) as well as
227     $\vec{a}_c$ and $\vec{a}_h$. Note that there are usually two roots
228     respectively for $c$ and $h$. However, the positive roots (which are
229     closer to 1) are chosen so that the perturbations to the system are
230     minimal. Figure \ref{method} illustrates the implementation of this
231     algorithm in an individual step.
232 skuang 3770
233 skuang 3772 \begin{figure}
234 gezelter 3786 \centering
235     \includegraphics[width=5in]{method}
236 gezelter 3788 \caption{A single step implementation of the VSS algorithm starts with
237     the velocity distributions for the two slabs in a shearing fluid
238     (solid lines). Equations \ref{eq:xformc} and \ref{eq:xformh} are
239     used to scale and shear velocities in both slabs to mimic the effect
240     of both a thermal and a momentum flux. Gaussian distributions are
241 gezelter 3786 preserved by both the scaling and shearing operations.}
242 skuang 3772 \label{method}
243     \end{figure}
244    
245 gezelter 3786 By implementing these operations at a fixed frequency, stable thermal
246     and momentum fluxes can both be applied and the corresponding
247     temperature and momentum gradients can be established.
248 skuang 3770
249 gezelter 3788 Compared to the previous NIVS method, the VSS-RNEMD approach is
250 gezelter 3786 computationally simpler in that only quadratic equations are involved
251     to determine a set of scaling coefficients, while the NIVS method
252     required solution of quartic equations. Furthermore, this method
253     implements {\it isotropic} scaling of velocities in respective slabs,
254     unlike NIVS, which required extra flexibility in the choice of scaling
255     coefficients to allow for the energy and momentum constraints. Most
256     importantly, separating the linear momentum flux from velocity scaling
257     avoids the underlying cause of the thermal anisotropy in NIVS. In
258     later sections, we demonstrate that this can improve the calculated
259     shear viscosities from RNEMD simulations.
260 gezelter 3769
261 gezelter 3788 The VSS-RNEMD approach has advantages over the original momentum
262 gezelter 3786 swapping in many respects. In the swapping method, the velocity
263     vectors involved are usually very different (or the generated flux
264     would be quite small), thus the swapping tends to create strong
265     perturbations in the neighborhood of the particles involved. In
266 gezelter 3788 comparison, the VSS approach distributes the flux widely to all
267 gezelter 3786 particles in a slab so that perturbations in the flux generating
268     region are minimized. Additionally, momentum swapping steps tend to
269     result in nonthermal velocity distributions when the imposed flux is
270     large and diffusion from the neighboring slabs cannot carry momentum
271     away quickly enough. In comparison, the scaling and shearing moves
272 gezelter 3788 made in the VSS approach preserve the shapes of the equilibrium
273 gezelter 3786 velocity disributions (e.g. Maxwell-Boltzmann). The results presented
274     in later sections will illustrate that this is quite helpful in
275     retaining reasonable thermal distributions in a simulation.
276 gezelter 3769
277 skuang 3772 \section{Computational Details}
278 skuang 3773 The algorithm has been implemented in our MD simulation code,
279 gezelter 3786 OpenMD.\cite{Meineke:2005gd,openmd} We will first compare the method
280     with previous RNEMD methods and equilibrium MD in homogeneous
281     fluids (Lennard-Jones and SPC/E water). We have also used the new
282     method to simulate the interfacial friction of different heterogeneous
283     interfaces (Au (111) with organic solvents, Au(111) with SPC/E water,
284 gezelter 3788 and the ice Ih - liquid water interface).
285 gezelter 3769
286 skuang 3773 \subsection{Simulation Protocols}
287 gezelter 3786 The systems we investigated were set up in orthorhombic simulation
288 skuang 3784 cells with periodic boundary conditions in all three dimensions. The
289 gezelter 3786 $z$-axes of these cells were typically quite long and served as the
290     temperature and/or momentum gradient axes. The cells were evenly
291     divided into $N$ slabs along this axis, with $N$ varying for the
292     individual system. The $x$ and $y$ axes were of similar lengths in all
293     simulations. In all cases, before introducing a nonequilibrium method
294     to establish steady thermal and/or momentum gradients, equilibration
295     simulations were run under the canonical ensemble with a Nos\'e-Hoover
296     thermostat\cite{hoover85} followed by further equilibration using
297     standard constant energy (NVE) conditions. For SPC/E water
298     simulations, isobaric-isothermal equilibrations\cite{melchionna93}
299     were performed before equilibration to reach standard densities at
300     atmospheric pressure (1 bar); for interfacial systems, similar
301     equilibrations with anisotropic box relaxations are used to relax the
302     surface tension in the $xy$ plane.
303 skuang 3772
304 gezelter 3786 While homogeneous fluid systems can be set up with random
305     configurations, interfacial systems are typically prepared with a
306     single crystal face presented perpendicular to the $z$-axis. This
307     crystal face is aligned in the x-y plane of the periodic cell, and the
308     solvent occupies the region directly above and below a crystalline
309 gezelter 3791 slab. The preparation and equilibration of solvated Au(111) surfaces,
310     as well as the solvation and equilibration processes used for these
311     interfaces are described in detail in reference \cite{kuang:AuThl}.
312 skuang 3773
313 gezelter 3786 For the ice / liquid water interfaces, the basal surface of ice
314     lattice was first constructed. Hirsch and Ojam\"{a}e
315     \cite{doi:10.1021/jp048434u} explored the energetics of ice lattices
316     with all possible proton ordered configurations. We utilized Hirsch
317     and Ojam\"{a}e's structure 6 ($P2_12_12_1$) which is an orthorhombic
318 gezelter 3788 cell that produces a proton-ordered version of ice Ih. The basal face
319     of ice in this unit cell geometry is the $\{0~0~1\}$ face. Although
320 skuang 3784 experimental solid/liquid coexistant temperature under normal pressure
321 gezelter 3786 are close to 273K, Bryk and Haymet's simulations of ice/liquid water
322     interfaces with different models suggest that for SPC/E, the most
323     stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
324     Therefore, our ice/liquid water simulations were carried out at an
325     average temperature of 225K. Molecular translation and orientational
326     restraints were applied in the early stages of equilibration to
327     prevent melting of the ice slab. These restraints were removed during
328     NVT equilibration, well before data collection was carried out.
329 gezelter 3769
330     \subsection{Force Field Parameters}
331 gezelter 3788 For comparing the VSS-RNEMD method with previous work, we retained
332 skuang 3784 force field parameters consistent with previous simulations. Argon is
333     the Lennard-Jones fluid used here, and its results are reported in
334 gezelter 3786 reduced units for purposes of direct comparison with previous
335     calculations.
336 gezelter 3769
337 gezelter 3786 For our water simulations, we utilized the SPC/E model throughout this
338     work. Previous work for transport properties of SPC/E water model is
339     available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct
340     comparison with previous calculation methods is possible.
341 gezelter 3769
342 skuang 3774 The Au-Au interaction parameters in all simulations are described by
343     the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
344     QSC potentials include zero-point quantum corrections and are
345 gezelter 3769 reparametrized for accurate surface energies compared to the
346 skuang 3775 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
347     Spohr potential was adopted\cite{ISI:000167766600035} to depict
348     Au-H$_2$O interactions.
349 gezelter 3769
350 gezelter 3791 For our gold / organic liquid interfaces, the small-molecule organic
351     solvents included in our simulations are hexane and toluene. The
352     United-Atom models\cite{TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes} for
353     these components were used in this work for computational efficiency,
354     while maintaining good accuracy. We refer readers to our previous
355     work\cite{kuang:AuThl} for further details of these models, as well as
356     the interactions between Au and the above organic molecule components.
357 gezelter 3769
358 skuang 3774 \subsection{Thermal conductivities}
359 gezelter 3786 When the linear momentum flux $\vec{j}_z(\vec{p})$ is set to zero and
360 gezelter 3788 the target $J_z$ is non-zero, VSS-RNEMD imposes kinetic energy
361     transfer between the slabs, which can be used for computation of
362     thermal conductivities. As with other RNEMD methods, we assume that we
363 gezelter 3786 are in the linear response regime of the temperature gradient with
364     respect to the thermal flux. The thermal conductivity ($\lambda$) can
365 gezelter 3788 then be calculated using the imposed kinetic energy flux and the
366     measured thermal gradient:
367 skuang 3775 \begin{equation}
368     J_z = -\lambda \frac{\partial T}{\partial z}
369     \end{equation}
370     Like other imposed-flux methods, the energy flux was calculated using
371     the total non-physical energy transferred (${E_{total}}$) from slab
372 gezelter 3788 {\it C} to slab {\it H}, which was recorded throughout the simulation, and
373 gezelter 3786 the total data collection time $t$:
374 skuang 3775 \begin{equation}
375 gezelter 3786 J_z = \frac{E_{total}}{2 t L_x L_y}.
376 skuang 3775 \end{equation}
377 gezelter 3788 Here, $L_x$ and $L_y$ denote the dimensions of the plane in the
378     simulation cell that is perpendicular to the thermal gradient, and a
379     factor of two in the denominator is necessary as the heat transport
380     occurs in both the $+z$ and $-z$ directions. The average temperature
381     gradient ${\langle\partial T/\partial z\rangle}$ can be obtained by a
382     linear regression of the temperature profile, which is recorded during
383     a simulation for each slab in a cell. For Lennard-Jones simulations,
384 skuang 3775 thermal conductivities are reported in reduced units
385     (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
386    
387 skuang 3774 \subsection{Shear viscosities}
388 gezelter 3786 Alternatively, when the linear momentum flux $\vec{j}_z(\vec{p})$ is
389     non-zero in either the $x$ or $y$ directions, the method can be used
390     to compute the shear viscosity. For isotropic systems, the direction
391     of $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
392     ability to arbitarily specify the vector direction in our method could
393 gezelter 3788 be convenient when working with anisotropic interfaces.
394 skuang 3775
395 gezelter 3786 In a manner similar to the thermal conductivity calculations, a linear
396     response of the momentum gradient with respect to the shear stress is
397     assumed, and the shear viscosity ($\eta$) can be obtained with the
398     imposed momentum flux (e.g. in $x$ direction) and the measured
399     velocity gradient:
400 skuang 3775 \begin{equation}
401     j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
402     \end{equation}
403     where the flux is similarly defined:
404     \begin{equation}
405     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
406     \end{equation}
407     with $P_x$ being the total non-physical momentum transferred within
408 skuang 3784 the data collection time. Also, the averaged velocity gradient
409 skuang 3775 ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
410 skuang 3784 regression of the $x$ component of the mean velocity ($\langle
411     v_x\rangle$) in each of the bins. For Lennard-Jones simulations, shear
412     viscosities are also reported in reduced units
413 skuang 3775 (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
414    
415 gezelter 3786 Although $J_z$ may be switched off for shear viscosity simulations,
416 gezelter 3788 the VSS-RNEMD method allows the user the ability to {\it
417     simultaneously} impose both a thermal and a momentum flux during a
418     single simulation. This creates configurations with coincident
419     temperature and a velocity gradients. Since the viscosity is
420     generally a function of temperature, the local viscosity depends on
421     the local temperature in the fluid. Therefore, in a single simulation,
422     viscosity at $z$ (corresponding to a certain $T$) can be computed with
423     the applied shear flux and the local velocity gradient (which can be
424     obtained by finite difference approximation). This means that the
425     temperature dependence of the viscosity can be mapped out in only one
426 gezelter 3786 trajectory. Results for shear viscosity computations of SPC/E water
427 gezelter 3788 will demonstrate VSS-RNEMD's efficiency in this respect.
428 skuang 3784
429 gezelter 3786 \subsection{Interfacial friction and slip length}
430     Shear stress creates a velocity gradient within bulk fluid phases, but
431     at solid-liquid interfaces, the effects of a shear stress depend on
432     the molecular details of the interface. The interfacial friction
433     coefficient, $\kappa$, relates the shear stress (e.g. along the
434 gezelter 3788 $x$-axis) with the relative velocity of the fluid tangent to the
435     interface:
436 skuang 3775 \begin{equation}
437 gezelter 3786 j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right]
438 skuang 3775 \end{equation}
439 gezelter 3786 where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured
440     directly adjacent to the interface. Under ``stick'' boundary
441     condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads
442     to $\kappa\rightarrow\infty$. However, for ``slip'' boundary
443     conditions at a solid-liquid interface, $\kappa$ becomes finite. To
444     characterize the interfacial boundary conditions, the slip length,
445     $\delta$, is defined by the ratio of the fluid-phase viscosity to the
446     friction coefficient of the interface:
447 skuang 3775 \begin{equation}
448     \delta = \frac{\eta}{\kappa}
449     \end{equation}
450 gezelter 3788 In ``stick'' boundary conditions, $\delta\rightarrow 0$, so $\delta$
451     is a measure of how ``slippery'' an interface is. Figure
452     \ref{slipLength} illustrates how this quantity is defined and computed
453     for a solid-liquid interface.
454 gezelter 3769
455 skuang 3775 \begin{figure}
456     \includegraphics[width=\linewidth]{defDelta}
457 gezelter 3788 \caption{The slip length ($\delta$) can be obtained from a velocity
458     profile along an axis perpendicular to the interface. The data shown
459     is for an Au / hexane interface -- the crystalline region (Au) is
460     moving as a block (lower dots), while the measured velocity gradient
461     in the hexane phase is discontinuous at the interface.}
462 skuang 3775 \label{slipLength}
463     \end{figure}
464 gezelter 3769
465 gezelter 3788 Since the VSS-RNEMD method can be applied for interfaces as well as
466     for bulk materials, the shear stress is applied in an identical manner
467     to the shear viscosity computations, e.g. by applying an unphysical
468     momentum flux, $j_z(\vec{p})$. With the correct choice of $\vec{p}$
469     in the $x-y$ plane, one can compute friction coefficients and slip
470     lengths for a number of different dragging vectors on a given slab.
471     The corresponding velocity profiles can be obtained as shown in Figure
472 gezelter 3786 \ref{slipLength}, in which the velocity gradients within the liquid
473     phase and the velocity difference at the liquid-solid interface can be
474     easily measured from saved simulation data.
475 skuang 3775
476 skuang 3776 \section{Results and Discussions}
477     \subsection{Lennard-Jones fluid}
478 gezelter 3786 Our orthorhombic simulation cell for the Lennard-Jones fluid has
479 gezelter 3788 identical parameters to previous work to facilitate
480     comparison.\cite{kuang:164101} Thermal conductivities and shear
481 gezelter 3789 viscosities were computed with the new VSS algorithm, and the results
482     were compared with the previous NIVS algorithm. However, since the
483     NIVS algorithm produces temperature anisotropy under shear stress,
484     these results are also compared to the previous momentum swapping
485 gezelter 3788 approaches. Table \ref{LJ} lists these values with various fluxes in
486     reduced units.
487 skuang 3770
488 gezelter 3791 \begin{table}
489     \tbl{Thermal conductivity ($\lambda^*$) and shear viscosity
490     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
491     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
492     at various momentum fluxes. The new VSS method yields similar
493     results to previous RNEMD methods. All results are reported in
494     reduced unit. Uncertainties are indicated in parentheses.}
495 skuang 3776
496 gezelter 3791 {\begin{tabular}{cc|ccc|cc}\toprule
497     Swap & Equivalent &
498     \multicolumn{3}{|c}{$\lambda^*$} &
499     \multicolumn{2}{|c}{$\eta^*$} \\
500     Interval & $J_z^*$ or $j_z^*(p_x)$ & Swapping &
501 skuang 3785 NIVS\cite{kuang:164101} & This work & Swapping & This work \\
502 gezelter 3791 \colrule
503 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)\\
504     0.232 & 0.09 & 7.03(0.14) & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
505     0.463 & 0.047 & 6.91(0.42) & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
506     0.926 & 0.024 & 7.52(0.15) & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
507     1.158 & 0.019 & 7.41(0.29) & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
508 gezelter 3791 \botrule
509     \end{tabular}}
510     \label{LJ}
511     \end{table}
512 gezelter 3769
513 skuang 3776 \subsubsection{Thermal conductivity}
514 gezelter 3789 Our thermal conductivity calculations yield comparable results to the
515     previous NIVS algorithm. This indicates that the thermal gradients
516     produced using this method are also quite close to previous RNEMD
517     methods. Simulations with moderately large thermal fluxes tend to
518     yield more reliable thermal gradients than under NIVS and thus avoid
519     large errors. Extreme values for the applied thermal flux can
520     introduce side effects such as non-linear temperature gradients and
521     inadvertent phase transitions, so these are avoided.
522 gezelter 3769
523 gezelter 3789 Since the scaling operation is isotropic in VSS, the temperature
524     anisotropy that was observed under the earlier NIVS approach should be
525     absent. Furthermore, the VSS method avoids unintended momentum flux
526     when only thermal flux is being imposed. This was not always possible
527     with swapping or NIVS approaches. The thermal energy exchange in
528     swapping ($\vec{p}_i$ in slab {\it C} swapped with $\vec{p}_j$ in slab
529     {\it H}) or NIVS (total slab momentum components $P^\alpha$ scaled to
530     $\alpha P^\alpha$) do not achieve this unless thermal flux vanishes
531     (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$). In this sense, the VSS
532     method achieves minimal perturbation to a simulation while imposing
533     thermal flux.
534 skuang 3776
535     \subsubsection{Shear viscosity}
536 skuang 3785 Table \ref{LJ} also compares our shear viscosity results with the
537 gezelter 3789 momentum swapping approach. Our calculations show that the VSS method
538     predicted similar values for shear viscosities to the momentum
539     swapping approach, as well as providing similar velocity gradient
540     profiles. Moderately large momentum fluxes are helpful in reducing the
541     errors in measured velocity gradients and thus the shear viscosity
542     values. However, it should be noted that the momentum swapping
543     approach tends to produce non-thermal velocity distributions in the
544     two slabs.\cite{Maginn:2010}
545 skuang 3776
546 gezelter 3789 To show that the temperature is isotropic within each slab under VSS,
547     we measured the three one-dimensional temperatures in each of the
548     slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
549 skuang 3785 temperatures were calculated after subtracting the contribution from
550 gezelter 3789 bulk (shearing) velocities of the slabs. The one-dimensional
551     temperature profiles show no observable differences between the three
552     box dimensions. This ensures that VSS automatically preserves
553     temperature isotropy during shear viscosity calculations.
554 skuang 3776
555 gezelter 3769 \begin{figure}
556 skuang 3776 \includegraphics[width=\linewidth]{tempXyz}
557 gezelter 3788 \caption{Unlike the previous NIVS algorithm, the VSS method does not
558     produce thermal anisotropy under shearing stress. Temperature
559     differences along the different box axes were significantly smaller
560     than the error bars. Note that the two regions of elevated
561     temperature are caused by the shear stress. This is the same
562     frictional heating reported previously by Tenney and
563     Maginn.\cite{Maginn:2010}}
564 skuang 3776 \label{tempXyz}
565 gezelter 3769 \end{figure}
566    
567 gezelter 3789 The velocity distribution profiles were further tested by imposing a
568     large shear stress. Figure \ref{vDist} demonstrates how the VSS method
569     is able to maintain nearly ideal Maxwell-Boltzmann velocity
570     distributions even under large imposed fluxes. Previous swapping
571     methods tend to deplete particles with positive velocities in the
572     negative velocity slab ({\it C}) and deplete particles with negative
573     velocities in the {\it H} slab, leaving significant `notches' in the
574     velocity distributions. The problematic velocity distributions become
575     significant when the imposed-flux becomes so large that diffusion from
576     neighboring slabs cannot offset the depletion. Simultaneously,
577     abnormal peaks appear in the other regions of the velocity
578     distributions as high (or low) momentum particles are introduced in to
579     the corresponding slab. These nonthermal distributions limit
580     application of the swapping approach in shear stress simulations. The
581     VSS method avoids the above problematic distributions by altering the
582     means of applying momentum flux. Comparatively, velocity distributions
583     recorded from simulations with the VSS method is so close to the ideal
584     Maxwell-Boltzmann distributions that no obvious difference is visible
585     in Figure \ref{vDist}.
586 gezelter 3769
587 skuang 3776 \begin{figure}
588 gezelter 3788 \centering
589     \includegraphics[width=5.5in]{velDist}
590     \caption{Velocity distributions that develop under VSS more closely
591     resemble ideal Maxwell-Boltzmann distributions than earlier
592     momentum-swapping RNEMD approaches. This data was obtained from
593 skuang 3779 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
594 skuang 3777 swapping interval of 20 time steps). This is a relatively large flux
595 gezelter 3788 which demonstrates some of the non-thermal velocity distributions
596     that can develop under swapping RNEMD.}
597 skuang 3776 \label{vDist}
598     \end{figure}
599 gezelter 3769
600 skuang 3776 \subsection{Bulk SPC/E water}
601 gezelter 3789 We also computed the thermal conductivity and shear viscosity of SPC/E
602     water. A simulation cell with 1000 molecules was set up in a similar
603     manner to Ref. \cite{kuang:164101}. For thermal conductivity
604     calculations, measurements were taken to compare with previous RNEMD
605     methods; for shear viscosity computations, simulations were run under
606     a series of temperatures (with corresponding pressure relaxation using
607     the isobaric-isothermal ensemble\cite{melchionna93}), and results were
608     compared to available data from equilibrium molecular dynamics
609     calculations.\cite{10.1063/1.3330544,Medina2011} Additionally, a
610     simulation with {\it both} applied thermal and momentum gradients was
611     carried out to map out shear viscosity as a function of temperature.
612 skuang 3777
613 skuang 3776 \subsubsection{Thermal conductivity}
614 gezelter 3789 Table \ref{spceThermal} summarizes the thermal conductivity of SPC/E
615     under different temperatures and thermal gradients compared with the
616     previous NIVS results\cite{kuang:164101} and experimental
617     measurements.\cite{WagnerKruse} Note that no appreciable drift of
618     total system energy or temperature was observed when the VSS method
619     was applied, which indicates that our algorithm conserves total energy
620     well for systems involving electrostatic interactions.
621 gezelter 3769
622 gezelter 3789 Measurements using the VSS method established similar temperature
623 skuang 3778 gradients to the previous NIVS method. Our simulation results are in
624 gezelter 3789 fairly good agreement with those from previous simulations. Both
625     methods yield values in reasonable agreement with experimental
626     values. Simulations using larger thermal gradients and those with
627     longer gradient axes ($z$) have less measurable noise.
628 skuang 3778
629 gezelter 3791 \begin{table}
630    
631     \tbl{Thermal conductivity of SPC/E water under various
632     imposed thermal gradients. Uncertainties are indicated in
633     parentheses.}
634    
635     {\begin{tabular}{cc|ccc}\toprule
636 skuang 3778 $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
637     {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
638 gezelter 3789 (K) & (K/\AA) & This work & NIVS\cite{kuang:164101} &
639 skuang 3778 Experiment\cite{WagnerKruse} \\
640 gezelter 3791 \colrule
641     300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\ \colrule
642 skuang 3778 318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
643 gezelter 3789 & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
644 gezelter 3791 & 0.8 & 0.786(0.009)$^{\rm a}$ & & \\
645     \botrule
646     \end{tabular}}
647     \tabnote{$^{\rm a}$Simulation with 2000 SPC/E molecules}
648     \label{spceThermal}
649     \end{table}
650 skuang 3778
651 skuang 3776 \subsubsection{Shear viscosity}
652 gezelter 3789 Shear viscosity was computed for SPC/E water at a range of
653     temperatures that span the liquidus range for water under atmospheric
654     pressure. VSS-RNEMD simulations predict a similar trend of $\eta$
655     vs. $T$ to equilibrium molecular dynamics (EMD) results and are
656     presented in table \ref{spceShear}. Our results show no significant
657     differences from the earlier EMD calculations. Since each value
658 gezelter 3791 reported using our method takes a single 1.5 ns trajectory, instead of
659     average from many trajectories when using EMD, the VSS method provides
660     an efficient means for computing the shear viscosity.
661 gezelter 3769
662 gezelter 3791 \begin{table}
663     \tbl{Computed shear viscosity of SPC/E water under different
664 gezelter 3789 temperatures. Results are compared to those obtained with
665     equilibrium molecular dynamics.\cite{Medina2011} Uncertainties
666     are indicated in parentheses.}
667 skuang 3778
668 gezelter 3791 {\begin{tabular}{cc|cc}\toprule
669 skuang 3778 $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
670     {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
671 skuang 3785 (K) & (10$^{10}$s$^{-1}$) & This work & Previous
672     simulations\cite{Medina2011} \\
673 gezelter 3791 \colrule
674 skuang 3785 273 & 1.12 & 1.218(0.004) & 1.282(0.048) \\
675     & 1.79 & 1.140(0.012) & \\
676 gezelter 3791 \colrule
677 skuang 3785 303 & 2.09 & 0.646(0.008) & 0.643(0.019) \\
678 gezelter 3791 \colrule
679 skuang 3785 318 & 2.50 & 0.536(0.007) & \\
680     & 5.25 & 0.510(0.007) & \\
681 gezelter 3791 & 2.82 & 0.474(0.003)$^{\rm a}$ & \\
682     \colrule
683 skuang 3785 333 & 3.10 & 0.428(0.002) & 0.421(0.008) \\
684 gezelter 3791 \colrule
685 skuang 3785 363 & 2.34 & 0.279(0.014) & 0.291(0.005) \\
686     & 4.26 & 0.306(0.001) & \\
687 gezelter 3791 \botrule
688     \end{tabular}}
689     \tabnote{$^{\rm a}$Simulation with 2000 SPC/E molecules}\label{spceShear}
690     \end{table}
691 gezelter 3769
692 gezelter 3791 A much more efficient way to map out $\eta$ vs $T$ is to combine the
693     momentum flux with a thermal flux. Figure \ref{Tvxdvdz} shows the
694     thermal and velocity gradient in one such simulation. The local
695     temperature varies along the $z$-axis, and the velocity gradient can
696     be computed locally via finite difference approximations. With the
697     data provided in Figure \ref{Tvxdvdz}, the entire temperature
698     dependent viscosity $\eta(T)$ can be calculated from a {\it single} 2
699     ns trajectory. This data is shown in figure \ref{etaT}. A linear fit
700     can be used to approximate $\partial v_x/\partial z$ vs. $z$ and the
701     fit velocity gradient produced the solid curve in Fig. \ref{etaT}.
702 skuang 3785
703     \begin{figure}
704 gezelter 3788 \centering
705     \includegraphics[width=5.5in]{tvxdvdz}
706 skuang 3785 \caption{With a combination of a thermal and a momentum flux, a
707 gezelter 3788 simulation can exhibit both a temperature (top) and a velocity
708     (middle) gradient. $\partial v_x/\partial z$ depends on the local
709     temperature, so it varies along the gradient axis. This derivative
710     can be computed using finite difference approximations (lower) and
711     can be used to calculate $\eta$ vs $T$ (Figure \ref{etaT}).}
712 skuang 3785 \label{Tvxdvdz}
713     \end{figure}
714    
715 gezelter 3791 In Figure \ref{etaT}, the fit gradient yields viscosity values that
716     are in agreement with VSS-RNEMD simulations carried out at fixed
717     temperatures, as well as results reported using EMD methods through
718     most of the simulated temperature
719     range.\cite{10.1063/1.3330544,Medina2011} However, there is larger
720     error in lower temperature regions and the single-trajectory approach
721     has difficulty predicting $\eta$ near 273K. Longer trajectories and
722     larger box geometries would be required to converge these values with
723     single-temperature calculations. Since previous work already pointed
724     out that SPC/E water tends to yield lower viscosities than
725     experimental data,\cite{Medina2011} experimental comparisons are not
726     given here.
727 skuang 3785
728     \begin{figure}
729     \includegraphics[width=\linewidth]{etaT}
730 gezelter 3788 \caption{The temperature dependence of the shear viscosity generated
731     by a {\it single} VSS-RNEMD simulation. Applying both thermal and
732     momentum gradient predicts reasonable values in much of the
733     temperature range being tested.}
734 skuang 3785 \label{etaT}
735     \end{figure}
736    
737 gezelter 3791 \subsection{Interfacial friction and slip length}
738     Another attractive aspect of the VSS-RNEMD method is the ability to
739     apply shear stress to interfacial and heterogeneous systems, where
740 skuang 3785 molecules of different identities (or phases) are segregated in
741 gezelter 3791 different regions of space. We have previously studied the interfacial
742     thermal conductivity of a series of metal-liquid
743     interfaces,\cite{kuang:164101,kuang:AuThl} and would like to further
744 skuang 3785 investigate the relationship between this phenomenon and the
745 gezelter 3791 interfacial friction. Knowing the friction coefficients and slip
746     lengths of interfaces could also help predict the mass- and
747     heat-transport properties of colloidal suspensions of nanocrystalline
748     materials.
749 gezelter 3769
750 gezelter 3791 Table \ref{etaKappaDelta} includes results of a few model interfaces.
751     For bare Au(111) surfaces, slip boundary conditions were observed for
752     both organic and aqueous liquid phases, corresponding to previously
753     computed low interfacial thermal conductance. This supports
754     discussions on the relationship between surface wetting and slip
755     effect and thermal conductance of the
756 skuang 3785 interface.\cite{PhysRevLett.82.4671,doi:10.1080/0026897031000068578,garde:PhysRevLett2009}
757 skuang 3776
758 gezelter 3791 \begin{table}
759     \tbl{Computed interfacial friction coefficients and slip
760     lengths for various solid -- liquid interfaces. All of the
761     solvated bare metal surfaces appear to exist in the ``slip''
762     boundary conditions under shear stress, while the ice/water
763     interface is in the opposite or ``stick'' boundary
764     limit. Error estimates are indicated in parentheses.}
765    
766     {\begin{tabular}{ccccccc}
767     \toprule
768 skuang 3779 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
769 gezelter 3791 & $\delta$ \\
770     surface & phase & K & MPa & mPa$\cdot$s &
771     10$^4$Pa$\cdot$s/m & nm \\
772     \colrule
773 skuang 3785 Au(111) & hexane & 200 & 1.08 & 0.197(0.009) & 5.30(0.36) &
774 gezelter 3791 3.72(0.25) \\
775 skuang 3785 & & & 2.15 & 0.141(0.002) & 5.31(0.26) &
776 gezelter 3791 2.76(0.14) \\
777     \colrule
778 skuang 3785 Au(111) & toluene & 200 & 1.08 & 0.722(0.035) & 15.7(0.7) &
779 gezelter 3791 4.60(0.22) \\
780 skuang 3785 & & & 2.16 & 0.544(0.030) & 11.2(0.5) &
781 gezelter 3791 4.86(0.27) \\
782     \colrule
783 skuang 3785 Au(111) & water & 300 & 1.08 & 0.399(0.050) & 1.928(0.022) &
784 gezelter 3791 20.7(2.6) \\
785 skuang 3785 & & & 2.16 & 0.794(0.255) & 1.895(0.003) &
786 gezelter 3791 41.9(13.5) \\
787     \colrule
788     ice Ih (basal) & water & 225 & 19.4 & 15.8(0.2) & $\infty$ & 0 \\
789     \botrule
790     \end{tabular}}
791     \label{etaKappaDelta}
792     \end{table}
793 gezelter 3769
794 gezelter 3791 In addition to these previously studied interfaces, we also
795     constructed ice Ih (basal face) -- water interfaces. In contrast with
796     the Au(111) -- water interface, where the friction coefficient is
797     small and large slip effect is present, the ice -- water interface
798     demonstrates strong solid-liquid interactions and appears to be in the
799     stick boundary limit. The supercooled liquid phase is an order of
800     magnitude more viscous than measurements of liquid water done in the
801     previous section. It would be of interst to investigate the effects of
802     different ice lattice planes (e.g. the prism, secondary prism, and
803     pyramidal faces of ice) on interfacial friction and the corresponding
804     liquid viscosity.
805 gezelter 3769
806 skuang 3776 \section{Conclusions}
807 gezelter 3791 We have developed a novel method for simultaneously applying both
808     shear stress and thermal flux during molecular dynamics
809     simulations. This method, the VSS-RNEMD approach, enables efficient
810     computation of both thermal conductivity and shear viscosity in bulk
811     liquids. The method maintains thermal velocity distributions and
812     avoids thermal anisotropy in previous NIVS shear stress simulations,
813     while retaining the attractive features of previous RNEMD
814     methods. There is no {\it a priori} restrictions on the method to
815     particular statistical ensembles, so use of this method within
816     extended-system simulations (to maintain constant temperature and
817     pressure) is also possible.
818 gezelter 3769
819 gezelter 3791 Most importantly, the VSS-RNEMD method is capable of simultaneously
820     imposing thermal flux and shear stress accross a heterogeneous
821     interface. This will facilitate the study of interfacial properties
822     that depend on the chemical details of the interface. Further
823     investigations will be carried out to characterize interfacial
824     friction behavior using the method.
825 gezelter 3769
826 gezelter 3791 Since VSS-RNEMD can simultaneously introduce thermal and momentum
827     gradients, it can also be used to efficiently map out the viscosity as
828     a function of temperature with a single simulation. Complex systems
829     that involve thermal and momentum gradients might potentially benefit
830     from this feature. For example, the behavior of the Soret effect
831     (solute migration across a thermal gradient) in a solution flowing
832     through a medium would be of interest to those simulating purification
833     and separation processes.
834 gezelter 3769
835 gezelter 3791 \section*{Acknowledgments}
836 gezelter 3769 Support for this project was provided by the National Science
837     Foundation under grant CHE-0848243. Computational time was provided by
838     the Center for Research Computing (CRC) at the University of Notre
839     Dame.
840    
841     \newpage
842    
843 skuang 3770 \bibliography{stokes}
844 gezelter 3769
845 gezelter 3791 %\end{doublespace}
846 gezelter 3769 \end{document}