ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3615
Committed: Mon Jul 19 19:59:11 2010 UTC (14 years, 1 month ago) by skuang
Content type: application/x-tex
File size: 41013 byte(s)
Log Message:
changes to shear's legends. new results added. and other minor changes.

File Contents

# User Rev Content
1 gezelter 3524 \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 gezelter 3609 \usepackage{multirow}
10 gezelter 3524 %\usepackage{booktabs}
11     %\usepackage{bibentry}
12     %\usepackage{mathrsfs}
13     \usepackage[ref]{overcite}
14     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
15     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
16     9.0in \textwidth 6.5in \brokenpenalty=10000
17    
18     % double space list of tables and figures
19     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
20     \setlength{\abovecaptionskip}{20 pt}
21     \setlength{\belowcaptionskip}{30 pt}
22    
23     \renewcommand\citemid{\ } % no comma in optional referenc note
24    
25     \begin{document}
26    
27     \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
28    
29     \author{Shenyu Kuang and J. Daniel
30     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
31     Department of Chemistry and Biochemistry,\\
32     University of Notre Dame\\
33     Notre Dame, Indiana 46556}
34    
35     \date{\today}
36    
37     \maketitle
38    
39     \begin{doublespace}
40    
41     \begin{abstract}
42 gezelter 3583 We present a new method for introducing stable non-equilibrium
43     velocity and temperature distributions in molecular dynamics
44 gezelter 3609 simulations of heterogeneous systems. This method extends earlier
45     Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
46     momentum exchange swapping moves that can create non-thermal
47     velocity distributions and are difficult to use for interfacial
48     calculations. By using non-isotropic velocity scaling (NIVS) on the
49     molecules in specific regions of a system, it is possible to impose
50     momentum or thermal flux between regions of a simulation and stable
51     thermal and momentum gradients can then be established. The scaling
52     method we have developed conserves the total linear momentum and
53     total energy of the system. To test the methods, we have computed
54     the thermal conductivity of model liquid and solid systems as well
55     as the interfacial thermal conductivity of a metal-water interface.
56     We find that the NIVS-RNEMD improves the problematic velocity
57     distributions that develop in other RNEMD methods.
58 gezelter 3524 \end{abstract}
59    
60     \newpage
61    
62     %\narrowtext
63    
64     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65     % BODY OF TEXT
66     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67    
68     \section{Introduction}
69     The original formulation of Reverse Non-equilibrium Molecular Dynamics
70     (RNEMD) obtains transport coefficients (thermal conductivity and shear
71     viscosity) in a fluid by imposing an artificial momentum flux between
72     two thin parallel slabs of material that are spatially separated in
73 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 gezelter 3583 artificial flux is typically created by periodically ``swapping''
75     either the entire momentum vector $\vec{p}$ or single components of
76     this vector ($p_x$) between molecules in each of the two slabs. If
77     the two slabs are separated along the $z$ coordinate, the imposed flux
78     is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
79     response of a simulated system to the imposed momentum flux will
80     typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
81     The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
82     easily obtained by assuming linear response of the system,
83 gezelter 3524 \begin{eqnarray}
84 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
86 gezelter 3524 \end{eqnarray}
87 gezelter 3600 RNEMD has been widely used to provide computational estimates of
88     thermal conductivities and shear viscosities in a wide range of
89     materials, from liquid copper to both monatomic and molecular fluids
90     (e.g. ionic
91     liquids).\cite{Bedrov:2000-1,Bedrov:2000,Muller-Plathe:2002,ISI:000184808400018,ISI:000231042800044,Maginn:2007,Muller-Plathe:2008,ISI:000258460400020,ISI:000258840700015,ISI:000261835100054}
92 gezelter 3524
93 skuang 3574 \begin{figure}
94     \includegraphics[width=\linewidth]{thermalDemo}
95 gezelter 3583 \caption{RNEMD methods impose an unphysical transfer of momentum or
96     kinetic energy between a ``hot'' slab and a ``cold'' slab in the
97     simulation box. The molecular system responds to this imposed flux
98     by generating a momentum or temperature gradient. The slope of the
99     gradient can then be used to compute transport properties (e.g.
100     shear viscosity and thermal conductivity).}
101 skuang 3574 \label{thermalDemo}
102     \end{figure}
103    
104 skuang 3591 RNEMD is preferable in many ways to the forward NEMD
105 skuang 3592 methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
106     because it imposes what is typically difficult to measure (a flux or
107 gezelter 3600 stress) and it is typically much easier to compute the response
108 gezelter 3609 (momentum gradients or strains). For similar reasons, RNEMD is also
109 skuang 3592 preferable to slowly-converging equilibrium methods for measuring
110     thermal conductivity and shear viscosity (using Green-Kubo
111 skuang 3591 relations\cite{daivis:541,mondello:9327} or the Helfand moment
112     approach of Viscardy {\it et
113 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
114 gezelter 3524 computing difficult to measure quantities.
115    
116     Another attractive feature of RNEMD is that it conserves both total
117     linear momentum and total energy during the swaps (as long as the two
118     molecules have the same identity), so the swapped configurations are
119     typically samples from the same manifold of states in the
120     microcanonical ensemble.
121    
122 skuang 3588 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
123 skuang 3565 some problems with the original RNEMD swap technique. Notably, large
124     momentum fluxes (equivalent to frequent momentum swaps between the
125 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
126     momentum distributions in the two slabs, as well as non-linear thermal
127     and velocity distributions along the direction of the imposed flux
128     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
129     and self-adjusting metrics for retaining the usability of the method.
130 gezelter 3524
131     In this paper, we develop and test a method for non-isotropic velocity
132 gezelter 3600 scaling (NIVS) which retains the desirable features of RNEMD
133 gezelter 3524 (conservation of linear momentum and total energy, compatibility with
134     periodic boundary conditions) while establishing true thermal
135 gezelter 3600 distributions in each of the two slabs. In the next section, we
136 gezelter 3583 present the method for determining the scaling constraints. We then
137 gezelter 3600 test the method on both liquids and solids as well as a non-isotropic
138     liquid-solid interface and show that it is capable of providing
139 gezelter 3524 reasonable estimates of the thermal conductivity and shear viscosity
140 gezelter 3600 in all of these cases.
141 gezelter 3524
142     \section{Methodology}
143 gezelter 3583 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
144     periodic system is partitioned into a series of thin slabs along one
145 gezelter 3524 axis ($z$). One of the slabs at the end of the periodic box is
146     designated the ``hot'' slab, while the slab in the center of the box
147     is designated the ``cold'' slab. The artificial momentum flux will be
148     established by transferring momentum from the cold slab and into the
149     hot slab.
150    
151     Rather than using momentum swaps, we use a series of velocity scaling
152 gezelter 3583 moves. For molecules $\{i\}$ located within the cold slab,
153 gezelter 3524 \begin{equation}
154 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
155     x & 0 & 0 \\
156     0 & y & 0 \\
157     0 & 0 & z \\
158 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
159     \end{equation}
160 gezelter 3600 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
161     of the three directions in the system. Likewise, the molecules
162     $\{j\}$ located in the hot slab will see a concomitant scaling of
163     velocities,
164 gezelter 3524 \begin{equation}
165 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
166     x^\prime & 0 & 0 \\
167     0 & y^\prime & 0 \\
168     0 & 0 & z^\prime \\
169 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
170     \end{equation}
171    
172     Conservation of linear momentum in each of the three directions
173 gezelter 3583 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
174 gezelter 3524 parameters together:
175     \begin{equation}
176 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
177 gezelter 3524 \end{equation}
178     where
179 skuang 3565 \begin{eqnarray}
180 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
181 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
182 gezelter 3524 \label{eq:momentumdef}
183 skuang 3565 \end{eqnarray}
184 skuang 3528 Therefore, for each of the three directions, the hot scaling
185     parameters are a simple function of the cold scaling parameters and
186 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
187     \begin{equation}
188 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
189 gezelter 3524 \label{eq:hotcoldscaling}
190     \end{equation}
191 skuang 3528 where
192     \begin{equation}
193     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
194     \end{equation}
195     for convenience.
196 gezelter 3524
197     Conservation of total energy also places constraints on the scaling:
198     \begin{equation}
199     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
200 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
201 gezelter 3524 \end{equation}
202 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
203     $K_c^\alpha$, are computed for each of the three directions in a
204     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
205     Substituting in the expressions for the hot scaling parameters
206     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
207 gezelter 3583 {\it constraint ellipsoid}:
208 gezelter 3524 \begin{equation}
209 gezelter 3600 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
210     c_\alpha \right) = 0
211 gezelter 3524 \label{eq:constraintEllipsoid}
212     \end{equation}
213     where the constants are obtained from the instantaneous values of the
214     linear momenta and kinetic energies for the hot and cold slabs,
215 skuang 3565 \begin{eqnarray}
216 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
217     \left(p_\alpha\right)^2\right) \\
218     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
219 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
220 gezelter 3524 \label{eq:constraintEllipsoidConsts}
221 skuang 3565 \end{eqnarray}
222 gezelter 3583 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
223 gezelter 3600 cold slab scaling parameters which, when applied, preserve the linear
224     momentum of the system in all three directions as well as total
225     kinetic energy.
226 gezelter 3524
227 gezelter 3600 The goal of using these velocity scaling variables is to transfer
228 gezelter 3609 kinetic energy from the cold slab to the hot slab. If the hot and
229     cold slabs are separated along the z-axis, the energy flux is given
230     simply by the decrease in kinetic energy of the cold bin:
231 gezelter 3524 \begin{equation}
232 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
233 gezelter 3524 \end{equation}
234     The expression for the energy flux can be re-written as another
235     ellipsoid centered on $(x,y,z) = 0$:
236     \begin{equation}
237 skuang 3604 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
238     K_c^\alpha -J_z \Delta t
239 gezelter 3524 \label{eq:fluxEllipsoid}
240     \end{equation}
241 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
242 gezelter 3600 both by the target flux, $J_z$ as well as the instantaneous values of
243     the kinetic energy components in the cold bin.
244 gezelter 3524
245     To satisfy an energetic flux as well as the conservation constraints,
246 gezelter 3600 we must determine the points ${x,y,z}$ that lie on both the constraint
247     ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
248     (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
249     ellipsoids in 3-dimensional space.
250 gezelter 3524
251 gezelter 3569 \begin{figure}
252     \includegraphics[width=\linewidth]{ellipsoids}
253 gezelter 3600 \caption{Velocity scaling coefficients which maintain both constant
254     energy and constant linear momentum of the system lie on the surface
255     of the {\it constraint ellipsoid} while points which generate the
256     target momentum flux lie on the surface of the {\it flux ellipsoid}.
257     The velocity distributions in the cold bin are scaled by only those
258     points which lie on both ellipsoids.}
259 gezelter 3569 \label{ellipsoids}
260     \end{figure}
261    
262 gezelter 3600 Since ellipsoids can be expressed as polynomials up to second order in
263     each of the three coordinates, finding the the intersection points of
264     two ellipsoids is isomorphic to finding the roots a polynomial of
265     degree 16. There are a number of polynomial root-finding methods in
266 skuang 3614 the literature,\cite{Hoffman:2001sf,384119} but numerically finding
267     the roots of high-degree polynomials is generally an ill-conditioned
268 skuang 3613 problem.\cite{Hoffman:2001sf}[P157] One simplification is to maintain velocity
269 gezelter 3609 scalings that are {\it as isotropic as possible}. To do this, we
270     impose $x=y$, and to treat both the constraint and flux ellipsoids as
271     2-dimensional ellipses. In reduced dimensionality, the
272 gezelter 3600 intersecting-ellipse problem reduces to finding the roots of
273 gezelter 3609 polynomials of degree 4.
274 gezelter 3600
275     Depending on the target flux and current velocity distributions, the
276     ellipsoids can have between 0 and 4 intersection points. If there are
277     no intersection points, it is not possible to satisfy the constraints
278     while performing a non-equilibrium scaling move, and no change is made
279     to the dynamics.
280    
281     With multiple intersection points, any of the scaling points will
282     conserve the linear momentum and kinetic energy of the system and will
283     generate the correct target flux. Although this method is inherently
284     non-isotropic, the goal is still to maintain the system as close to an
285     isotropic fluid as possible. With this in mind, we would like the
286     kinetic energies in the three different directions could become as
287     close as each other as possible after each scaling. Simultaneously,
288     one would also like each scaling as gentle as possible, i.e. ${x,y,z
289     \rightarrow 1}$, in order to avoid large perturbation to the system.
290 gezelter 3609 To do this, we pick the intersection point which maintains the three
291     scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
292 gezelter 3600 ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
293    
294     After the valid scaling parameters are arrived at by solving geometric
295     intersection problems in $x, y, z$ space in order to obtain cold slab
296     scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
297     determine the conjugate hot slab scaling variables.
298    
299     \subsection{Introducing shear stress via velocity scaling}
300 gezelter 3609 It is also possible to use this method to magnify the random
301     fluctuations of the average momentum in each of the bins to induce a
302     momentum flux. Doing this repeatedly will create a shear stress on
303     the system which will respond with an easily-measured strain. The
304     momentum flux (say along the $x$-direction) may be defined as:
305 gezelter 3524 \begin{equation}
306 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
307 skuang 3531 \label{eq:fluxPlane}
308 gezelter 3524 \end{equation}
309 gezelter 3600 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
310     its position governed both by a target value, $j_z(p_x)$ as well as
311     the instantaneous value of the momentum along the $x$-direction.
312 gezelter 3524
313 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
314     constraints, we must determine the points ${x,y,z}$ which lie on both
315     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
316     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
317 gezelter 3600 ellipsoid and a plane in 3-dimensional space.
318 gezelter 3524
319 gezelter 3600 In the case of momentum flux transfer, we also impose another
320 gezelter 3609 constraint to set the kinetic energy transfer as zero. In other
321     words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
322 gezelter 3600 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
323     set of quartic equations to the above kinetic energy transfer problem.
324 gezelter 3524
325 gezelter 3600 \section{Computational Details}
326 gezelter 3583
327 gezelter 3609 We have implemented this methodology in our molecular dynamics code,
328     OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
329 skuang 3613 after an MD step with a variable frequency. We have tested the method
330     in a variety of different systems, including homogeneous fluids
331     (Lennard-Jones and SPC/E water), crystalline solids ({\sc
332 skuang 3615 eam})~\cite{PhysRevB.33.7983} and quantum Sutton-Chen ({\sc
333 skuang 3613 q-sc})~\cite{PhysRevB.59.3527} models for Gold), and heterogeneous
334 skuang 3615 interfaces ({\sc q-sc} gold - SPC/E water). The last of these systems would
335 skuang 3613 have been difficult to study using previous RNEMD methods, but using
336     velocity scaling moves, we can even obtain estimates of the
337     interfacial thermal conductivities ($G$).
338 gezelter 3524
339 gezelter 3609 \subsection{Simulation Cells}
340 gezelter 3524
341 gezelter 3609 In each of the systems studied, the dynamics was carried out in a
342     rectangular simulation cell using periodic boundary conditions in all
343     three dimensions. The cells were longer along the $z$ axis and the
344     space was divided into $N$ slabs along this axis (typically $N=20$).
345 skuang 3613 The top slab ($n=1$) was designated the ``hot'' slab, while the
346     central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
347 gezelter 3609 cases, simulations were first thermalized in canonical ensemble (NVT)
348     using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
349 gezelter 3600 microcanonical ensemble (NVE) before introducing any non-equilibrium
350     method.
351 skuang 3531
352 gezelter 3609 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
353 skuang 3531
354 gezelter 3609 In order to compare our new methodology with the original
355     M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
356     first performed simulations using the original technique.
357 skuang 3531
358 gezelter 3609 \subsection{RNEMD with NIVS scaling}
359    
360     For each simulation utilizing the swapping method, a corresponding
361     NIVS-RNEMD simulation was carried out using a target momentum flux set
362     to produce a the same momentum or energy flux exhibited in the
363     swapping simulation.
364    
365     To test the temperature homogeneity (and to compute transport
366     coefficients), directional momentum and temperature distributions were
367     accumulated for molecules in each of the slabs.
368    
369     \subsection{Shear viscosities}
370    
371     The momentum flux was calculated using the total non-physical momentum
372     transferred (${P_x}$) and the data collection time ($t$):
373 skuang 3534 \begin{equation}
374     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
375     \end{equation}
376 gezelter 3609 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
377     box. The factor of two in the denominator is present because physical
378     momentum transfer occurs in two directions due to our periodic
379     boundary conditions. The velocity gradient ${\langle \partial v_x
380     /\partial z \rangle}$ was obtained using linear regression of the
381     velocity profiles in the bins. For Lennard-Jones simulations, shear
382     viscosities are reporte in reduced units (${\eta^* = \eta \sigma^2
383     (\varepsilon m)^{-1/2}}$).
384 skuang 3532
385 gezelter 3609 \subsection{Thermal Conductivities}
386 skuang 3534
387 gezelter 3609 The energy flux was calculated similarly to the momentum flux, using
388     the total non-physical energy transferred (${E_{total}}$) and the data
389     collection time $t$:
390 skuang 3534 \begin{equation}
391     J_z = \frac{E_{total}}{2 t L_x L_y}
392     \end{equation}
393 gezelter 3609 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
394     obtained by a linear regression of the temperature profile. For
395     Lennard-Jones simulations, thermal conductivities are reported in
396     reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
397     k_B^{-1}\varepsilon^{-1/2}}$).
398 skuang 3534
399 gezelter 3609 \subsection{Interfacial Thermal Conductivities}
400 skuang 3563
401 gezelter 3609 For materials with a relatively low interfacial conductance, and in
402     cases where the flux between the materials is small, the bulk regions
403     on either side of an interface rapidly come to a state in which the
404     two phases have relatively homogeneous (but distinct) temperatures.
405     In calculating the interfacial thermal conductivity $G$, this
406     assumption was made, and the conductance can be approximated as:
407 skuang 3573
408     \begin{equation}
409     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
410     \langle T_{water}\rangle \right)}
411     \label{interfaceCalc}
412     \end{equation}
413 gezelter 3609 where ${E_{total}}$ is the imposed non-physical kinetic energy
414     transfer and ${\langle T_{gold}\rangle}$ and ${\langle
415     T_{water}\rangle}$ are the average observed temperature of gold and
416     water phases respectively.
417 skuang 3573
418 gezelter 3609 \section{Results}
419 skuang 3538
420 gezelter 3609 \subsection{Lennard-Jones Fluid}
421     2592 Lennard-Jones atoms were placed in an orthorhombic cell
422     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
423     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
424     direct comparison between our results and previous methods. These
425     simulations were carried out with a reduced timestep ${\tau^* =
426     4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
427     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
428     conductivity calculations, simulations were first run under reduced
429     temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
430     ensemble, but other temperatures ([XXX, YYY, and ZZZ]) were also
431     sampled. The simulations included $10^5$ steps of equilibration
432     without any momentum flux, $10^5$ steps of stablization with an
433     imposed momentum transfer to create a gradient, and $10^6$ steps of
434     data collection under RNEMD.
435    
436 gezelter 3611 \subsubsection*{Thermal Conductivity}
437    
438 gezelter 3609 Our thermal conductivity calculations show that the NIVS method agrees
439     well with the swapping method. Four different swap intervals were
440 skuang 3613 tested (Table \ref{LJ}). With a fixed scaling interval of 10 time steps,
441     the target exchange kinetic energy produced equivalent kinetic energy
442     flux as in the swapping method. Similar thermal gradients were
443     observed with similar thermal flux under the two different methods
444     (Figure \ref{thermalGrad}).
445 gezelter 3609
446 skuang 3563 \begin{table*}
447 gezelter 3609 \begin{minipage}{\linewidth}
448     \begin{center}
449 skuang 3538
450 gezelter 3612 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
451     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
452     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
453     at various momentum fluxes. The original swapping method and
454     the velocity scaling method give similar results.
455     Uncertainties are indicated in parentheses.}
456 gezelter 3609
457 gezelter 3612 \begin{tabular}{|cc|cc|cc|}
458 gezelter 3609 \hline
459 gezelter 3612 \multicolumn{2}{|c}{Momentum Exchange} &
460     \multicolumn{2}{|c}{Swapping RNEMD} &
461 gezelter 3609 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
462     \hline
463 gezelter 3612 \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
464    
465     \multirow{2}{*}{$\lambda^*_{swap}$} &
466     \multirow{2}{*}{$\eta^*_{swap}$} &
467     \multirow{2}{*}{$\lambda^*_{scale}$} &
468     \multirow{2}{*}{$\eta^*_{scale}$} \\
469     & $j_p^*(v_x)$ (reduced units) & & & & \\
470 gezelter 3609 \hline
471 gezelter 3612 250 & 0.16 & 7.03(0.34) & & 7.30(0.10) & \\
472     500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
473     1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
474     2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
475     2500 & 0.019 & & 3.42(0.06) & & 3.43(0.08)\\
476 gezelter 3609 \hline
477     \end{tabular}
478 gezelter 3612 \label{LJ}
479 gezelter 3609 \end{center}
480     \end{minipage}
481 skuang 3563 \end{table*}
482    
483     \begin{figure}
484 gezelter 3612 \includegraphics[width=\linewidth]{thermalGrad}
485     \caption{NIVS-RNEMD method creates similar temperature gradients
486     compared with the swapping method under a variety of imposed kinetic
487     energy flux values.}
488     \label{thermalGrad}
489 skuang 3563 \end{figure}
490    
491 gezelter 3612 \subsubsection*{Velocity Distributions}
492    
493 gezelter 3609 During these simulations, velocities were recorded every 1000 steps
494     and was used to produce distributions of both velocity and speed in
495     each of the slabs. From these distributions, we observed that under
496 skuang 3613 relatively high non-physical kinetic energy flux, the speed of
497 gezelter 3609 molecules in slabs where swapping occured could deviate from the
498     Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
499     and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
500     distributions deviate from an ideal distribution. In the ``hot'' slab,
501     the probability density is notched at low speeds and has a substantial
502     shoulder at higher speeds relative to the ideal MB distribution. In
503     the cold slab, the opposite notching and shouldering occurs. This
504     phenomenon is more obvious at higher swapping rates.
505 skuang 3563
506 gezelter 3609 In the velocity distributions, the ideal Gaussian peak is
507     substantially flattened in the hot slab, and is overly sharp (with
508     truncated wings) in the cold slab. This problem is rooted in the
509     mechanism of the swapping method. Continually depleting low (high)
510     speed particles in the high (low) temperature slab is not complemented
511     by diffusions of low (high) speed particles from neighboring slabs,
512     unless the swapping rate is sufficiently small. Simutaneously, surplus
513     low speed particles in the low temperature slab do not have sufficient
514     time to diffuse to neighboring slabs. Since the thermal exchange rate
515     must reach a minimum level to produce an observable thermal gradient,
516     the swapping-method RNEMD has a relatively narrow choice of exchange
517     times that can be utilized.
518 skuang 3578
519 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
520     Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
521     this is simple; upon velocity scaling, a Gaussian distribution remains
522     Gaussian. Although a single scaling move is non-isotropic in three
523     dimensions, our criteria for choosing a set of scaling coefficients
524     helps maintain the distributions as close to isotropic as possible.
525     Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
526     as the previous RNEMD methods but without large perturbations to the
527     velocity distributions in the two slabs.
528    
529 skuang 3568 \begin{figure}
530 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
531     \caption{Speed distribution for thermal conductivity using a)
532     ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
533     simulations with an exchange or equilvalent exchange interval of 250
534 skuang 3593 fs. In circled areas, distributions from ``swapping'' RNEMD
535     simulation have deviation from ideal Maxwell-Boltzmann distribution
536     (curves fit for each distribution).}
537 skuang 3589 \label{thermalHist}
538 skuang 3568 \end{figure}
539    
540 gezelter 3611
541     \subsubsection*{Shear Viscosity}
542 gezelter 3612 Our calculations (Table \ref{LJ}) show that velocity-scaling
543 gezelter 3611 RNEMD predicted comparable shear viscosities to swap RNEMD method. All
544     the scale method results were from simulations that had a scaling
545     interval of 10 time steps. The average molecular momentum gradients of
546     these samples are shown in Figure \ref{shear} (a) and (b).
547    
548     \begin{figure}
549     \includegraphics[width=\linewidth]{shear}
550     \caption{Average momentum gradients in shear viscosity simulations,
551     using (a) ``swapping'' method and (b) NIVS-RNEMD method
552     respectively. (c) Temperature difference among x and y, z dimensions
553     observed when using NIVS-RNEMD with equivalent exchange interval of
554     500 fs.}
555     \label{shear}
556     \end{figure}
557    
558     However, observations of temperatures along three dimensions show that
559     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
560     two slabs which were scaled. Figure \ref{shear} (c) indicate that with
561     relatively large imposed momentum flux, the temperature difference among $x$
562     and the other two dimensions was significant. This would result from the
563     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
564     momentum gradient is set up, $P_c^x$ would be roughly stable
565     ($<0$). Consequently, scaling factor $x$ would most probably larger
566     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
567     keep increase after most scaling steps. And if there is not enough time
568     for the kinetic energy to exchange among different dimensions and
569     different slabs, the system would finally build up temperature
570     (kinetic energy) difference among the three dimensions. Also, between
571     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
572     are closer to neighbor slabs. This is due to momentum transfer along
573     $z$ dimension between slabs.
574    
575     Although results between scaling and swapping methods are comparable,
576     the inherent temperature inhomogeneity even in relatively low imposed
577     exchange momentum flux simulations makes scaling RNEMD method less
578     attractive than swapping RNEMD in shear viscosity calculation.
579    
580    
581 gezelter 3609 \subsection{Bulk SPC/E water}
582    
583     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
584     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
585     the original swapping RNEMD method. Bedrov {\it et
586 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
587 skuang 3579 center-of-mass velocities instead of single atom velocities in a
588 gezelter 3609 molecule conserves the total kinetic energy and linear momentum. This
589     principle is also adopted in our simulations. Scaling was applied to
590     the center-of-mass velocities of the rigid bodies of SPC/E model water
591     molecules.
592 skuang 3563
593 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
594     molecules were first equilibrated under ambient pressure and
595     temperature conditions using the isobaric-isothermal (NPT)
596     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
597     average volume observed in the NPT simulations, and this was followed
598     by equilibration, first in the canonical (NVT) ensemble, followed by a
599 skuang 3613 100ps period under constant-NVE conditions without any momentum
600     flux. 100ps was allowed to stabilize the system with an imposed
601     momentum transfer to create a gradient, and 1ns was alotted for
602 gezelter 3609 data collection under RNEMD.
603    
604     As shown in Figure \ref{spceGrad}, temperature gradients were
605 skuang 3615 established similar to the previous work. Our simulation results under
606     318K are in well agreement with those from Bedrov {\it et al.} (Table
607     \ref{spceThermal}). And both methods yield values in reasonable
608     agreement with experimental value. A larger difference between
609     simulation result and experiment is found under 300K. This could
610     result from the force field that is used in simulation.
611 gezelter 3609
612 skuang 3570 \begin{figure}
613 gezelter 3609 \includegraphics[width=\linewidth]{spceGrad}
614     \caption{Temperature gradients in SPC/E water thermal conductivity
615     simulations.}
616     \label{spceGrad}
617 skuang 3570 \end{figure}
618    
619     \begin{table*}
620 gezelter 3609 \begin{minipage}{\linewidth}
621     \begin{center}
622    
623     \caption{Thermal conductivity of SPC/E water under various
624     imposed thermal gradients. Uncertainties are indicated in
625     parentheses.}
626    
627 skuang 3615 \begin{tabular}{|c|c|ccc|}
628 gezelter 3609 \hline
629 skuang 3615 \multirow{2}{*}{$\langle T\rangle$(K)} &
630     \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
631     \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
632     \mathrm{K}^{-1})$} \\
633     & & This work & Previous simulations\cite{Bedrov:2000} &
634 gezelter 3609 Experiment\cite{WagnerKruse}\\
635     \hline
636 skuang 3615 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
637     \multirow{3}{*}{0.61}\\
638     & 0.81 & 0.770(0.008) & & \\
639     & 1.54 & 0.813(0.007) & & \\
640 gezelter 3609 \hline
641 skuang 3615 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
642     \multirow{2}{*}{0.64}\\
643     & 1.59 & 0.778(0.019) & 0.730 & \\
644     \hline
645 gezelter 3609 \end{tabular}
646     \label{spceThermal}
647     \end{center}
648     \end{minipage}
649     \end{table*}
650 skuang 3570
651 gezelter 3609 \subsection{Crystalline Gold}
652 skuang 3570
653 gezelter 3609 To see how the method performed in a solid, we calculated thermal
654     conductivities using two atomistic models for gold. Several different
655     potential models have been developed that reasonably describe
656     interactions in transition metals. In particular, the Embedded Atom
657     Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc
658     sc})~\cite{Chen90} potential have been used to study a wide range of
659     phenomena in both bulk materials and
660     nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
661     Both potentials are based on a model of a metal which treats the
662     nuclei and core electrons as pseudo-atoms embedded in the electron
663     density due to the valence electrons on all of the other atoms in the
664     system. The {\sc sc} potential has a simple form that closely
665     resembles the Lennard Jones potential,
666     \begin{equation}
667     \label{eq:SCP1}
668     U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
669     \end{equation}
670     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
671     \begin{equation}
672     \label{eq:SCP2}
673     V^{pair}_{ij}(r)=\left( \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}.
674     \end{equation}
675     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
676     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
677     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
678     the interactions between the valence electrons and the cores of the
679     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
680     scale, $c_i$ scales the attractive portion of the potential relative
681     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
682     that assures a dimensionless form for $\rho$. These parameters are
683     tuned to various experimental properties such as the density, cohesive
684     energy, and elastic moduli for FCC transition metals. The quantum
685     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
686     including zero-point quantum corrections for different transition
687     metals.\cite{PhysRevB.59.3527} The {\sc eam} functional forms differ
688     slightly from {\sc sc} but the overall method is very similar.
689 skuang 3570
690 gezelter 3609 In this work, we have utilized both the {\sc eam} and the {\sc q-sc}
691     potentials to test the behavior of scaling RNEMD.
692 skuang 3570
693 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
694 skuang 3613 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
695     both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
696 gezelter 3609 pre-equilibration at a target pressure of 1 atm. When equilibrated
697     under NPT conditions, our simulation box expanded by approximately 1\%
698 skuang 3613 in volume. Following adjustment of the box volume, equilibrations in
699     both the canonical and microcanonical ensembles were carried out. With
700     the simulation cell divided evenly into 10 slabs, different thermal
701     gradients were established by applying a set of target thermal
702     transfer fluxes.
703 skuang 3570
704 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
705     \ref{AuThermal}. In these calculations, the end and middle slabs were
706 gezelter 3610 excluded in thermal gradient linear regession. {\sc eam} predicts
707 gezelter 3609 slightly larger thermal conductivities than {\sc q-sc}. However, both
708     values are smaller than experimental value by a factor of more than
709     200. This behavior has been observed previously by Richardson and
710 skuang 3615 Clancy, and has been attributed to the lack of electronic contribution
711     in these force fields.\cite{Clancy:1992} The non-equilibrium MD method
712 skuang 3613 employed in their simulations gave an thermal conductance estimation
713 skuang 3615 of {\sc eam} gold as 1.74$\mathrm{W m}^{-1}\mathrm{K}^{-1}$. As stated
714     in their work, this was a rough estimation in the temperature range
715     300K-800K. Therefore, our results could be more accurate. It should be
716     noted that the density of the metal being simulated also affects the
717     thermal conductivity significantly. With an expanded lattice, lower
718     thermal conductance is expected (and observed). We also observed a
719     decrease in thermal conductance at higher temperatures, a trend that
720     agrees with experimental measurements [PAGE NUMBERS?].\cite{AshcroftMermin}
721 skuang 3570
722 gezelter 3609 \begin{table*}
723     \begin{minipage}{\linewidth}
724     \begin{center}
725    
726     \caption{Calculated thermal conductivity of crystalline gold
727     using two related force fields. Calculations were done at both
728     experimental and equilibrated densities and at a range of
729     temperatures and thermal flux rates. Uncertainties are
730 skuang 3613 indicated in parentheses. [SWAPPING COMPARISON?]}
731 gezelter 3609
732     \begin{tabular}{|c|c|c|cc|}
733     \hline
734     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
735     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
736     \hline
737     \multirow{7}{*}{\sc q-sc} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
738     & & & 2.86 & 1.08(0.05)\\
739     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
740     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
741     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
742     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
743     & & & 4.84 & 0.92(0.05)\\
744     \hline
745     \multirow{8}{*}{\sc eam} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
746     & & & 2.06 & 1.37(0.04)\\
747     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
748     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
749     & & & 2.04 & 1.41(0.07)\\
750     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
751     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
752     & & & 4.14 & 1.08(0.05)\\
753     \hline
754     \end{tabular}
755     \label{AuThermal}
756     \end{center}
757     \end{minipage}
758 skuang 3580 \end{table*}
759    
760 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
761     The most attractive aspect of the scaling approach for RNEMD is the
762     ability to use the method in non-homogeneous systems, where molecules
763     of different identities are segregated in different slabs. To test
764     this application, we simulated a Gold (111) / water interface. To
765     construct the interface, a box containing a lattice of 1188 Au atoms
766     (with the 111 surface in the +z and -z directions) was allowed to
767     relax under ambient temperature and pressure. A separate (but
768     identically sized) box of SPC/E water was also equilibrated at ambient
769     conditions. The two boxes were combined by removing all water
770 skuang 3613 molecules within 3 \AA radius of any gold atom. The final
771 gezelter 3609 configuration contained 1862 SPC/E water molecules.
772 skuang 3580
773 gezelter 3609 After simulations of bulk water and crystal gold, a mixture system was
774     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
775     molecules. Spohr potential was adopted in depicting the interaction
776     between metal atom and water molecule.\cite{ISI:000167766600035} A
777     similar protocol of equilibration was followed. Several thermal
778     gradients was built under different target thermal flux. It was found
779     out that compared to our previous simulation systems, the two phases
780     could have large temperature difference even under a relatively low
781     thermal flux.
782    
783    
784 skuang 3581 After simulations of homogeneous water and gold systems using
785     NIVS-RNEMD method were proved valid, calculation of gold/water
786     interfacial thermal conductivity was followed. It is found out that
787     the low interfacial conductance is probably due to the hydrophobic
788 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
789 skuang 3581 density change along $z$-axis, which is perpendicular to the
790     gold/water interface. It is observed that water density significantly
791     decreases when approaching the surface. Under this low thermal
792     conductance, both gold and water phase have sufficient time to
793     eliminate temperature difference inside respectively (Figure
794 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
795 skuang 3581 within respective phase, it is valid to assume that the temperature
796     difference between gold and water on surface would be approximately
797     the same as the difference between the gold and water phase. This
798     assumption enables convenient calculation of $G$ using
799     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
800     layer of water and gold close enough to surface, which would have
801     greater fluctuation and lower accuracy. Reported results (Table
802     \ref{interfaceRes}) are of two orders of magnitude smaller than our
803     calculations on homogeneous systems, and thus have larger relative
804     errors than our calculation results on homogeneous systems.
805 skuang 3573
806 skuang 3571 \begin{figure}
807 skuang 3595 \includegraphics[width=\linewidth]{interface}
808     \caption{Simulation results for Gold/Water interfacial thermal
809     conductivity: (a) Significant water density decrease is observed on
810 skuang 3597 crystalline gold surface, which indicates low surface contact and
811     leads to low thermal conductance. (b) Temperature profiles for a
812     series of simulations. Temperatures of different slabs in the same
813     phase show no significant differences.}
814 skuang 3595 \label{interface}
815 skuang 3571 \end{figure}
816    
817 skuang 3572 \begin{table*}
818 gezelter 3612 \begin{minipage}{\linewidth}
819     \begin{center}
820    
821     \caption{Computed interfacial thermal conductivity ($G$) values
822     for the Au(111) / water interface at ${\langle T\rangle \sim}$
823     300K using a range of energy fluxes. Uncertainties are
824     indicated in parentheses. }
825    
826     \begin{tabular}{|cccc|}
827     \hline
828     $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
829     T_{water} \rangle$ (K) & $G$
830     (MW/m$^2$/K)\\
831     \hline
832     98.0 & 355.2 & 295.8 & 1.65(0.21) \\
833     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
834     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
835     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
836     \hline
837     \end{tabular}
838     \label{interfaceRes}
839     \end{center}
840     \end{minipage}
841 skuang 3572 \end{table*}
842    
843 skuang 3576
844 skuang 3574 \section{Conclusions}
845     NIVS-RNEMD simulation method is developed and tested on various
846 skuang 3581 systems. Simulation results demonstrate its validity in thermal
847     conductivity calculations, from Lennard-Jones fluid to multi-atom
848     molecule like water and metal crystals. NIVS-RNEMD improves
849     non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
850     methods. Furthermore, it develops a valid means for unphysical thermal
851     transfer between different species of molecules, and thus extends its
852     applicability to interfacial systems. Our calculation of gold/water
853     interfacial thermal conductivity demonstrates this advantage over
854     previous RNEMD methods. NIVS-RNEMD has also limited application on
855     shear viscosity calculations, but could cause temperature difference
856     among different dimensions under high momentum flux. Modification is
857     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
858     calculations.
859 skuang 3572
860 gezelter 3524 \section{Acknowledgments}
861     Support for this project was provided by the National Science
862     Foundation under grant CHE-0848243. Computational time was provided by
863     the Center for Research Computing (CRC) at the University of Notre
864     Dame. \newpage
865    
866 gezelter 3583 \bibliographystyle{aip}
867 gezelter 3524 \bibliography{nivsRnemd}
868 gezelter 3583
869 gezelter 3524 \end{doublespace}
870     \end{document}
871