ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3619
Committed: Fri Jul 30 21:13:43 2010 UTC (14 years ago) by skuang
Content type: application/x-tex
File size: 41610 byte(s)
Log Message:
changes in two LJ plots. and other small 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 gezelter 3616 %\usepackage[ref]{overcite}
14     \usepackage[square, comma, sort&compress]{natbib}
15     \usepackage{url}
16 gezelter 3524 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18     9.0in \textwidth 6.5in \brokenpenalty=10000
19    
20     % double space list of tables and figures
21     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22     \setlength{\abovecaptionskip}{20 pt}
23     \setlength{\belowcaptionskip}{30 pt}
24    
25 gezelter 3616 %\renewcommand\citemid{\ } % no comma in optional referenc note
26     \bibpunct{[}{]}{,}{s}{}{;}
27     \bibliographystyle{aip}
28 gezelter 3524
29     \begin{document}
30    
31     \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
32    
33     \author{Shenyu Kuang and J. Daniel
34     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
35     Department of Chemistry and Biochemistry,\\
36     University of Notre Dame\\
37     Notre Dame, Indiana 46556}
38    
39     \date{\today}
40    
41     \maketitle
42    
43     \begin{doublespace}
44    
45     \begin{abstract}
46 gezelter 3583 We present a new method for introducing stable non-equilibrium
47     velocity and temperature distributions in molecular dynamics
48 gezelter 3609 simulations of heterogeneous systems. This method extends earlier
49     Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
50     momentum exchange swapping moves that can create non-thermal
51     velocity distributions and are difficult to use for interfacial
52     calculations. By using non-isotropic velocity scaling (NIVS) on the
53     molecules in specific regions of a system, it is possible to impose
54     momentum or thermal flux between regions of a simulation and stable
55     thermal and momentum gradients can then be established. The scaling
56     method we have developed conserves the total linear momentum and
57     total energy of the system. To test the methods, we have computed
58     the thermal conductivity of model liquid and solid systems as well
59     as the interfacial thermal conductivity of a metal-water interface.
60     We find that the NIVS-RNEMD improves the problematic velocity
61     distributions that develop in other RNEMD methods.
62 gezelter 3524 \end{abstract}
63    
64     \newpage
65    
66     %\narrowtext
67    
68     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69     % BODY OF TEXT
70     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71    
72     \section{Introduction}
73     The original formulation of Reverse Non-equilibrium Molecular Dynamics
74     (RNEMD) obtains transport coefficients (thermal conductivity and shear
75     viscosity) in a fluid by imposing an artificial momentum flux between
76     two thin parallel slabs of material that are spatially separated in
77 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
78 gezelter 3583 artificial flux is typically created by periodically ``swapping''
79     either the entire momentum vector $\vec{p}$ or single components of
80     this vector ($p_x$) between molecules in each of the two slabs. If
81     the two slabs are separated along the $z$ coordinate, the imposed flux
82     is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
83     response of a simulated system to the imposed momentum flux will
84     typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
85     The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
86     easily obtained by assuming linear response of the system,
87 gezelter 3524 \begin{eqnarray}
88 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
89 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
90 gezelter 3524 \end{eqnarray}
91 gezelter 3600 RNEMD has been widely used to provide computational estimates of
92     thermal conductivities and shear viscosities in a wide range of
93     materials, from liquid copper to both monatomic and molecular fluids
94     (e.g. ionic
95     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}
96 gezelter 3524
97 skuang 3574 \begin{figure}
98     \includegraphics[width=\linewidth]{thermalDemo}
99 gezelter 3583 \caption{RNEMD methods impose an unphysical transfer of momentum or
100     kinetic energy between a ``hot'' slab and a ``cold'' slab in the
101     simulation box. The molecular system responds to this imposed flux
102     by generating a momentum or temperature gradient. The slope of the
103     gradient can then be used to compute transport properties (e.g.
104     shear viscosity and thermal conductivity).}
105 skuang 3574 \label{thermalDemo}
106     \end{figure}
107    
108 skuang 3591 RNEMD is preferable in many ways to the forward NEMD
109 skuang 3592 methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
110     because it imposes what is typically difficult to measure (a flux or
111 gezelter 3600 stress) and it is typically much easier to compute the response
112 gezelter 3609 (momentum gradients or strains). For similar reasons, RNEMD is also
113 skuang 3592 preferable to slowly-converging equilibrium methods for measuring
114     thermal conductivity and shear viscosity (using Green-Kubo
115 skuang 3591 relations\cite{daivis:541,mondello:9327} or the Helfand moment
116     approach of Viscardy {\it et
117 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
118 gezelter 3524 computing difficult to measure quantities.
119    
120     Another attractive feature of RNEMD is that it conserves both total
121     linear momentum and total energy during the swaps (as long as the two
122     molecules have the same identity), so the swapped configurations are
123     typically samples from the same manifold of states in the
124     microcanonical ensemble.
125    
126 skuang 3588 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
127 skuang 3565 some problems with the original RNEMD swap technique. Notably, large
128     momentum fluxes (equivalent to frequent momentum swaps between the
129 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
130     momentum distributions in the two slabs, as well as non-linear thermal
131     and velocity distributions along the direction of the imposed flux
132     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
133     and self-adjusting metrics for retaining the usability of the method.
134 gezelter 3524
135     In this paper, we develop and test a method for non-isotropic velocity
136 gezelter 3600 scaling (NIVS) which retains the desirable features of RNEMD
137 gezelter 3524 (conservation of linear momentum and total energy, compatibility with
138     periodic boundary conditions) while establishing true thermal
139 gezelter 3600 distributions in each of the two slabs. In the next section, we
140 gezelter 3583 present the method for determining the scaling constraints. We then
141 gezelter 3600 test the method on both liquids and solids as well as a non-isotropic
142     liquid-solid interface and show that it is capable of providing
143 gezelter 3524 reasonable estimates of the thermal conductivity and shear viscosity
144 gezelter 3600 in all of these cases.
145 gezelter 3524
146     \section{Methodology}
147 gezelter 3583 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
148     periodic system is partitioned into a series of thin slabs along one
149 gezelter 3524 axis ($z$). One of the slabs at the end of the periodic box is
150     designated the ``hot'' slab, while the slab in the center of the box
151     is designated the ``cold'' slab. The artificial momentum flux will be
152     established by transferring momentum from the cold slab and into the
153     hot slab.
154    
155     Rather than using momentum swaps, we use a series of velocity scaling
156 gezelter 3583 moves. For molecules $\{i\}$ located within the cold slab,
157 gezelter 3524 \begin{equation}
158 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
159     x & 0 & 0 \\
160     0 & y & 0 \\
161     0 & 0 & z \\
162 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
163     \end{equation}
164 gezelter 3600 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
165     of the three directions in the system. Likewise, the molecules
166     $\{j\}$ located in the hot slab will see a concomitant scaling of
167     velocities,
168 gezelter 3524 \begin{equation}
169 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
170     x^\prime & 0 & 0 \\
171     0 & y^\prime & 0 \\
172     0 & 0 & z^\prime \\
173 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
174     \end{equation}
175    
176     Conservation of linear momentum in each of the three directions
177 gezelter 3583 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
178 gezelter 3524 parameters together:
179     \begin{equation}
180 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
181 gezelter 3524 \end{equation}
182     where
183 skuang 3565 \begin{eqnarray}
184 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
185 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
186 gezelter 3524 \label{eq:momentumdef}
187 skuang 3565 \end{eqnarray}
188 skuang 3528 Therefore, for each of the three directions, the hot scaling
189     parameters are a simple function of the cold scaling parameters and
190 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
191     \begin{equation}
192 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
193 gezelter 3524 \label{eq:hotcoldscaling}
194     \end{equation}
195 skuang 3528 where
196     \begin{equation}
197     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
198     \end{equation}
199     for convenience.
200 gezelter 3524
201     Conservation of total energy also places constraints on the scaling:
202     \begin{equation}
203     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
204 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
205 gezelter 3524 \end{equation}
206 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
207     $K_c^\alpha$, are computed for each of the three directions in a
208     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
209     Substituting in the expressions for the hot scaling parameters
210     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
211 gezelter 3583 {\it constraint ellipsoid}:
212 gezelter 3524 \begin{equation}
213 gezelter 3600 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
214     c_\alpha \right) = 0
215 gezelter 3524 \label{eq:constraintEllipsoid}
216     \end{equation}
217     where the constants are obtained from the instantaneous values of the
218     linear momenta and kinetic energies for the hot and cold slabs,
219 skuang 3565 \begin{eqnarray}
220 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
221     \left(p_\alpha\right)^2\right) \\
222     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
223 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
224 gezelter 3524 \label{eq:constraintEllipsoidConsts}
225 skuang 3565 \end{eqnarray}
226 gezelter 3583 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
227 gezelter 3600 cold slab scaling parameters which, when applied, preserve the linear
228     momentum of the system in all three directions as well as total
229     kinetic energy.
230 gezelter 3524
231 gezelter 3600 The goal of using these velocity scaling variables is to transfer
232 gezelter 3609 kinetic energy from the cold slab to the hot slab. If the hot and
233     cold slabs are separated along the z-axis, the energy flux is given
234     simply by the decrease in kinetic energy of the cold bin:
235 gezelter 3524 \begin{equation}
236 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
237 gezelter 3524 \end{equation}
238     The expression for the energy flux can be re-written as another
239     ellipsoid centered on $(x,y,z) = 0$:
240     \begin{equation}
241 skuang 3604 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
242     K_c^\alpha -J_z \Delta t
243 gezelter 3524 \label{eq:fluxEllipsoid}
244     \end{equation}
245 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
246 gezelter 3600 both by the target flux, $J_z$ as well as the instantaneous values of
247     the kinetic energy components in the cold bin.
248 gezelter 3524
249     To satisfy an energetic flux as well as the conservation constraints,
250 gezelter 3600 we must determine the points ${x,y,z}$ that lie on both the constraint
251     ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
252     (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
253     ellipsoids in 3-dimensional space.
254 gezelter 3524
255 gezelter 3569 \begin{figure}
256     \includegraphics[width=\linewidth]{ellipsoids}
257 gezelter 3600 \caption{Velocity scaling coefficients which maintain both constant
258     energy and constant linear momentum of the system lie on the surface
259     of the {\it constraint ellipsoid} while points which generate the
260     target momentum flux lie on the surface of the {\it flux ellipsoid}.
261     The velocity distributions in the cold bin are scaled by only those
262     points which lie on both ellipsoids.}
263 gezelter 3569 \label{ellipsoids}
264     \end{figure}
265    
266 gezelter 3600 Since ellipsoids can be expressed as polynomials up to second order in
267     each of the three coordinates, finding the the intersection points of
268     two ellipsoids is isomorphic to finding the roots a polynomial of
269     degree 16. There are a number of polynomial root-finding methods in
270 skuang 3614 the literature,\cite{Hoffman:2001sf,384119} but numerically finding
271     the roots of high-degree polynomials is generally an ill-conditioned
272 skuang 3617 problem.\cite{Hoffman:2001sf} One simplification is to maintain velocity
273 gezelter 3609 scalings that are {\it as isotropic as possible}. To do this, we
274     impose $x=y$, and to treat both the constraint and flux ellipsoids as
275     2-dimensional ellipses. In reduced dimensionality, the
276 gezelter 3600 intersecting-ellipse problem reduces to finding the roots of
277 gezelter 3609 polynomials of degree 4.
278 gezelter 3600
279     Depending on the target flux and current velocity distributions, the
280     ellipsoids can have between 0 and 4 intersection points. If there are
281     no intersection points, it is not possible to satisfy the constraints
282     while performing a non-equilibrium scaling move, and no change is made
283     to the dynamics.
284    
285     With multiple intersection points, any of the scaling points will
286     conserve the linear momentum and kinetic energy of the system and will
287     generate the correct target flux. Although this method is inherently
288     non-isotropic, the goal is still to maintain the system as close to an
289     isotropic fluid as possible. With this in mind, we would like the
290     kinetic energies in the three different directions could become as
291     close as each other as possible after each scaling. Simultaneously,
292     one would also like each scaling as gentle as possible, i.e. ${x,y,z
293     \rightarrow 1}$, in order to avoid large perturbation to the system.
294 gezelter 3609 To do this, we pick the intersection point which maintains the three
295     scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
296 gezelter 3600 ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
297    
298     After the valid scaling parameters are arrived at by solving geometric
299     intersection problems in $x, y, z$ space in order to obtain cold slab
300     scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
301     determine the conjugate hot slab scaling variables.
302    
303     \subsection{Introducing shear stress via velocity scaling}
304 gezelter 3609 It is also possible to use this method to magnify the random
305     fluctuations of the average momentum in each of the bins to induce a
306     momentum flux. Doing this repeatedly will create a shear stress on
307     the system which will respond with an easily-measured strain. The
308     momentum flux (say along the $x$-direction) may be defined as:
309 gezelter 3524 \begin{equation}
310 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
311 skuang 3531 \label{eq:fluxPlane}
312 gezelter 3524 \end{equation}
313 gezelter 3600 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
314     its position governed both by a target value, $j_z(p_x)$ as well as
315     the instantaneous value of the momentum along the $x$-direction.
316 gezelter 3524
317 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
318     constraints, we must determine the points ${x,y,z}$ which lie on both
319     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
320     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
321 gezelter 3600 ellipsoid and a plane in 3-dimensional space.
322 gezelter 3524
323 gezelter 3600 In the case of momentum flux transfer, we also impose another
324 gezelter 3609 constraint to set the kinetic energy transfer as zero. In other
325     words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
326 gezelter 3600 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
327     set of quartic equations to the above kinetic energy transfer problem.
328 gezelter 3524
329 gezelter 3600 \section{Computational Details}
330 gezelter 3583
331 gezelter 3609 We have implemented this methodology in our molecular dynamics code,
332     OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
333 skuang 3613 after an MD step with a variable frequency. We have tested the method
334     in a variety of different systems, including homogeneous fluids
335     (Lennard-Jones and SPC/E water), crystalline solids ({\sc
336 skuang 3615 eam})~\cite{PhysRevB.33.7983} and quantum Sutton-Chen ({\sc
337 skuang 3613 q-sc})~\cite{PhysRevB.59.3527} models for Gold), and heterogeneous
338 skuang 3615 interfaces ({\sc q-sc} gold - SPC/E water). The last of these systems would
339 skuang 3613 have been difficult to study using previous RNEMD methods, but using
340     velocity scaling moves, we can even obtain estimates of the
341     interfacial thermal conductivities ($G$).
342 gezelter 3524
343 gezelter 3609 \subsection{Simulation Cells}
344 gezelter 3524
345 gezelter 3609 In each of the systems studied, the dynamics was carried out in a
346     rectangular simulation cell using periodic boundary conditions in all
347     three dimensions. The cells were longer along the $z$ axis and the
348     space was divided into $N$ slabs along this axis (typically $N=20$).
349 skuang 3613 The top slab ($n=1$) was designated the ``hot'' slab, while the
350     central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
351 gezelter 3609 cases, simulations were first thermalized in canonical ensemble (NVT)
352     using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
353 gezelter 3600 microcanonical ensemble (NVE) before introducing any non-equilibrium
354     method.
355 skuang 3531
356 gezelter 3609 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
357 skuang 3531
358 gezelter 3609 In order to compare our new methodology with the original
359     M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
360     first performed simulations using the original technique.
361 skuang 3531
362 gezelter 3609 \subsection{RNEMD with NIVS scaling}
363    
364     For each simulation utilizing the swapping method, a corresponding
365     NIVS-RNEMD simulation was carried out using a target momentum flux set
366     to produce a the same momentum or energy flux exhibited in the
367     swapping simulation.
368    
369     To test the temperature homogeneity (and to compute transport
370     coefficients), directional momentum and temperature distributions were
371     accumulated for molecules in each of the slabs.
372    
373     \subsection{Shear viscosities}
374    
375     The momentum flux was calculated using the total non-physical momentum
376     transferred (${P_x}$) and the data collection time ($t$):
377 skuang 3534 \begin{equation}
378     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
379     \end{equation}
380 gezelter 3609 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
381     box. The factor of two in the denominator is present because physical
382     momentum transfer occurs in two directions due to our periodic
383     boundary conditions. The velocity gradient ${\langle \partial v_x
384     /\partial z \rangle}$ was obtained using linear regression of the
385     velocity profiles in the bins. For Lennard-Jones simulations, shear
386     viscosities are reporte in reduced units (${\eta^* = \eta \sigma^2
387     (\varepsilon m)^{-1/2}}$).
388 skuang 3532
389 gezelter 3609 \subsection{Thermal Conductivities}
390 skuang 3534
391 gezelter 3609 The energy flux was calculated similarly to the momentum flux, using
392     the total non-physical energy transferred (${E_{total}}$) and the data
393     collection time $t$:
394 skuang 3534 \begin{equation}
395     J_z = \frac{E_{total}}{2 t L_x L_y}
396     \end{equation}
397 gezelter 3609 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
398     obtained by a linear regression of the temperature profile. For
399     Lennard-Jones simulations, thermal conductivities are reported in
400     reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
401     k_B^{-1}\varepsilon^{-1/2}}$).
402 skuang 3534
403 gezelter 3609 \subsection{Interfacial Thermal Conductivities}
404 skuang 3563
405 gezelter 3609 For materials with a relatively low interfacial conductance, and in
406     cases where the flux between the materials is small, the bulk regions
407     on either side of an interface rapidly come to a state in which the
408     two phases have relatively homogeneous (but distinct) temperatures.
409     In calculating the interfacial thermal conductivity $G$, this
410     assumption was made, and the conductance can be approximated as:
411 skuang 3573
412     \begin{equation}
413     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
414     \langle T_{water}\rangle \right)}
415     \label{interfaceCalc}
416     \end{equation}
417 gezelter 3609 where ${E_{total}}$ is the imposed non-physical kinetic energy
418     transfer and ${\langle T_{gold}\rangle}$ and ${\langle
419     T_{water}\rangle}$ are the average observed temperature of gold and
420     water phases respectively.
421 skuang 3573
422 gezelter 3609 \section{Results}
423 skuang 3538
424 gezelter 3609 \subsection{Lennard-Jones Fluid}
425     2592 Lennard-Jones atoms were placed in an orthorhombic cell
426     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
427     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
428     direct comparison between our results and previous methods. These
429     simulations were carried out with a reduced timestep ${\tau^* =
430     4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
431     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
432 skuang 3617 conductivity calculations, simulations were run under reduced
433 gezelter 3609 temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
434 skuang 3617 ensemble. The simulations included $10^5$ steps of equilibration
435 gezelter 3609 without any momentum flux, $10^5$ steps of stablization with an
436     imposed momentum transfer to create a gradient, and $10^6$ steps of
437     data collection under RNEMD.
438    
439 gezelter 3611 \subsubsection*{Thermal Conductivity}
440    
441 gezelter 3609 Our thermal conductivity calculations show that the NIVS method agrees
442 skuang 3618 well with the swapping method. Five different swap intervals were
443 skuang 3613 tested (Table \ref{LJ}). With a fixed scaling interval of 10 time steps,
444     the target exchange kinetic energy produced equivalent kinetic energy
445     flux as in the swapping method. Similar thermal gradients were
446     observed with similar thermal flux under the two different methods
447 skuang 3618 (Figure \ref{thermalGrad}). Furthermore, with appropriate choice of
448     scaling variables, temperature along $x$, $y$ and $z$ axis has no
449     observable difference(Figure TO BE ADDED). The system is able
450     to maintain temperature homogeneity even under high flux.
451 gezelter 3609
452 skuang 3563 \begin{table*}
453 gezelter 3609 \begin{minipage}{\linewidth}
454     \begin{center}
455 skuang 3538
456 gezelter 3612 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
457     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
458     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
459     at various momentum fluxes. The original swapping method and
460     the velocity scaling method give similar results.
461     Uncertainties are indicated in parentheses.}
462 gezelter 3609
463 gezelter 3612 \begin{tabular}{|cc|cc|cc|}
464 gezelter 3609 \hline
465 gezelter 3612 \multicolumn{2}{|c}{Momentum Exchange} &
466     \multicolumn{2}{|c}{Swapping RNEMD} &
467 gezelter 3609 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
468     \hline
469 gezelter 3612 \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
470     \multirow{2}{*}{$\lambda^*_{swap}$} &
471     \multirow{2}{*}{$\eta^*_{swap}$} &
472     \multirow{2}{*}{$\lambda^*_{scale}$} &
473     \multirow{2}{*}{$\eta^*_{scale}$} \\
474 skuang 3617 & $j_z^*(p_x)$ (reduced units) & & & & \\
475 gezelter 3609 \hline
476 skuang 3617 250 & 0.16 & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
477 gezelter 3612 500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
478     1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
479     2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
480 skuang 3617 2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
481 gezelter 3609 \hline
482     \end{tabular}
483 gezelter 3612 \label{LJ}
484 gezelter 3609 \end{center}
485     \end{minipage}
486 skuang 3563 \end{table*}
487    
488     \begin{figure}
489 gezelter 3612 \includegraphics[width=\linewidth]{thermalGrad}
490 skuang 3619 \caption{NIVS-RNEMD method (b) creates similar temperature gradients
491     compared with the swapping method (a) under a variety of imposed
492     kinetic energy flux values. Furthermore, the implementation of
493     Non-Isotropic Velocity Scaling does not cause temperature
494     difference among the three dimensions (c).}
495 gezelter 3612 \label{thermalGrad}
496 skuang 3563 \end{figure}
497    
498 gezelter 3612 \subsubsection*{Velocity Distributions}
499    
500 gezelter 3609 During these simulations, velocities were recorded every 1000 steps
501     and was used to produce distributions of both velocity and speed in
502     each of the slabs. From these distributions, we observed that under
503 skuang 3613 relatively high non-physical kinetic energy flux, the speed of
504 gezelter 3609 molecules in slabs where swapping occured could deviate from the
505     Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
506 skuang 3619 and Maginn\cite{Maginn:2010} in their simulations for shear viscosity
507     calculations. Figure \ref{thermalHist} shows how these distributions
508     deviate from an ideal distribution. In the ``hot'' slab, the
509     probability density is notched at low speeds and has a substantial
510     shoulder at higher speeds relative to the ideal MB distribution. In
511 gezelter 3609 the cold slab, the opposite notching and shouldering occurs. This
512 skuang 3619 phenomenon is more obvious at higher swapping rates.
513 skuang 3563
514 gezelter 3609 In the velocity distributions, the ideal Gaussian peak is
515     substantially flattened in the hot slab, and is overly sharp (with
516     truncated wings) in the cold slab. This problem is rooted in the
517     mechanism of the swapping method. Continually depleting low (high)
518     speed particles in the high (low) temperature slab is not complemented
519     by diffusions of low (high) speed particles from neighboring slabs,
520     unless the swapping rate is sufficiently small. Simutaneously, surplus
521     low speed particles in the low temperature slab do not have sufficient
522     time to diffuse to neighboring slabs. Since the thermal exchange rate
523     must reach a minimum level to produce an observable thermal gradient,
524     the swapping-method RNEMD has a relatively narrow choice of exchange
525     times that can be utilized.
526 skuang 3578
527 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
528     Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
529     this is simple; upon velocity scaling, a Gaussian distribution remains
530     Gaussian. Although a single scaling move is non-isotropic in three
531     dimensions, our criteria for choosing a set of scaling coefficients
532     helps maintain the distributions as close to isotropic as possible.
533     Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
534     as the previous RNEMD methods but without large perturbations to the
535     velocity distributions in the two slabs.
536    
537 skuang 3568 \begin{figure}
538 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
539 skuang 3619 \caption{Speed distribution for thermal conductivity using
540     ``swapping'' and NIVS-RNEMD methods. Shown is from simulations under
541     ${\langle T^* \rangle = 0.8}$ with an swapping interval of 200
542     time steps (equivalent ${J_z^*\sim 0.2}$). In circled areas,
543     distributions from ``swapping'' RNEMD simulation have deviations
544     from ideal Maxwell-Boltzmann distributions.}
545 skuang 3589 \label{thermalHist}
546 skuang 3568 \end{figure}
547    
548 gezelter 3611
549     \subsubsection*{Shear Viscosity}
550 gezelter 3612 Our calculations (Table \ref{LJ}) show that velocity-scaling
551 gezelter 3611 RNEMD predicted comparable shear viscosities to swap RNEMD method. All
552     the scale method results were from simulations that had a scaling
553     interval of 10 time steps. The average molecular momentum gradients of
554     these samples are shown in Figure \ref{shear} (a) and (b).
555    
556     \begin{figure}
557     \includegraphics[width=\linewidth]{shear}
558     \caption{Average momentum gradients in shear viscosity simulations,
559     using (a) ``swapping'' method and (b) NIVS-RNEMD method
560 skuang 3619 respectively. (c) Temperature difference among $x$ and $y, z$
561     dimensions observed when using NIVS-RNEMD with ${j_z^*(p_x)\sim 0.09}$.}
562 gezelter 3611 \label{shear}
563     \end{figure}
564    
565     However, observations of temperatures along three dimensions show that
566     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
567     two slabs which were scaled. Figure \ref{shear} (c) indicate that with
568     relatively large imposed momentum flux, the temperature difference among $x$
569     and the other two dimensions was significant. This would result from the
570     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
571     momentum gradient is set up, $P_c^x$ would be roughly stable
572     ($<0$). Consequently, scaling factor $x$ would most probably larger
573     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
574     keep increase after most scaling steps. And if there is not enough time
575     for the kinetic energy to exchange among different dimensions and
576     different slabs, the system would finally build up temperature
577     (kinetic energy) difference among the three dimensions. Also, between
578     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
579     are closer to neighbor slabs. This is due to momentum transfer along
580     $z$ dimension between slabs.
581    
582     Although results between scaling and swapping methods are comparable,
583     the inherent temperature inhomogeneity even in relatively low imposed
584     exchange momentum flux simulations makes scaling RNEMD method less
585     attractive than swapping RNEMD in shear viscosity calculation.
586    
587    
588 gezelter 3609 \subsection{Bulk SPC/E water}
589    
590     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
591     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
592     the original swapping RNEMD method. Bedrov {\it et
593 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
594 skuang 3579 center-of-mass velocities instead of single atom velocities in a
595 gezelter 3609 molecule conserves the total kinetic energy and linear momentum. This
596     principle is also adopted in our simulations. Scaling was applied to
597     the center-of-mass velocities of the rigid bodies of SPC/E model water
598     molecules.
599 skuang 3563
600 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
601     molecules were first equilibrated under ambient pressure and
602     temperature conditions using the isobaric-isothermal (NPT)
603     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
604     average volume observed in the NPT simulations, and this was followed
605     by equilibration, first in the canonical (NVT) ensemble, followed by a
606 skuang 3613 100ps period under constant-NVE conditions without any momentum
607     flux. 100ps was allowed to stabilize the system with an imposed
608     momentum transfer to create a gradient, and 1ns was alotted for
609 gezelter 3609 data collection under RNEMD.
610    
611 skuang 3619 In our simulations, temperature gradients were established similar to
612     the previous work. Our simulation results under 318K are in well
613     agreement with those from Bedrov {\it et al.} (Table
614 skuang 3615 \ref{spceThermal}). And both methods yield values in reasonable
615     agreement with experimental value. A larger difference between
616     simulation result and experiment is found under 300K. This could
617     result from the force field that is used in simulation.
618 gezelter 3609
619 skuang 3570 \begin{figure}
620 gezelter 3609 \includegraphics[width=\linewidth]{spceGrad}
621     \caption{Temperature gradients in SPC/E water thermal conductivity
622     simulations.}
623     \label{spceGrad}
624 skuang 3570 \end{figure}
625    
626     \begin{table*}
627 gezelter 3609 \begin{minipage}{\linewidth}
628     \begin{center}
629    
630     \caption{Thermal conductivity of SPC/E water under various
631     imposed thermal gradients. Uncertainties are indicated in
632     parentheses.}
633    
634 skuang 3615 \begin{tabular}{|c|c|ccc|}
635 gezelter 3609 \hline
636 skuang 3615 \multirow{2}{*}{$\langle T\rangle$(K)} &
637     \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
638     \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
639     \mathrm{K}^{-1})$} \\
640     & & This work & Previous simulations\cite{Bedrov:2000} &
641 gezelter 3609 Experiment\cite{WagnerKruse}\\
642     \hline
643 skuang 3615 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
644     \multirow{3}{*}{0.61}\\
645     & 0.81 & 0.770(0.008) & & \\
646     & 1.54 & 0.813(0.007) & & \\
647 gezelter 3609 \hline
648 skuang 3615 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
649     \multirow{2}{*}{0.64}\\
650     & 1.59 & 0.778(0.019) & 0.730 & \\
651     \hline
652 gezelter 3609 \end{tabular}
653     \label{spceThermal}
654     \end{center}
655     \end{minipage}
656     \end{table*}
657 skuang 3570
658 gezelter 3609 \subsection{Crystalline Gold}
659 skuang 3570
660 gezelter 3609 To see how the method performed in a solid, we calculated thermal
661     conductivities using two atomistic models for gold. Several different
662     potential models have been developed that reasonably describe
663     interactions in transition metals. In particular, the Embedded Atom
664     Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc
665     sc})~\cite{Chen90} potential have been used to study a wide range of
666     phenomena in both bulk materials and
667     nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
668     Both potentials are based on a model of a metal which treats the
669     nuclei and core electrons as pseudo-atoms embedded in the electron
670     density due to the valence electrons on all of the other atoms in the
671     system. The {\sc sc} potential has a simple form that closely
672     resembles the Lennard Jones potential,
673     \begin{equation}
674     \label{eq:SCP1}
675     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] ,
676     \end{equation}
677     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
678     \begin{equation}
679     \label{eq:SCP2}
680     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}}.
681     \end{equation}
682     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
683     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
684     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
685     the interactions between the valence electrons and the cores of the
686     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
687     scale, $c_i$ scales the attractive portion of the potential relative
688     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
689     that assures a dimensionless form for $\rho$. These parameters are
690     tuned to various experimental properties such as the density, cohesive
691     energy, and elastic moduli for FCC transition metals. The quantum
692     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
693     including zero-point quantum corrections for different transition
694     metals.\cite{PhysRevB.59.3527} The {\sc eam} functional forms differ
695     slightly from {\sc sc} but the overall method is very similar.
696 skuang 3570
697 gezelter 3609 In this work, we have utilized both the {\sc eam} and the {\sc q-sc}
698     potentials to test the behavior of scaling RNEMD.
699 skuang 3570
700 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
701 skuang 3613 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
702     both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
703 gezelter 3609 pre-equilibration at a target pressure of 1 atm. When equilibrated
704     under NPT conditions, our simulation box expanded by approximately 1\%
705 skuang 3613 in volume. Following adjustment of the box volume, equilibrations in
706     both the canonical and microcanonical ensembles were carried out. With
707     the simulation cell divided evenly into 10 slabs, different thermal
708     gradients were established by applying a set of target thermal
709     transfer fluxes.
710 skuang 3570
711 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
712     \ref{AuThermal}. In these calculations, the end and middle slabs were
713 gezelter 3610 excluded in thermal gradient linear regession. {\sc eam} predicts
714 gezelter 3609 slightly larger thermal conductivities than {\sc q-sc}. However, both
715     values are smaller than experimental value by a factor of more than
716     200. This behavior has been observed previously by Richardson and
717 skuang 3615 Clancy, and has been attributed to the lack of electronic contribution
718     in these force fields.\cite{Clancy:1992} The non-equilibrium MD method
719 skuang 3617 employed in their simulations was only able to give a rough estimation
720     of thermal conductance for {\sc eam} gold, and the result was an
721     average over a wide temperature range (300-800K). Comparatively, our
722     results were based on measurements with linear temperature gradients,
723     and thus of higher reliability and accuracy. It should be noted that
724     the density of the metal being simulated also has an observable effect
725     on thermal conductance. With an expanded lattice, lower thermal
726     conductance is expected (and observed). We also observed a decrease in
727     thermal conductance at higher temperatures, a trend that agrees with
728     experimental measurements.\cite{AshcroftMermin}
729 skuang 3570
730 gezelter 3609 \begin{table*}
731     \begin{minipage}{\linewidth}
732     \begin{center}
733    
734     \caption{Calculated thermal conductivity of crystalline gold
735     using two related force fields. Calculations were done at both
736     experimental and equilibrated densities and at a range of
737 skuang 3617 temperatures and thermal flux rates. Uncertainties are
738     indicated in parentheses. Richardson {\it et
739     al.}\cite{Clancy:1992} gave an estimatioin for {\sc eam} gold
740     of 1.74$\mathrm{W m}^{-1}\mathrm{K}^{-1}$.}
741 gezelter 3609
742     \begin{tabular}{|c|c|c|cc|}
743     \hline
744     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
745     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
746     \hline
747     \multirow{7}{*}{\sc q-sc} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
748     & & & 2.86 & 1.08(0.05)\\
749     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
750     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
751     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
752     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
753     & & & 4.84 & 0.92(0.05)\\
754     \hline
755     \multirow{8}{*}{\sc eam} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
756     & & & 2.06 & 1.37(0.04)\\
757     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
758     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
759     & & & 2.04 & 1.41(0.07)\\
760     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
761     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
762     & & & 4.14 & 1.08(0.05)\\
763     \hline
764     \end{tabular}
765     \label{AuThermal}
766     \end{center}
767     \end{minipage}
768 skuang 3580 \end{table*}
769    
770 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
771     The most attractive aspect of the scaling approach for RNEMD is the
772     ability to use the method in non-homogeneous systems, where molecules
773     of different identities are segregated in different slabs. To test
774     this application, we simulated a Gold (111) / water interface. To
775     construct the interface, a box containing a lattice of 1188 Au atoms
776 skuang 3619 (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
777 gezelter 3609 relax under ambient temperature and pressure. A separate (but
778     identically sized) box of SPC/E water was also equilibrated at ambient
779     conditions. The two boxes were combined by removing all water
780 skuang 3613 molecules within 3 \AA radius of any gold atom. The final
781 gezelter 3609 configuration contained 1862 SPC/E water molecules.
782 skuang 3580
783 gezelter 3609 After simulations of bulk water and crystal gold, a mixture system was
784     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
785     molecules. Spohr potential was adopted in depicting the interaction
786     between metal atom and water molecule.\cite{ISI:000167766600035} A
787     similar protocol of equilibration was followed. Several thermal
788     gradients was built under different target thermal flux. It was found
789     out that compared to our previous simulation systems, the two phases
790     could have large temperature difference even under a relatively low
791     thermal flux.
792    
793    
794 skuang 3581 After simulations of homogeneous water and gold systems using
795     NIVS-RNEMD method were proved valid, calculation of gold/water
796     interfacial thermal conductivity was followed. It is found out that
797     the low interfacial conductance is probably due to the hydrophobic
798 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
799 skuang 3581 density change along $z$-axis, which is perpendicular to the
800     gold/water interface. It is observed that water density significantly
801     decreases when approaching the surface. Under this low thermal
802     conductance, both gold and water phase have sufficient time to
803     eliminate temperature difference inside respectively (Figure
804 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
805 skuang 3581 within respective phase, it is valid to assume that the temperature
806     difference between gold and water on surface would be approximately
807     the same as the difference between the gold and water phase. This
808     assumption enables convenient calculation of $G$ using
809     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
810     layer of water and gold close enough to surface, which would have
811     greater fluctuation and lower accuracy. Reported results (Table
812     \ref{interfaceRes}) are of two orders of magnitude smaller than our
813     calculations on homogeneous systems, and thus have larger relative
814     errors than our calculation results on homogeneous systems.
815 skuang 3573
816 skuang 3571 \begin{figure}
817 skuang 3595 \includegraphics[width=\linewidth]{interface}
818     \caption{Simulation results for Gold/Water interfacial thermal
819     conductivity: (a) Significant water density decrease is observed on
820 skuang 3597 crystalline gold surface, which indicates low surface contact and
821     leads to low thermal conductance. (b) Temperature profiles for a
822     series of simulations. Temperatures of different slabs in the same
823     phase show no significant differences.}
824 skuang 3595 \label{interface}
825 skuang 3571 \end{figure}
826    
827 skuang 3572 \begin{table*}
828 gezelter 3612 \begin{minipage}{\linewidth}
829     \begin{center}
830    
831     \caption{Computed interfacial thermal conductivity ($G$) values
832     for the Au(111) / water interface at ${\langle T\rangle \sim}$
833     300K using a range of energy fluxes. Uncertainties are
834     indicated in parentheses. }
835    
836 gezelter 3616 \begin{tabular}{|cccc| }
837 gezelter 3612 \hline
838     $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
839     T_{water} \rangle$ (K) & $G$
840     (MW/m$^2$/K)\\
841     \hline
842     98.0 & 355.2 & 295.8 & 1.65(0.21) \\
843     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
844     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
845     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
846     \hline
847     \end{tabular}
848     \label{interfaceRes}
849     \end{center}
850     \end{minipage}
851 skuang 3572 \end{table*}
852    
853 skuang 3576
854 skuang 3574 \section{Conclusions}
855     NIVS-RNEMD simulation method is developed and tested on various
856 skuang 3581 systems. Simulation results demonstrate its validity in thermal
857     conductivity calculations, from Lennard-Jones fluid to multi-atom
858     molecule like water and metal crystals. NIVS-RNEMD improves
859 gezelter 3616 non-Boltzmann-Maxwell distributions, which exist inb previous RNEMD
860 skuang 3581 methods. Furthermore, it develops a valid means for unphysical thermal
861     transfer between different species of molecules, and thus extends its
862     applicability to interfacial systems. Our calculation of gold/water
863     interfacial thermal conductivity demonstrates this advantage over
864     previous RNEMD methods. NIVS-RNEMD has also limited application on
865     shear viscosity calculations, but could cause temperature difference
866     among different dimensions under high momentum flux. Modification is
867     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
868     calculations.
869 skuang 3572
870 gezelter 3524 \section{Acknowledgments}
871     Support for this project was provided by the National Science
872     Foundation under grant CHE-0848243. Computational time was provided by
873     the Center for Research Computing (CRC) at the University of Notre
874     Dame. \newpage
875    
876     \bibliography{nivsRnemd}
877 gezelter 3583
878 gezelter 3524 \end{doublespace}
879     \end{document}
880