ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3622
Committed: Thu Aug 5 02:39:16 2010 UTC (13 years, 10 months ago) by skuang
Content type: application/x-tex
File size: 40753 byte(s)
Log Message:
drop a plot, add more swapping details, and some 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 gezelter 3620 velocity and temperature gradients in molecular dynamics simulations
48     of heterogeneous systems. This method extends earlier Reverse
49     Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
50     momentum exchange swapping moves. The standard swapping moves can
51     create non-thermal velocity distributions and are difficult to use
52     for interfacial calculations. By using non-isotropic velocity
53     scaling (NIVS) on the molecules in specific regions of a system, it
54     is possible to impose momentum or thermal flux between regions of a
55     simulation while conserving the linear momentum and total energy of
56     the system. To test the methods, we have computed the thermal
57     conductivity of model liquid and solid systems as well as the
58     interfacial thermal conductivity of a metal-water interface. We
59     find that the NIVS-RNEMD improves the problematic velocity
60 gezelter 3609 distributions that develop in other RNEMD methods.
61 gezelter 3524 \end{abstract}
62    
63     \newpage
64    
65     %\narrowtext
66    
67     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68     % BODY OF TEXT
69     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70    
71     \section{Introduction}
72     The original formulation of Reverse Non-equilibrium Molecular Dynamics
73     (RNEMD) obtains transport coefficients (thermal conductivity and shear
74     viscosity) in a fluid by imposing an artificial momentum flux between
75     two thin parallel slabs of material that are spatially separated in
76 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
77 gezelter 3583 artificial flux is typically created by periodically ``swapping''
78     either the entire momentum vector $\vec{p}$ or single components of
79     this vector ($p_x$) between molecules in each of the two slabs. If
80     the two slabs are separated along the $z$ coordinate, the imposed flux
81     is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
82     response of a simulated system to the imposed momentum flux will
83     typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
84     The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
85     easily obtained by assuming linear response of the system,
86 gezelter 3524 \begin{eqnarray}
87 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
88 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
89 gezelter 3524 \end{eqnarray}
90 gezelter 3600 RNEMD has been widely used to provide computational estimates of
91     thermal conductivities and shear viscosities in a wide range of
92     materials, from liquid copper to both monatomic and molecular fluids
93     (e.g. ionic
94     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}
95 gezelter 3524
96 skuang 3574 \begin{figure}
97     \includegraphics[width=\linewidth]{thermalDemo}
98 gezelter 3583 \caption{RNEMD methods impose an unphysical transfer of momentum or
99     kinetic energy between a ``hot'' slab and a ``cold'' slab in the
100     simulation box. The molecular system responds to this imposed flux
101     by generating a momentum or temperature gradient. The slope of the
102     gradient can then be used to compute transport properties (e.g.
103     shear viscosity and thermal conductivity).}
104 skuang 3574 \label{thermalDemo}
105     \end{figure}
106    
107 skuang 3591 RNEMD is preferable in many ways to the forward NEMD
108 skuang 3592 methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
109     because it imposes what is typically difficult to measure (a flux or
110 gezelter 3600 stress) and it is typically much easier to compute the response
111 gezelter 3609 (momentum gradients or strains). For similar reasons, RNEMD is also
112 skuang 3592 preferable to slowly-converging equilibrium methods for measuring
113     thermal conductivity and shear viscosity (using Green-Kubo
114 skuang 3591 relations\cite{daivis:541,mondello:9327} or the Helfand moment
115     approach of Viscardy {\it et
116 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
117 gezelter 3524 computing difficult to measure quantities.
118    
119     Another attractive feature of RNEMD is that it conserves both total
120     linear momentum and total energy during the swaps (as long as the two
121     molecules have the same identity), so the swapped configurations are
122     typically samples from the same manifold of states in the
123     microcanonical ensemble.
124    
125 gezelter 3620 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered some
126     problems with the original RNEMD swap technique. Notably, large
127 skuang 3565 momentum fluxes (equivalent to frequent momentum swaps between the
128 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
129     momentum distributions in the two slabs, as well as non-linear thermal
130     and velocity distributions along the direction of the imposed flux
131     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
132 gezelter 3620 and proposed self-adjusting metrics for retaining the usability of the
133     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 gezelter 3620 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i v_{i\alpha} \\
185     P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j v_{j\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 3620 the instantaneous linear momenta in each of the two slabs.
191 gezelter 3524 \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 gezelter 3620 \sum_{\alpha = x,y,z} \left\{ K_h^\alpha + K_c^\alpha \right\} = \sum_{\alpha = x,y,z}
204     \left\{ \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha \right\}
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 gezelter 3620 problem.\cite{Hoffman:2001sf} One simplification is to maintain
273     velocity scalings that are {\it as isotropic as possible}. To do
274     this, we impose $x=y$, and treat both the constraint and flux
275     ellipsoids as 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 gezelter 3620 with a variable frequency after the molecular dynamics (MD) steps. We
334     have tested the method in a variety of different systems, including:
335     homogeneous fluids (Lennard-Jones and SPC/E water), crystalline
336     solids, using both the embedded atom method
337     (EAM)~\cite{PhysRevB.33.7983} and quantum Sutton-Chen
338     (QSC)~\cite{PhysRevB.59.3527} models for Gold, and heterogeneous
339     interfaces (QSC gold - SPC/E water). The last of these systems would
340     have been difficult to study using previous RNEMD methods, but the
341     current method can easily provide estimates of the interfacial thermal
342     conductivity ($G$).
343 gezelter 3524
344 gezelter 3609 \subsection{Simulation Cells}
345 gezelter 3524
346 gezelter 3609 In each of the systems studied, the dynamics was carried out in a
347     rectangular simulation cell using periodic boundary conditions in all
348     three dimensions. The cells were longer along the $z$ axis and the
349     space was divided into $N$ slabs along this axis (typically $N=20$).
350 skuang 3613 The top slab ($n=1$) was designated the ``hot'' slab, while the
351     central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
352 gezelter 3609 cases, simulations were first thermalized in canonical ensemble (NVT)
353     using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
354 gezelter 3600 microcanonical ensemble (NVE) before introducing any non-equilibrium
355     method.
356 skuang 3531
357 gezelter 3609 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
358 skuang 3531
359 gezelter 3609 In order to compare our new methodology with the original
360     M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
361 skuang 3622 first performed simulations using the original technique. With a
362     certain interval, swapping is performed between the hot and the cold
363     slab. For thermal exchange, particle with minimum speed in the hot
364     slab and the one with maximum speed in the cold slab swap all three
365     velocity components; for shear stress simulations, particle with the
366     most negative $v_x$ in the hot slab and the one with the most positive
367     $v_x$ in the cold slab swap only the $v_x$ component.
368 skuang 3531
369 gezelter 3609 \subsection{RNEMD with NIVS scaling}
370    
371     For each simulation utilizing the swapping method, a corresponding
372     NIVS-RNEMD simulation was carried out using a target momentum flux set
373 gezelter 3620 to produce the same flux experienced in the swapping simulation.
374 gezelter 3609
375 gezelter 3620 To test the temperature homogeneity, directional momentum and
376     temperature distributions were accumulated for molecules in each of
377     the slabs. Transport coefficients were computed using the temperature
378     (and momentum) gradients across the $z$-axis as well as the total
379     momentum flux the system experienced during data collection portion of
380     the simulation.
381 gezelter 3609
382     \subsection{Shear viscosities}
383    
384     The momentum flux was calculated using the total non-physical momentum
385     transferred (${P_x}$) and the data collection time ($t$):
386 skuang 3534 \begin{equation}
387     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
388     \end{equation}
389 gezelter 3609 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
390     box. The factor of two in the denominator is present because physical
391 gezelter 3620 momentum transfer between the slabs occurs in two directions ($+z$ and
392     $-z$). The velocity gradient ${\langle \partial v_x /\partial z
393     \rangle}$ was obtained using linear regression of the mean $x$
394     component of the velocity, $\langle v_x \rangle$, in each of the bins.
395     For Lennard-Jones simulations, shear viscosities are reported in
396     reduced units (${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$).
397 skuang 3532
398 gezelter 3609 \subsection{Thermal Conductivities}
399 skuang 3534
400 gezelter 3620 The energy flux was calculated in a similar manner to the momentum
401     flux, using the total non-physical energy transferred (${E_{total}}$)
402     and the data collection time $t$:
403 skuang 3534 \begin{equation}
404     J_z = \frac{E_{total}}{2 t L_x L_y}
405     \end{equation}
406 gezelter 3609 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
407     obtained by a linear regression of the temperature profile. For
408     Lennard-Jones simulations, thermal conductivities are reported in
409     reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
410     k_B^{-1}\varepsilon^{-1/2}}$).
411 skuang 3534
412 gezelter 3609 \subsection{Interfacial Thermal Conductivities}
413 skuang 3563
414 gezelter 3620 For interfaces with a relatively low interfacial conductance, the bulk
415     regions on either side of an interface rapidly come to a state in
416     which the two phases have relatively homogeneous (but distinct)
417     temperatures. The interfacial thermal conductivity $G$ can therefore
418     be approximated as:
419 skuang 3573
420     \begin{equation}
421     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
422     \langle T_{water}\rangle \right)}
423     \label{interfaceCalc}
424     \end{equation}
425 gezelter 3609 where ${E_{total}}$ is the imposed non-physical kinetic energy
426     transfer and ${\langle T_{gold}\rangle}$ and ${\langle
427     T_{water}\rangle}$ are the average observed temperature of gold and
428 gezelter 3620 water phases respectively. If the interfacial conductance is {\it
429     not} small, it is also be possible to compute the interfacial
430     thermal conductivity using this method utilizing the change in the
431     slope of the thermal gradient ($\partial^2 \langle T \rangle / \partial
432     z^2$) at the interface.
433 skuang 3573
434 gezelter 3609 \section{Results}
435 skuang 3538
436 gezelter 3609 \subsection{Lennard-Jones Fluid}
437     2592 Lennard-Jones atoms were placed in an orthorhombic cell
438     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
439     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
440     direct comparison between our results and previous methods. These
441     simulations were carried out with a reduced timestep ${\tau^* =
442     4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
443     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
444 skuang 3617 conductivity calculations, simulations were run under reduced
445 gezelter 3609 temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
446 skuang 3617 ensemble. The simulations included $10^5$ steps of equilibration
447 gezelter 3609 without any momentum flux, $10^5$ steps of stablization with an
448     imposed momentum transfer to create a gradient, and $10^6$ steps of
449     data collection under RNEMD.
450    
451 gezelter 3611 \subsubsection*{Thermal Conductivity}
452    
453 gezelter 3609 Our thermal conductivity calculations show that the NIVS method agrees
454 skuang 3618 well with the swapping method. Five different swap intervals were
455 gezelter 3620 tested (Table \ref{LJ}). Similar thermal gradients were observed with
456     similar thermal flux under the two different methods (Figure
457     \ref{thermalGrad}). Furthermore, with appropriate choice of scaling
458 skuang 3622 variables, the temperatures along $x$, $y$ and $z$ axes showed no
459     observable difference (Figure \ref{thermalGrad} c). The
460 gezelter 3620 system is able to maintain temperature homogeneity even under high
461     flux.
462 gezelter 3609
463 skuang 3563 \begin{table*}
464 gezelter 3609 \begin{minipage}{\linewidth}
465     \begin{center}
466 skuang 3538
467 gezelter 3612 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
468     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
469     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
470     at various momentum fluxes. The original swapping method and
471     the velocity scaling method give similar results.
472     Uncertainties are indicated in parentheses.}
473 gezelter 3609
474 gezelter 3612 \begin{tabular}{|cc|cc|cc|}
475 gezelter 3609 \hline
476 gezelter 3612 \multicolumn{2}{|c}{Momentum Exchange} &
477     \multicolumn{2}{|c}{Swapping RNEMD} &
478 gezelter 3609 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
479     \hline
480 gezelter 3612 \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
481     \multirow{2}{*}{$\lambda^*_{swap}$} &
482     \multirow{2}{*}{$\eta^*_{swap}$} &
483     \multirow{2}{*}{$\lambda^*_{scale}$} &
484     \multirow{2}{*}{$\eta^*_{scale}$} \\
485 skuang 3617 & $j_z^*(p_x)$ (reduced units) & & & & \\
486 gezelter 3609 \hline
487 skuang 3617 250 & 0.16 & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
488 gezelter 3612 500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
489     1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
490     2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
491 skuang 3617 2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
492 gezelter 3609 \hline
493     \end{tabular}
494 gezelter 3612 \label{LJ}
495 gezelter 3609 \end{center}
496     \end{minipage}
497 skuang 3563 \end{table*}
498    
499     \begin{figure}
500 gezelter 3612 \includegraphics[width=\linewidth]{thermalGrad}
501 gezelter 3620 \caption{The NIVS-RNEMD method (b) creates similar temperature gradients
502 skuang 3619 compared with the swapping method (a) under a variety of imposed
503     kinetic energy flux values. Furthermore, the implementation of
504     Non-Isotropic Velocity Scaling does not cause temperature
505 gezelter 3620 differences among the three dimensions (c).}
506 gezelter 3612 \label{thermalGrad}
507 skuang 3563 \end{figure}
508    
509 gezelter 3612 \subsubsection*{Velocity Distributions}
510    
511 gezelter 3609 During these simulations, velocities were recorded every 1000 steps
512 gezelter 3620 and were used to produce distributions of both velocity and speed in
513 gezelter 3609 each of the slabs. From these distributions, we observed that under
514 skuang 3613 relatively high non-physical kinetic energy flux, the speed of
515 gezelter 3609 molecules in slabs where swapping occured could deviate from the
516     Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
517 gezelter 3620 and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
518     distributions deviate from an ideal distribution. In the ``hot'' slab,
519     the probability density is notched at low speeds and has a substantial
520 skuang 3619 shoulder at higher speeds relative to the ideal MB distribution. In
521 gezelter 3609 the cold slab, the opposite notching and shouldering occurs. This
522 skuang 3619 phenomenon is more obvious at higher swapping rates.
523 skuang 3563
524 gezelter 3620 The peak of the velocity distribution is substantially flattened in
525     the hot slab, and is overly sharp (with truncated wings) in the cold
526     slab. This problem is rooted in the mechanism of the swapping method.
527     Continually depleting low (high) speed particles in the high (low)
528     temperature slab is not complemented by diffusions of low (high) speed
529     particles from neighboring slabs, unless the swapping rate is
530     sufficiently small. Simutaneously, surplus low speed particles in the
531     low temperature slab do not have sufficient time to diffuse to
532     neighboring slabs. Since the thermal exchange rate must reach a
533     minimum level to produce an observable thermal gradient, the
534     swapping-method RNEMD has a relatively narrow choice of exchange times
535     that can be utilized.
536 skuang 3578
537 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
538     Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
539     this is simple; upon velocity scaling, a Gaussian distribution remains
540     Gaussian. Although a single scaling move is non-isotropic in three
541     dimensions, our criteria for choosing a set of scaling coefficients
542     helps maintain the distributions as close to isotropic as possible.
543     Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
544     as the previous RNEMD methods but without large perturbations to the
545     velocity distributions in the two slabs.
546    
547 skuang 3568 \begin{figure}
548 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
549 skuang 3619 \caption{Speed distribution for thermal conductivity using
550     ``swapping'' and NIVS-RNEMD methods. Shown is from simulations under
551     ${\langle T^* \rangle = 0.8}$ with an swapping interval of 200
552     time steps (equivalent ${J_z^*\sim 0.2}$). In circled areas,
553     distributions from ``swapping'' RNEMD simulation have deviations
554     from ideal Maxwell-Boltzmann distributions.}
555 skuang 3589 \label{thermalHist}
556 skuang 3568 \end{figure}
557    
558 gezelter 3611
559     \subsubsection*{Shear Viscosity}
560 gezelter 3620 Our calculations (Table \ref{LJ}) show that velocity-scaling RNEMD
561     predicted comparable shear viscosities to swap RNEMD method. The
562     average molecular momentum gradients of these samples are shown in
563     Figure \ref{shear} (a) and (b).
564 gezelter 3611
565     \begin{figure}
566     \includegraphics[width=\linewidth]{shear}
567     \caption{Average momentum gradients in shear viscosity simulations,
568     using (a) ``swapping'' method and (b) NIVS-RNEMD method
569 skuang 3619 respectively. (c) Temperature difference among $x$ and $y, z$
570     dimensions observed when using NIVS-RNEMD with ${j_z^*(p_x)\sim 0.09}$.}
571 gezelter 3611 \label{shear}
572     \end{figure}
573    
574 gezelter 3620 Observations of the three one-dimensional temperatures in each of the
575     slabs shows that NIVS-RNEMD does produce some thermal anisotropy,
576     particularly in the hot and cold slabs. Figure \ref{shear} (c)
577     indicates that with a relatively large imposed momentum flux,
578     $j_z(p_x)$, the $x$ direction approaches a different temperature from
579     the $y$ and $z$ directions in both the hot and cold bins. This is an
580     artifact of the scaling constraints. After the momentum gradient has
581     been established, $P_c^x < 0$. Consequently, the scaling factor $x$
582     is nearly always greater than one in order to satisfy the constraints.
583     This will continually increase the kinetic energy in $x$-dimension,
584     $K_c^x$. If there is not enough time for the kinetic energy to
585     exchange among different directions and different slabs, the system
586     will exhibit the observed thermal anisotropy in the hot and cold bins.
587 gezelter 3611
588     Although results between scaling and swapping methods are comparable,
589 gezelter 3620 the inherent temperature anisotropy does make NIVS-RNEMD method less
590     attractive than swapping RNEMD for shear viscosity calculations. We
591     note that this problem appears only when momentum flux is applied, and
592     does not appear in thermal flux calculations.
593 gezelter 3611
594 gezelter 3609 \subsection{Bulk SPC/E water}
595    
596     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
597     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
598     the original swapping RNEMD method. Bedrov {\it et
599 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
600 skuang 3579 center-of-mass velocities instead of single atom velocities in a
601 gezelter 3609 molecule conserves the total kinetic energy and linear momentum. This
602 gezelter 3620 principle is also adopted Fin our simulations. Scaling was applied to
603 gezelter 3609 the center-of-mass velocities of the rigid bodies of SPC/E model water
604     molecules.
605 skuang 3563
606 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
607     molecules were first equilibrated under ambient pressure and
608     temperature conditions using the isobaric-isothermal (NPT)
609     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
610     average volume observed in the NPT simulations, and this was followed
611     by equilibration, first in the canonical (NVT) ensemble, followed by a
612 gezelter 3620 100~ps period under constant-NVE conditions without any momentum flux.
613     Another 100~ps was allowed to stabilize the system with an imposed
614     momentum transfer to create a gradient, and 1~ns was allotted for data
615     collection under RNEMD.
616 gezelter 3609
617 gezelter 3620 In our simulations, the established temperature gradients were similar
618     to the previous work. Our simulation results at 318K are in good
619 skuang 3619 agreement with those from Bedrov {\it et al.} (Table
620 skuang 3615 \ref{spceThermal}). And both methods yield values in reasonable
621 gezelter 3620 agreement with experimental values.
622 gezelter 3609
623 skuang 3570 \begin{table*}
624 gezelter 3609 \begin{minipage}{\linewidth}
625     \begin{center}
626    
627     \caption{Thermal conductivity of SPC/E water under various
628     imposed thermal gradients. Uncertainties are indicated in
629     parentheses.}
630    
631 skuang 3615 \begin{tabular}{|c|c|ccc|}
632 gezelter 3609 \hline
633 skuang 3615 \multirow{2}{*}{$\langle T\rangle$(K)} &
634     \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
635     \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
636     \mathrm{K}^{-1})$} \\
637     & & This work & Previous simulations\cite{Bedrov:2000} &
638 gezelter 3609 Experiment\cite{WagnerKruse}\\
639     \hline
640 skuang 3615 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
641     \multirow{3}{*}{0.61}\\
642     & 0.81 & 0.770(0.008) & & \\
643     & 1.54 & 0.813(0.007) & & \\
644 gezelter 3609 \hline
645 skuang 3615 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
646     \multirow{2}{*}{0.64}\\
647     & 1.59 & 0.778(0.019) & 0.730 & \\
648     \hline
649 gezelter 3609 \end{tabular}
650     \label{spceThermal}
651     \end{center}
652     \end{minipage}
653     \end{table*}
654 skuang 3570
655 gezelter 3609 \subsection{Crystalline Gold}
656 skuang 3570
657 gezelter 3609 To see how the method performed in a solid, we calculated thermal
658     conductivities using two atomistic models for gold. Several different
659     potential models have been developed that reasonably describe
660     interactions in transition metals. In particular, the Embedded Atom
661 gezelter 3620 Model (EAM)~\cite{PhysRevB.33.7983} and Sutton-Chen (SC)~\cite{Chen90}
662     potential have been used to study a wide range of phenomena in both
663     bulk materials and
664 gezelter 3609 nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
665     Both potentials are based on a model of a metal which treats the
666     nuclei and core electrons as pseudo-atoms embedded in the electron
667     density due to the valence electrons on all of the other atoms in the
668 gezelter 3620 system. The SC potential has a simple form that closely resembles the
669     Lennard Jones potential,
670 gezelter 3609 \begin{equation}
671     \label{eq:SCP1}
672     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] ,
673     \end{equation}
674     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
675     \begin{equation}
676     \label{eq:SCP2}
677     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}}.
678     \end{equation}
679     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
680     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
681     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
682     the interactions between the valence electrons and the cores of the
683     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
684     scale, $c_i$ scales the attractive portion of the potential relative
685     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
686     that assures a dimensionless form for $\rho$. These parameters are
687     tuned to various experimental properties such as the density, cohesive
688     energy, and elastic moduli for FCC transition metals. The quantum
689 gezelter 3620 Sutton-Chen (QSC) formulation matches these properties while including
690     zero-point quantum corrections for different transition
691     metals.\cite{PhysRevB.59.3527} The EAM functional forms differ
692     slightly from SC but the overall method is very similar.
693 skuang 3570
694 gezelter 3620 In this work, we have utilized both the EAM and the QSC potentials to
695     test the behavior of scaling RNEMD.
696 skuang 3570
697 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
698 skuang 3613 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
699     both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
700 gezelter 3609 pre-equilibration at a target pressure of 1 atm. When equilibrated
701     under NPT conditions, our simulation box expanded by approximately 1\%
702 skuang 3613 in volume. Following adjustment of the box volume, equilibrations in
703     both the canonical and microcanonical ensembles were carried out. With
704     the simulation cell divided evenly into 10 slabs, different thermal
705     gradients were established by applying a set of target thermal
706     transfer fluxes.
707 skuang 3570
708 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
709     \ref{AuThermal}. In these calculations, the end and middle slabs were
710 gezelter 3620 excluded in thermal gradient linear regession. EAM predicts slightly
711     larger thermal conductivities than QSC. However, both values are
712     smaller than experimental value by a factor of more than 200. This
713     behavior has been observed previously by Richardson and Clancy, and
714     has been attributed to the lack of electronic contribution in these
715     force fields.\cite{Clancy:1992} It should be noted that the density of
716     the metal being simulated has an effect on thermal conductance. With
717     an expanded lattice, lower thermal conductance is expected (and
718     observed). We also observed a decrease in thermal conductance at
719     higher temperatures, a trend that agrees with experimental
720     measurements.\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 skuang 3617 temperatures and thermal flux rates. Uncertainties are
730     indicated in parentheses. Richardson {\it et
731 gezelter 3621 al.}\cite{Clancy:1992} give an estimate of 1.74 $\mathrm{W
732     m}^{-1}\mathrm{K}^{-1}$ for EAM gold
733     at a density of 19.263 g / cm$^3$.}
734 gezelter 3609
735     \begin{tabular}{|c|c|c|cc|}
736     \hline
737     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
738     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
739     \hline
740 gezelter 3621 \multirow{7}{*}{QSC} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
741 gezelter 3609 & & & 2.86 & 1.08(0.05)\\
742     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
743     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
744     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
745     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
746     & & & 4.84 & 0.92(0.05)\\
747     \hline
748 gezelter 3621 \multirow{8}{*}{EAM} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
749 gezelter 3609 & & & 2.06 & 1.37(0.04)\\
750     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
751     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
752     & & & 2.04 & 1.41(0.07)\\
753     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
754     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
755     & & & 4.14 & 1.08(0.05)\\
756     \hline
757     \end{tabular}
758     \label{AuThermal}
759     \end{center}
760     \end{minipage}
761 skuang 3580 \end{table*}
762    
763 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
764     The most attractive aspect of the scaling approach for RNEMD is the
765     ability to use the method in non-homogeneous systems, where molecules
766     of different identities are segregated in different slabs. To test
767     this application, we simulated a Gold (111) / water interface. To
768     construct the interface, a box containing a lattice of 1188 Au atoms
769 skuang 3619 (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
770 gezelter 3609 relax under ambient temperature and pressure. A separate (but
771     identically sized) box of SPC/E water was also equilibrated at ambient
772     conditions. The two boxes were combined by removing all water
773 skuang 3613 molecules within 3 \AA radius of any gold atom. The final
774 gezelter 3609 configuration contained 1862 SPC/E water molecules.
775 skuang 3580
776 gezelter 3620 The Spohr potential was adopted in depicting the interaction between
777     metal atoms and water molecules.\cite{ISI:000167766600035} A similar
778     protocol of equilibration to our water simulations was followed. We
779     observed that the two phases developed large temperature differences
780     even under a relatively low thermal flux.
781 gezelter 3609
782 gezelter 3620 The low interfacial conductance is probably due to the hydrophobic
783 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
784 skuang 3581 density change along $z$-axis, which is perpendicular to the
785     gold/water interface. It is observed that water density significantly
786     decreases when approaching the surface. Under this low thermal
787     conductance, both gold and water phase have sufficient time to
788     eliminate temperature difference inside respectively (Figure
789 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
790 skuang 3581 within respective phase, it is valid to assume that the temperature
791     difference between gold and water on surface would be approximately
792     the same as the difference between the gold and water phase. This
793 gezelter 3620 assumption enables convenient calculation of $G$ using Eq.
794     \ref{interfaceCalc} instead of measuring temperatures of thin layer of
795     water and gold close enough to surface, which would have greater
796     fluctuation and lower accuracy. Reported results (Table
797 skuang 3581 \ref{interfaceRes}) are of two orders of magnitude smaller than our
798     calculations on homogeneous systems, and thus have larger relative
799     errors than our calculation results on homogeneous systems.
800 skuang 3573
801 skuang 3571 \begin{figure}
802 skuang 3595 \includegraphics[width=\linewidth]{interface}
803     \caption{Simulation results for Gold/Water interfacial thermal
804     conductivity: (a) Significant water density decrease is observed on
805 skuang 3597 crystalline gold surface, which indicates low surface contact and
806     leads to low thermal conductance. (b) Temperature profiles for a
807     series of simulations. Temperatures of different slabs in the same
808     phase show no significant differences.}
809 skuang 3595 \label{interface}
810 skuang 3571 \end{figure}
811    
812 skuang 3572 \begin{table*}
813 gezelter 3612 \begin{minipage}{\linewidth}
814     \begin{center}
815    
816     \caption{Computed interfacial thermal conductivity ($G$) values
817     for the Au(111) / water interface at ${\langle T\rangle \sim}$
818     300K using a range of energy fluxes. Uncertainties are
819     indicated in parentheses. }
820    
821 gezelter 3616 \begin{tabular}{|cccc| }
822 gezelter 3612 \hline
823     $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
824     T_{water} \rangle$ (K) & $G$
825     (MW/m$^2$/K)\\
826     \hline
827     98.0 & 355.2 & 295.8 & 1.65(0.21) \\
828     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
829     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
830     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
831     \hline
832     \end{tabular}
833     \label{interfaceRes}
834     \end{center}
835     \end{minipage}
836 skuang 3572 \end{table*}
837    
838 skuang 3576
839 skuang 3574 \section{Conclusions}
840     NIVS-RNEMD simulation method is developed and tested on various
841 skuang 3581 systems. Simulation results demonstrate its validity in thermal
842     conductivity calculations, from Lennard-Jones fluid to multi-atom
843     molecule like water and metal crystals. NIVS-RNEMD improves
844 gezelter 3616 non-Boltzmann-Maxwell distributions, which exist inb previous RNEMD
845 skuang 3581 methods. Furthermore, it develops a valid means for unphysical thermal
846     transfer between different species of molecules, and thus extends its
847     applicability to interfacial systems. Our calculation of gold/water
848     interfacial thermal conductivity demonstrates this advantage over
849     previous RNEMD methods. NIVS-RNEMD has also limited application on
850     shear viscosity calculations, but could cause temperature difference
851     among different dimensions under high momentum flux. Modification is
852     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
853     calculations.
854 skuang 3572
855 gezelter 3524 \section{Acknowledgments}
856     Support for this project was provided by the National Science
857     Foundation under grant CHE-0848243. Computational time was provided by
858     the Center for Research Computing (CRC) at the University of Notre
859     Dame. \newpage
860    
861     \bibliography{nivsRnemd}
862 gezelter 3583
863 gezelter 3524 \end{doublespace}
864     \end{document}
865