ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3647
Committed: Fri Sep 17 21:02:21 2010 UTC (13 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 43006 byte(s)
Log Message:
answer the question why nivs is better than swapping for inhomogeneous
system.

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 gezelter 3631 the system. To test the method, we have computed the thermal
57 gezelter 3620 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 skuang 3645 (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 3631 computing and integrating long-time correlation functions that are
118     subject to noise issues.
119 gezelter 3524
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 skuang 3645 microcanonical ensemble. Furthermore, the method is applicable with
125     different ensembles, unlike the heat-exchange algorithm proposed by
126     Hafskjold {\it et al.} \cite{HeX:1994,HeX:1993}, which is incompatible
127     with non-microcanonical ensemble.
128 gezelter 3524
129 gezelter 3620 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered some
130     problems with the original RNEMD swap technique. Notably, large
131 skuang 3565 momentum fluxes (equivalent to frequent momentum swaps between the
132 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
133     momentum distributions in the two slabs, as well as non-linear thermal
134     and velocity distributions along the direction of the imposed flux
135     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
136 gezelter 3620 and proposed self-adjusting metrics for retaining the usability of the
137     method.
138 gezelter 3524
139     In this paper, we develop and test a method for non-isotropic velocity
140 gezelter 3600 scaling (NIVS) which retains the desirable features of RNEMD
141 gezelter 3524 (conservation of linear momentum and total energy, compatibility with
142     periodic boundary conditions) while establishing true thermal
143 gezelter 3600 distributions in each of the two slabs. In the next section, we
144 gezelter 3583 present the method for determining the scaling constraints. We then
145 gezelter 3600 test the method on both liquids and solids as well as a non-isotropic
146     liquid-solid interface and show that it is capable of providing
147 gezelter 3524 reasonable estimates of the thermal conductivity and shear viscosity
148 gezelter 3600 in all of these cases.
149 gezelter 3524
150     \section{Methodology}
151 gezelter 3583 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
152     periodic system is partitioned into a series of thin slabs along one
153 gezelter 3524 axis ($z$). One of the slabs at the end of the periodic box is
154     designated the ``hot'' slab, while the slab in the center of the box
155     is designated the ``cold'' slab. The artificial momentum flux will be
156     established by transferring momentum from the cold slab and into the
157     hot slab.
158    
159     Rather than using momentum swaps, we use a series of velocity scaling
160 gezelter 3583 moves. For molecules $\{i\}$ located within the cold slab,
161 gezelter 3524 \begin{equation}
162 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
163     x & 0 & 0 \\
164     0 & y & 0 \\
165     0 & 0 & z \\
166 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
167     \end{equation}
168 gezelter 3600 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
169     of the three directions in the system. Likewise, the molecules
170     $\{j\}$ located in the hot slab will see a concomitant scaling of
171     velocities,
172 gezelter 3524 \begin{equation}
173 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
174     x^\prime & 0 & 0 \\
175     0 & y^\prime & 0 \\
176     0 & 0 & z^\prime \\
177 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
178     \end{equation}
179    
180     Conservation of linear momentum in each of the three directions
181 gezelter 3583 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
182 gezelter 3524 parameters together:
183     \begin{equation}
184 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
185 gezelter 3524 \end{equation}
186     where
187 skuang 3565 \begin{eqnarray}
188 gezelter 3620 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i v_{i\alpha} \\
189     P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j v_{j\alpha}
190 gezelter 3524 \label{eq:momentumdef}
191 skuang 3565 \end{eqnarray}
192 skuang 3528 Therefore, for each of the three directions, the hot scaling
193     parameters are a simple function of the cold scaling parameters and
194 gezelter 3620 the instantaneous linear momenta in each of the two slabs.
195 gezelter 3524 \begin{equation}
196 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
197 gezelter 3524 \label{eq:hotcoldscaling}
198     \end{equation}
199 skuang 3528 where
200     \begin{equation}
201     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
202     \end{equation}
203     for convenience.
204 gezelter 3524
205     Conservation of total energy also places constraints on the scaling:
206     \begin{equation}
207 gezelter 3620 \sum_{\alpha = x,y,z} \left\{ K_h^\alpha + K_c^\alpha \right\} = \sum_{\alpha = x,y,z}
208     \left\{ \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha \right\}
209 gezelter 3524 \end{equation}
210 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
211     $K_c^\alpha$, are computed for each of the three directions in a
212     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
213     Substituting in the expressions for the hot scaling parameters
214     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
215 gezelter 3583 {\it constraint ellipsoid}:
216 gezelter 3524 \begin{equation}
217 gezelter 3600 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
218     c_\alpha \right) = 0
219 gezelter 3524 \label{eq:constraintEllipsoid}
220     \end{equation}
221     where the constants are obtained from the instantaneous values of the
222     linear momenta and kinetic energies for the hot and cold slabs,
223 skuang 3565 \begin{eqnarray}
224 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
225     \left(p_\alpha\right)^2\right) \\
226     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
227 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
228 gezelter 3524 \label{eq:constraintEllipsoidConsts}
229 skuang 3565 \end{eqnarray}
230 gezelter 3583 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
231 gezelter 3600 cold slab scaling parameters which, when applied, preserve the linear
232     momentum of the system in all three directions as well as total
233     kinetic energy.
234 gezelter 3524
235 gezelter 3600 The goal of using these velocity scaling variables is to transfer
236 gezelter 3609 kinetic energy from the cold slab to the hot slab. If the hot and
237     cold slabs are separated along the z-axis, the energy flux is given
238     simply by the decrease in kinetic energy of the cold bin:
239 gezelter 3524 \begin{equation}
240 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
241 gezelter 3524 \end{equation}
242     The expression for the energy flux can be re-written as another
243     ellipsoid centered on $(x,y,z) = 0$:
244     \begin{equation}
245 skuang 3604 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
246     K_c^\alpha -J_z \Delta t
247 gezelter 3524 \label{eq:fluxEllipsoid}
248     \end{equation}
249 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
250 gezelter 3600 both by the target flux, $J_z$ as well as the instantaneous values of
251     the kinetic energy components in the cold bin.
252 gezelter 3524
253     To satisfy an energetic flux as well as the conservation constraints,
254 gezelter 3600 we must determine the points ${x,y,z}$ that lie on both the constraint
255     ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
256     (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
257     ellipsoids in 3-dimensional space.
258 gezelter 3524
259 gezelter 3569 \begin{figure}
260     \includegraphics[width=\linewidth]{ellipsoids}
261 gezelter 3600 \caption{Velocity scaling coefficients which maintain both constant
262     energy and constant linear momentum of the system lie on the surface
263     of the {\it constraint ellipsoid} while points which generate the
264     target momentum flux lie on the surface of the {\it flux ellipsoid}.
265     The velocity distributions in the cold bin are scaled by only those
266     points which lie on both ellipsoids.}
267 gezelter 3569 \label{ellipsoids}
268     \end{figure}
269    
270 gezelter 3600 Since ellipsoids can be expressed as polynomials up to second order in
271     each of the three coordinates, finding the the intersection points of
272     two ellipsoids is isomorphic to finding the roots a polynomial of
273     degree 16. There are a number of polynomial root-finding methods in
274 skuang 3614 the literature,\cite{Hoffman:2001sf,384119} but numerically finding
275     the roots of high-degree polynomials is generally an ill-conditioned
276 gezelter 3620 problem.\cite{Hoffman:2001sf} One simplification is to maintain
277     velocity scalings that are {\it as isotropic as possible}. To do
278     this, we impose $x=y$, and treat both the constraint and flux
279     ellipsoids as 2-dimensional ellipses. In reduced dimensionality, the
280 gezelter 3600 intersecting-ellipse problem reduces to finding the roots of
281 gezelter 3609 polynomials of degree 4.
282 gezelter 3600
283     Depending on the target flux and current velocity distributions, the
284     ellipsoids can have between 0 and 4 intersection points. If there are
285     no intersection points, it is not possible to satisfy the constraints
286     while performing a non-equilibrium scaling move, and no change is made
287     to the dynamics.
288    
289     With multiple intersection points, any of the scaling points will
290     conserve the linear momentum and kinetic energy of the system and will
291     generate the correct target flux. Although this method is inherently
292     non-isotropic, the goal is still to maintain the system as close to an
293     isotropic fluid as possible. With this in mind, we would like the
294     kinetic energies in the three different directions could become as
295     close as each other as possible after each scaling. Simultaneously,
296 gezelter 3631 one would also like each scaling to be as gentle as possible, i.e.
297 gezelter 3632 ${x,y,z \rightarrow 1}$, in order to avoid large perturbations to the
298 gezelter 3631 system. To do this, we pick the intersection point which maintains
299     the three scaling variables ${x, y, z}$ as well as the ratio of
300     kinetic energies ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to
301     1.
302 gezelter 3600
303     After the valid scaling parameters are arrived at by solving geometric
304     intersection problems in $x, y, z$ space in order to obtain cold slab
305     scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
306     determine the conjugate hot slab scaling variables.
307    
308     \subsection{Introducing shear stress via velocity scaling}
309 gezelter 3609 It is also possible to use this method to magnify the random
310     fluctuations of the average momentum in each of the bins to induce a
311     momentum flux. Doing this repeatedly will create a shear stress on
312     the system which will respond with an easily-measured strain. The
313     momentum flux (say along the $x$-direction) may be defined as:
314 gezelter 3524 \begin{equation}
315 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
316 skuang 3531 \label{eq:fluxPlane}
317 gezelter 3524 \end{equation}
318 gezelter 3600 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
319     its position governed both by a target value, $j_z(p_x)$ as well as
320     the instantaneous value of the momentum along the $x$-direction.
321 gezelter 3524
322 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
323     constraints, we must determine the points ${x,y,z}$ which lie on both
324     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
325     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
326 gezelter 3600 ellipsoid and a plane in 3-dimensional space.
327 gezelter 3524
328 gezelter 3600 In the case of momentum flux transfer, we also impose another
329 gezelter 3609 constraint to set the kinetic energy transfer as zero. In other
330     words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
331 gezelter 3600 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
332     set of quartic equations to the above kinetic energy transfer problem.
333 gezelter 3524
334 gezelter 3600 \section{Computational Details}
335 gezelter 3583
336 gezelter 3609 We have implemented this methodology in our molecular dynamics code,
337     OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
338 gezelter 3620 with a variable frequency after the molecular dynamics (MD) steps. We
339     have tested the method in a variety of different systems, including:
340     homogeneous fluids (Lennard-Jones and SPC/E water), crystalline
341     solids, using both the embedded atom method
342     (EAM)~\cite{PhysRevB.33.7983} and quantum Sutton-Chen
343     (QSC)~\cite{PhysRevB.59.3527} models for Gold, and heterogeneous
344 skuang 3645 interfaces (QSC gold - SPC/E water). Even though previous RNEMD
345     methods might remain usable for the last of these systems, energy
346     transfer from imaginary elastic collisions would be less effective
347     when the two particles involved have larger mass difference, and thus
348 skuang 3647 affect the actuall implementation of these methods. In this case, an
349     imarginary elastic collision between gold atom and water molecule has
350     a kinetic energy transfer efficiency of only about 30\%. However, our
351 gezelter 3620 current method can easily provide estimates of the interfacial thermal
352     conductivity ($G$).
353 gezelter 3524
354 gezelter 3609 \subsection{Simulation Cells}
355 gezelter 3524
356 gezelter 3609 In each of the systems studied, the dynamics was carried out in a
357     rectangular simulation cell using periodic boundary conditions in all
358     three dimensions. The cells were longer along the $z$ axis and the
359     space was divided into $N$ slabs along this axis (typically $N=20$).
360 skuang 3613 The top slab ($n=1$) was designated the ``hot'' slab, while the
361     central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
362 gezelter 3609 cases, simulations were first thermalized in canonical ensemble (NVT)
363     using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
364 gezelter 3600 microcanonical ensemble (NVE) before introducing any non-equilibrium
365     method.
366 skuang 3531
367 gezelter 3609 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
368 skuang 3531
369 gezelter 3609 In order to compare our new methodology with the original
370     M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
371 gezelter 3625 first performed simulations using the original technique. At fixed
372     intervals, kinetic energy or momentum exchange moves were performed
373     between the hot and the cold slabs. The interval between exchange
374     moves governs the effective momentum flux ($j_z(p_x)$) or energy flux
375     ($J_z$) between the two slabs so to vary this quantity, we performed
376     simulations with a variety of delay intervals between the swapping moves.
377 skuang 3531
378 gezelter 3625 For thermal conductivity measurements, the particle with smallest
379     speed in the hot slab and the one with largest speed in the cold slab
380     had their entire momentum vectors swapped. In the test cases run
381     here, all particles had the same chemical identity and mass, so this
382     move preserves both total linear momentum and total energy. It is
383     also possible to exchange energy by assuming an elastic collision
384     between the two particles which are exchanging energy.
385    
386     For shear stress simulations, the particle with the most negative
387     $p_x$ in the hot slab and the one with the most positive $p_x$ in the
388     cold slab exchanged only this component of their momentum vectors.
389    
390 gezelter 3609 \subsection{RNEMD with NIVS scaling}
391    
392     For each simulation utilizing the swapping method, a corresponding
393     NIVS-RNEMD simulation was carried out using a target momentum flux set
394 gezelter 3620 to produce the same flux experienced in the swapping simulation.
395 gezelter 3609
396 gezelter 3631 To test the temperature homogeneity, momentum and temperature
397     distributions (for all three dimensions) were accumulated for
398     molecules in each of the slabs. Transport coefficients were computed
399     using the temperature (and momentum) gradients across the $z$-axis as
400     well as the total momentum flux the system experienced during data
401     collection portion of the simulation.
402 gezelter 3609
403     \subsection{Shear viscosities}
404    
405     The momentum flux was calculated using the total non-physical momentum
406     transferred (${P_x}$) and the data collection time ($t$):
407 skuang 3534 \begin{equation}
408     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
409     \end{equation}
410 gezelter 3609 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
411     box. The factor of two in the denominator is present because physical
412 gezelter 3620 momentum transfer between the slabs occurs in two directions ($+z$ and
413     $-z$). The velocity gradient ${\langle \partial v_x /\partial z
414     \rangle}$ was obtained using linear regression of the mean $x$
415     component of the velocity, $\langle v_x \rangle$, in each of the bins.
416     For Lennard-Jones simulations, shear viscosities are reported in
417     reduced units (${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$).
418 skuang 3532
419 gezelter 3609 \subsection{Thermal Conductivities}
420 skuang 3534
421 gezelter 3620 The energy flux was calculated in a similar manner to the momentum
422     flux, using the total non-physical energy transferred (${E_{total}}$)
423     and the data collection time $t$:
424 skuang 3534 \begin{equation}
425     J_z = \frac{E_{total}}{2 t L_x L_y}
426     \end{equation}
427 gezelter 3609 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
428     obtained by a linear regression of the temperature profile. For
429     Lennard-Jones simulations, thermal conductivities are reported in
430     reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
431     k_B^{-1}\varepsilon^{-1/2}}$).
432 skuang 3534
433 gezelter 3609 \subsection{Interfacial Thermal Conductivities}
434 skuang 3563
435 gezelter 3620 For interfaces with a relatively low interfacial conductance, the bulk
436     regions on either side of an interface rapidly come to a state in
437     which the two phases have relatively homogeneous (but distinct)
438     temperatures. The interfacial thermal conductivity $G$ can therefore
439     be approximated as:
440 skuang 3573
441     \begin{equation}
442 gezelter 3631 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
443     \langle T_\mathrm{cold}\rangle \right)}
444 skuang 3573 \label{interfaceCalc}
445     \end{equation}
446 gezelter 3609 where ${E_{total}}$ is the imposed non-physical kinetic energy
447 gezelter 3631 transfer and ${\langle T_\mathrm{hot}\rangle}$ and ${\langle
448     T_\mathrm{cold}\rangle}$ are the average observed temperature of the
449     two separated phases. If the interfacial conductance is {\it not}
450     small, it would also be possible to compute the interfacial thermal
451 gezelter 3632 conductivity using this method by computing the change in the slope of
452 gezelter 3631 the thermal gradient ($\partial^2 \langle T \rangle /
453     \partial z^2$) at the interface.
454 skuang 3573
455 gezelter 3609 \section{Results}
456 skuang 3538
457 gezelter 3609 \subsection{Lennard-Jones Fluid}
458     2592 Lennard-Jones atoms were placed in an orthorhombic cell
459     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
460     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
461     direct comparison between our results and previous methods. These
462     simulations were carried out with a reduced timestep ${\tau^* =
463     4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
464     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
465 skuang 3617 conductivity calculations, simulations were run under reduced
466 gezelter 3609 temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
467 skuang 3617 ensemble. The simulations included $10^5$ steps of equilibration
468 gezelter 3609 without any momentum flux, $10^5$ steps of stablization with an
469     imposed momentum transfer to create a gradient, and $10^6$ steps of
470     data collection under RNEMD.
471    
472 gezelter 3611 \subsubsection*{Thermal Conductivity}
473    
474 gezelter 3609 Our thermal conductivity calculations show that the NIVS method agrees
475 skuang 3618 well with the swapping method. Five different swap intervals were
476 gezelter 3620 tested (Table \ref{LJ}). Similar thermal gradients were observed with
477 gezelter 3631 similar thermal flux under the two different methods (Fig.
478 gezelter 3625 \ref{thermalGrad}). Furthermore, the 1-d temperature profiles showed
479 gezelter 3631 no observable differences between the $x$, $y$ and $z$ axes (lower
480     panel of Fig. \ref{thermalGrad}), so even though we are using a
481     non-isotropic scaling method, none of the three directions are
482     experience disproportionate heating due to the velocity scaling.
483 gezelter 3609
484 skuang 3563 \begin{table*}
485 gezelter 3609 \begin{minipage}{\linewidth}
486     \begin{center}
487 skuang 3538
488 gezelter 3612 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
489     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
490     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
491     at various momentum fluxes. The original swapping method and
492     the velocity scaling method give similar results.
493     Uncertainties are indicated in parentheses.}
494 gezelter 3609
495 gezelter 3612 \begin{tabular}{|cc|cc|cc|}
496 gezelter 3609 \hline
497 gezelter 3612 \multicolumn{2}{|c}{Momentum Exchange} &
498     \multicolumn{2}{|c}{Swapping RNEMD} &
499 gezelter 3609 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
500     \hline
501 skuang 3646 \multirow{2}{*}{Swap Interval} & Equivalent $J_z^*$ or &
502     \multirow{2}{*}{$\lambda^*_{swap}$} &
503 gezelter 3612 \multirow{2}{*}{$\eta^*_{swap}$} &
504     \multirow{2}{*}{$\lambda^*_{scale}$} &
505     \multirow{2}{*}{$\eta^*_{scale}$} \\
506 skuang 3646 (timesteps) & $j_z^*(p_x)$ (reduced units) & & & & \\
507 gezelter 3609 \hline
508 skuang 3617 250 & 0.16 & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
509 gezelter 3612 500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
510     1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
511     2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
512 skuang 3617 2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
513 gezelter 3609 \hline
514     \end{tabular}
515 gezelter 3612 \label{LJ}
516 gezelter 3609 \end{center}
517     \end{minipage}
518 skuang 3563 \end{table*}
519    
520     \begin{figure}
521 gezelter 3612 \includegraphics[width=\linewidth]{thermalGrad}
522 gezelter 3625 \caption{The NIVS-RNEMD method creates similar temperature gradients
523     compared with the swapping method under a variety of imposed
524 skuang 3619 kinetic energy flux values. Furthermore, the implementation of
525     Non-Isotropic Velocity Scaling does not cause temperature
526 gezelter 3625 anisotropy to develop in thermal conductivity calculations.}
527 gezelter 3612 \label{thermalGrad}
528 skuang 3563 \end{figure}
529    
530 gezelter 3612 \subsubsection*{Velocity Distributions}
531    
532 gezelter 3631 To test the effects on the velocity distributions, we accumulated
533     velocities every 100 steps and produced distributions of both velocity
534     and speed in each of the slabs. From these distributions, we observed
535     that under high non-physical kinetic energy flux, the speed of
536     molecules in slabs where {\it swapping} occured could deviate from the
537 gezelter 3609 Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
538 gezelter 3620 and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
539     distributions deviate from an ideal distribution. In the ``hot'' slab,
540     the probability density is notched at low speeds and has a substantial
541 gezelter 3631 shoulder at higher speeds relative to the ideal distribution. In the
542     cold slab, the opposite notching and shouldering occurs. This
543     phenomenon is more obvious at high swapping rates.
544 skuang 3563
545 gezelter 3620 The peak of the velocity distribution is substantially flattened in
546     the hot slab, and is overly sharp (with truncated wings) in the cold
547     slab. This problem is rooted in the mechanism of the swapping method.
548     Continually depleting low (high) speed particles in the high (low)
549 gezelter 3631 temperature slab is not complemented by diffusion of low (high) speed
550     particles from neighboring slabs unless the swapping rate is
551 gezelter 3620 sufficiently small. Simutaneously, surplus low speed particles in the
552 gezelter 3631 cold slab do not have sufficient time to diffuse to neighboring slabs.
553     Since the thermal exchange rate must reach a minimum level to produce
554     an observable thermal gradient, the swapping-method RNEMD has a
555     relatively narrow choice of exchange times that can be utilized.
556 skuang 3578
557 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
558 gezelter 3631 Maxwell-Boltzmann curve (Fig. \ref{thermalHist}). The reason for this
559     is simple; upon velocity scaling, a Gaussian distribution remains
560 gezelter 3609 Gaussian. Although a single scaling move is non-isotropic in three
561     dimensions, our criteria for choosing a set of scaling coefficients
562     helps maintain the distributions as close to isotropic as possible.
563 gezelter 3631 Consequently, NIVS-RNEMD is able to impose a non-physical thermal flux
564     without large perturbations to the velocity distributions in the two
565     slabs.
566 gezelter 3609
567 skuang 3568 \begin{figure}
568 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
569 gezelter 3629 \caption{Velocity and speed distributions that develop under the
570     swapping and NIVS-RNEMD methods at high flux. The distributions for
571     the hot bins (upper panels) and cold bins (lower panels) were
572     obtained from Lennard-Jones simulations with $\langle T^* \rangle =
573 skuang 3630 4.5$ with a flux of $J_z^* \sim 5$ (equivalent to a swapping interval
574 gezelter 3629 of 10 time steps). This is a relatively large flux which shows the
575     non-thermal distributions that develop under the swapping method.
576     NIVS does a better job of producing near-thermal distributions in
577     the bins.}
578 skuang 3589 \label{thermalHist}
579 skuang 3568 \end{figure}
580    
581 gezelter 3611
582     \subsubsection*{Shear Viscosity}
583 gezelter 3620 Our calculations (Table \ref{LJ}) show that velocity-scaling RNEMD
584 gezelter 3631 predicted similar values for shear viscosities to the swapping RNEMD
585     method. The average molecular momentum gradients of these samples are
586     shown in the upper two panels of Fig. \ref{shear}.
587 gezelter 3611
588     \begin{figure}
589     \includegraphics[width=\linewidth]{shear}
590     \caption{Average momentum gradients in shear viscosity simulations,
591 gezelter 3625 using ``swapping'' method (top panel) and NIVS-RNEMD method
592     (middle panel). NIVS-RNEMD produces a thermal anisotropy artifact
593     in the hot and cold bins when used for shear viscosity. This
594     artifact does not appear in thermal conductivity calculations.}
595 gezelter 3611 \label{shear}
596     \end{figure}
597    
598 gezelter 3620 Observations of the three one-dimensional temperatures in each of the
599     slabs shows that NIVS-RNEMD does produce some thermal anisotropy,
600 skuang 3645 particularly in the hot and cold slabs. Note that these temperature
601     measurements have been taken into account of the kinetic energy
602     contributed by the slab field velocity. However, this contribution has
603     only a comparable order of magnitude to the errors of data, and does
604     not significantly affect our observation. The lower panel of Fig.
605 gezelter 3631 \ref{shear} indicates that with a relatively large imposed momentum
606     flux, $j_z(p_x)$, the $x$ direction approaches a different temperature
607     from the $y$ and $z$ directions in both the hot and cold bins. This
608     is an artifact of the scaling constraints. After a momentum gradient
609     has been established, $P_c^x$ is always less than zero. Consequently,
610     the scaling factor $x$ is always greater than one in order to satisfy
611     the constraints. This will continually increase the kinetic energy in
612     $x$-dimension, $K_c^x$. If there is not enough time for the kinetic
613     energy to exchange among different directions and different slabs, the
614     system will exhibit the observed thermal anisotropy in the hot and
615     cold bins.
616 gezelter 3611
617     Although results between scaling and swapping methods are comparable,
618 gezelter 3620 the inherent temperature anisotropy does make NIVS-RNEMD method less
619     attractive than swapping RNEMD for shear viscosity calculations. We
620 gezelter 3631 note that this problem appears only when a large {\it linear} momentum
621     flux is applied, and does not appear in {\it thermal} flux
622     calculations.
623 gezelter 3611
624 gezelter 3609 \subsection{Bulk SPC/E water}
625    
626     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
627     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
628     the original swapping RNEMD method. Bedrov {\it et
629 gezelter 3631 al.}\cite{Bedrov:2000} argued that exchange of the molecular
630     center-of-mass velocities instead of single atom velocities conserves
631     the total kinetic energy and linear momentum. This principle is also
632     adopted in our simulations. Scaling was applied to the center-of-mass
633     velocities of SPC/E water molecules.
634 skuang 3563
635 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
636     molecules were first equilibrated under ambient pressure and
637     temperature conditions using the isobaric-isothermal (NPT)
638     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
639     average volume observed in the NPT simulations, and this was followed
640     by equilibration, first in the canonical (NVT) ensemble, followed by a
641 gezelter 3620 100~ps period under constant-NVE conditions without any momentum flux.
642     Another 100~ps was allowed to stabilize the system with an imposed
643     momentum transfer to create a gradient, and 1~ns was allotted for data
644 skuang 3645 collection under RNEMD. Total system energy is recorded to ensure that
645     it is not drifted noticeably without a thermostat although
646     electrostatic interactions are involved.
647 gezelter 3609
648 gezelter 3620 In our simulations, the established temperature gradients were similar
649     to the previous work. Our simulation results at 318K are in good
650 skuang 3619 agreement with those from Bedrov {\it et al.} (Table
651 skuang 3615 \ref{spceThermal}). And both methods yield values in reasonable
652 gezelter 3620 agreement with experimental values.
653 gezelter 3609
654 skuang 3570 \begin{table*}
655 gezelter 3609 \begin{minipage}{\linewidth}
656     \begin{center}
657    
658     \caption{Thermal conductivity of SPC/E water under various
659     imposed thermal gradients. Uncertainties are indicated in
660     parentheses.}
661    
662 skuang 3615 \begin{tabular}{|c|c|ccc|}
663 gezelter 3609 \hline
664 skuang 3615 \multirow{2}{*}{$\langle T\rangle$(K)} &
665     \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
666     \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
667     \mathrm{K}^{-1})$} \\
668     & & This work & Previous simulations\cite{Bedrov:2000} &
669 gezelter 3609 Experiment\cite{WagnerKruse}\\
670     \hline
671 skuang 3615 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
672     \multirow{3}{*}{0.61}\\
673     & 0.81 & 0.770(0.008) & & \\
674     & 1.54 & 0.813(0.007) & & \\
675 gezelter 3609 \hline
676 skuang 3615 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
677     \multirow{2}{*}{0.64}\\
678     & 1.59 & 0.778(0.019) & 0.730 & \\
679     \hline
680 gezelter 3609 \end{tabular}
681     \label{spceThermal}
682     \end{center}
683     \end{minipage}
684     \end{table*}
685 skuang 3570
686 gezelter 3609 \subsection{Crystalline Gold}
687 skuang 3570
688 gezelter 3609 To see how the method performed in a solid, we calculated thermal
689     conductivities using two atomistic models for gold. Several different
690     potential models have been developed that reasonably describe
691     interactions in transition metals. In particular, the Embedded Atom
692 gezelter 3620 Model (EAM)~\cite{PhysRevB.33.7983} and Sutton-Chen (SC)~\cite{Chen90}
693     potential have been used to study a wide range of phenomena in both
694     bulk materials and
695 gezelter 3609 nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
696     Both potentials are based on a model of a metal which treats the
697     nuclei and core electrons as pseudo-atoms embedded in the electron
698     density due to the valence electrons on all of the other atoms in the
699 gezelter 3620 system. The SC potential has a simple form that closely resembles the
700     Lennard Jones potential,
701 gezelter 3609 \begin{equation}
702     \label{eq:SCP1}
703     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] ,
704     \end{equation}
705     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
706     \begin{equation}
707     \label{eq:SCP2}
708     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}}.
709     \end{equation}
710     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
711     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
712     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
713     the interactions between the valence electrons and the cores of the
714     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
715     scale, $c_i$ scales the attractive portion of the potential relative
716     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
717     that assures a dimensionless form for $\rho$. These parameters are
718     tuned to various experimental properties such as the density, cohesive
719     energy, and elastic moduli for FCC transition metals. The quantum
720 gezelter 3620 Sutton-Chen (QSC) formulation matches these properties while including
721     zero-point quantum corrections for different transition
722     metals.\cite{PhysRevB.59.3527} The EAM functional forms differ
723     slightly from SC but the overall method is very similar.
724 skuang 3570
725 gezelter 3620 In this work, we have utilized both the EAM and the QSC potentials to
726     test the behavior of scaling RNEMD.
727 skuang 3570
728 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
729 skuang 3613 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
730     both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
731 gezelter 3609 pre-equilibration at a target pressure of 1 atm. When equilibrated
732     under NPT conditions, our simulation box expanded by approximately 1\%
733 skuang 3613 in volume. Following adjustment of the box volume, equilibrations in
734     both the canonical and microcanonical ensembles were carried out. With
735     the simulation cell divided evenly into 10 slabs, different thermal
736 gezelter 3631 gradients were established by applying a set of target thermal fluxes.
737 skuang 3570
738 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
739     \ref{AuThermal}. In these calculations, the end and middle slabs were
740 gezelter 3631 excluded from the thermal gradient linear regession. EAM predicts
741     slightly larger thermal conductivities than QSC. However, both values
742     are smaller than experimental value by a factor of more than 200. This
743 gezelter 3620 behavior has been observed previously by Richardson and Clancy, and
744     has been attributed to the lack of electronic contribution in these
745     force fields.\cite{Clancy:1992} It should be noted that the density of
746     the metal being simulated has an effect on thermal conductance. With
747     an expanded lattice, lower thermal conductance is expected (and
748     observed). We also observed a decrease in thermal conductance at
749     higher temperatures, a trend that agrees with experimental
750     measurements.\cite{AshcroftMermin}
751 skuang 3570
752 gezelter 3609 \begin{table*}
753     \begin{minipage}{\linewidth}
754     \begin{center}
755    
756     \caption{Calculated thermal conductivity of crystalline gold
757     using two related force fields. Calculations were done at both
758     experimental and equilibrated densities and at a range of
759 skuang 3617 temperatures and thermal flux rates. Uncertainties are
760     indicated in parentheses. Richardson {\it et
761 gezelter 3621 al.}\cite{Clancy:1992} give an estimate of 1.74 $\mathrm{W
762     m}^{-1}\mathrm{K}^{-1}$ for EAM gold
763     at a density of 19.263 g / cm$^3$.}
764 gezelter 3609
765     \begin{tabular}{|c|c|c|cc|}
766     \hline
767     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
768     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
769     \hline
770 gezelter 3621 \multirow{7}{*}{QSC} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
771 gezelter 3609 & & & 2.86 & 1.08(0.05)\\
772     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
773     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
774     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
775     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
776     & & & 4.84 & 0.92(0.05)\\
777     \hline
778 gezelter 3621 \multirow{8}{*}{EAM} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
779 gezelter 3609 & & & 2.06 & 1.37(0.04)\\
780     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
781     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
782     & & & 2.04 & 1.41(0.07)\\
783     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
784     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
785     & & & 4.14 & 1.08(0.05)\\
786     \hline
787     \end{tabular}
788     \label{AuThermal}
789     \end{center}
790     \end{minipage}
791 skuang 3580 \end{table*}
792    
793 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
794     The most attractive aspect of the scaling approach for RNEMD is the
795     ability to use the method in non-homogeneous systems, where molecules
796     of different identities are segregated in different slabs. To test
797     this application, we simulated a Gold (111) / water interface. To
798     construct the interface, a box containing a lattice of 1188 Au atoms
799 skuang 3619 (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
800 gezelter 3609 relax under ambient temperature and pressure. A separate (but
801     identically sized) box of SPC/E water was also equilibrated at ambient
802     conditions. The two boxes were combined by removing all water
803 gezelter 3631 molecules within 3 \AA~ radius of any gold atom. The final
804 gezelter 3609 configuration contained 1862 SPC/E water molecules.
805 skuang 3580
806 gezelter 3620 The Spohr potential was adopted in depicting the interaction between
807     metal atoms and water molecules.\cite{ISI:000167766600035} A similar
808     protocol of equilibration to our water simulations was followed. We
809     observed that the two phases developed large temperature differences
810     even under a relatively low thermal flux.
811 gezelter 3609
812 gezelter 3625 The low interfacial conductance is probably due to an acoustic
813     impedance mismatch between the solid and the liquid
814     phase.\cite{Cahill:793,RevModPhys.61.605} Experiments on the thermal
815     conductivity of gold nanoparticles and nanorods in solvent and in
816     glass cages have predicted values for $G$ between 100 and 350
817 gezelter 3632 (MW/m$^2$/K), two orders of magnitude larger than the value reported
818     here. The experiments typically have multiple surfaces that have been
819     protected by ionic surfactants, either
820     citrate\cite{Wilson:2002uq,plech:195423} or cetyltrimethylammonium
821     bromide (CTAB), or which are in direct contact with various glassy
822     solids. In these cases, the acoustic impedance mismatch would be
823     substantially reduced, leading to much higher interfacial
824     conductances. It is also possible, however, that the lack of
825     electronic effects that gives rise to the low thermal conductivity of
826     EAM gold is also causing a low reading for this particular interface.
827 gezelter 3625
828 skuang 3633 Under this low thermal conductance, both gold and water phases have
829 gezelter 3631 sufficient time to eliminate local temperature differences (Fig.
830 gezelter 3632 \ref{interface}). With flat thermal profiles within each phase, it is
831 gezelter 3631 valid to assume that the temperature difference between gold and water
832 skuang 3633 surfaces would be approximately the same as the difference between the
833     gold and water bulk regions. This assumption enables convenient
834     calculation of $G$ using Eq. \ref{interfaceCalc}.
835 skuang 3573
836 skuang 3571 \begin{figure}
837 skuang 3595 \includegraphics[width=\linewidth]{interface}
838 gezelter 3625 \caption{Temperature profiles of the Gold / Water interface at four
839     different values for the thermal flux. Temperatures for slabs
840     either in the gold or in the water show no significant differences,
841     although there is a large discontinuity between the materials
842     because of the relatively low interfacial thermal conductivity.}
843 skuang 3595 \label{interface}
844 skuang 3571 \end{figure}
845    
846 skuang 3572 \begin{table*}
847 gezelter 3612 \begin{minipage}{\linewidth}
848     \begin{center}
849    
850     \caption{Computed interfacial thermal conductivity ($G$) values
851     for the Au(111) / water interface at ${\langle T\rangle \sim}$
852     300K using a range of energy fluxes. Uncertainties are
853     indicated in parentheses. }
854    
855 gezelter 3616 \begin{tabular}{|cccc| }
856 gezelter 3612 \hline
857     $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
858     T_{water} \rangle$ (K) & $G$
859     (MW/m$^2$/K)\\
860     \hline
861     98.0 & 355.2 & 295.8 & 1.65(0.21) \\
862     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
863     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
864     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
865     \hline
866     \end{tabular}
867     \label{interfaceRes}
868     \end{center}
869     \end{minipage}
870 skuang 3572 \end{table*}
871    
872 skuang 3576
873 skuang 3574 \section{Conclusions}
874 gezelter 3631
875     Our simulations demonstrate that validity of non-isotropic velocity
876     scaling (NIVS) in RNEMD calculations of thermal conductivity in atomic
877     and molecular liquids and solids. NIVS-RNEMD improves the problematic
878     velocity distributions which can develop in other RNEMD methods.
879     Furthermore, it provides a means for carrying out non-physical thermal
880 skuang 3581 transfer between different species of molecules, and thus extends its
881 gezelter 3631 applicability to interfacial systems. Our calculation of the gold /
882     water interfacial thermal conductivity demonstrates this advantage
883     over previous RNEMD methods. NIVS-RNEMD also has limited applications
884     for shear viscosity calculations, but has the potential to cause
885     temperature anisotropy under high momentum fluxes. Further work will
886     be necessary to eliminate the one-dimensional heating if shear
887     viscosities are required.
888 skuang 3572
889 gezelter 3524 \section{Acknowledgments}
890 gezelter 3624 The authors would like to thank Craig Tenney and Ed Maginn for many
891     helpful discussions. Support for this project was provided by the
892     National Science Foundation under grant CHE-0848243. Computational
893     time was provided by the Center for Research Computing (CRC) at the
894     University of Notre Dame.
895     \newpage
896 gezelter 3524
897     \bibliography{nivsRnemd}
898 gezelter 3583
899 gezelter 3524 \end{doublespace}
900     \end{document}
901