ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3632
Committed: Thu Aug 12 15:39:32 2010 UTC (14 years, 1 month ago) by gezelter
Content type: application/x-tex
File size: 42163 byte(s)
Log Message:
more edits

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