ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
(Generate patch)

Comparing trunk/nivsRnemd/nivsRnemd.tex (file contents):
Revision 3527 by skuang, Mon Sep 21 20:54:06 2009 UTC vs.
Revision 3630 by skuang, Thu Aug 12 14:48:03 2010 UTC

# Line 6 | Line 6
6   \usepackage{caption}
7   %\usepackage{tabularx}
8   \usepackage{graphicx}
9 + \usepackage{multirow}
10   %\usepackage{booktabs}
11   %\usepackage{bibentry}
12   %\usepackage{mathrsfs}
13 < \usepackage[ref]{overcite}
13 > %\usepackage[ref]{overcite}
14 > \usepackage[square, comma, sort&compress]{natbib}
15 > \usepackage{url}
16   \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17   \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18   9.0in \textwidth 6.5in \brokenpenalty=10000
# Line 19 | Line 22
22   \setlength{\abovecaptionskip}{20 pt}
23   \setlength{\belowcaptionskip}{30 pt}
24  
25 < \renewcommand\citemid{\ } % no comma in optional referenc note
25 > %\renewcommand\citemid{\ } % no comma in optional referenc note
26 > \bibpunct{[}{]}{,}{s}{}{;}
27 > \bibliographystyle{aip}
28  
29   \begin{document}
30  
# Line 38 | Line 43 | Notre Dame, Indiana 46556}
43   \begin{doublespace}
44  
45   \begin{abstract}
46 <
46 >  We present a new method for introducing stable non-equilibrium
47 >  velocity and temperature gradients in molecular dynamics simulations
48 >  of heterogeneous systems.  This method extends earlier Reverse
49 >  Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
50 >  momentum exchange swapping moves. The standard swapping moves can
51 >  create non-thermal velocity distributions and are difficult to use
52 >  for interfacial calculations.  By using non-isotropic velocity
53 >  scaling (NIVS) on the molecules in specific regions of a system, it
54 >  is possible to impose momentum or thermal flux between regions of a
55 >  simulation while conserving the linear momentum and total energy of
56 >  the system.  To test the methods, we have computed the thermal
57 >  conductivity of model liquid and solid systems as well as the
58 >  interfacial thermal conductivity of a metal-water interface.  We
59 >  find that the NIVS-RNEMD improves the problematic velocity
60 >  distributions that develop in other RNEMD methods.
61   \end{abstract}
62  
63   \newpage
# Line 49 | Line 68 | Notre Dame, Indiana 46556}
68   %                          BODY OF TEXT
69   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70  
52
53
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 < the simulation cell.\cite{MullerPlathe:1997xw,Muller-Plathe:1999ek} The
77 < artificial flux is typically created by periodically "swapping" either
78 < the entire momentum vector $\vec{p}$ or single components of this
79 < vector ($p_x$) between molecules in each of the two slabs.  If the two
80 < slabs are separated along the z coordinate, the imposed flux is either
81 < directional ($J_z(p_x)$) or isotropic ($J_z$), and the response of a
82 < simulated system to the imposed momentum flux will typically be a
83 < velocity or thermal gradient.  The transport coefficients (shear
84 < viscosity and thermal conductivity) are easily obtained by assuming
85 < linear response of the system,
76 > the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
77 > 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   \begin{eqnarray}
87 < J_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
88 < J & = & \lambda \frac{\partial T}{\partial z}
87 > j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
88 > J_z & = & \lambda \frac{\partial T}{\partial z}
89   \end{eqnarray}
90 < RNEMD been widely used to provide computational estimates of thermal
91 < conductivities and shear viscosities in a wide range of materials,
92 < from liquid copper to monatomic liquids to molecular fluids
93 < (e.g. ionic liquids).
90 > 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  
96 < RNEMD is preferable in many ways to the forward NEMD methods because
97 < it imposes what is typically difficult to measure (a flux or stress)
98 < and it is typically much easier to compute momentum gradients or
99 < strains (the response).  For similar reasons, RNEMD is also preferable
100 < to slowly-converging equilibrium methods for measuring thermal
101 < conductivity and shear viscosity (using Green-Kubo relations or the
102 < Helfand moment approach of Viscardy {\it et
96 > \begin{figure}
97 > \includegraphics[width=\linewidth]{thermalDemo}
98 > \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 > \label{thermalDemo}
105 > \end{figure}
106 >
107 > RNEMD is preferable in many ways to the forward NEMD
108 > 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 > stress) and it is typically much easier to compute the response
111 > (momentum gradients or strains).  For similar reasons, RNEMD is also
112 > preferable to slowly-converging equilibrium methods for measuring
113 > thermal conductivity and shear viscosity (using Green-Kubo
114 > relations\cite{daivis:541,mondello:9327} or the Helfand moment
115 > approach of Viscardy {\it et
116    al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
117   computing difficult to measure quantities.
118  
# Line 91 | Line 122 | Recently, Tenney and Maginn have discovered some probl
122   typically samples from the same manifold of states in the
123   microcanonical ensemble.
124  
125 < Recently, Tenney and Maginn have discovered some problems with the
126 < original RNEMD swap technique.  Notably, large momentum fluxes
127 < (equivalent to frequent momentum swaps between the slabs) can result
128 < in "notched", "peaked" and generally non-thermal momentum
129 < distributions in the two slabs, as well as non-linear thermal and
130 < velocity distributions along the direction of the imposed flux ($z$).
131 < Tenney and Maginn obtained reasonable limits on imposed flux and
132 < self-adjusting metrics for retaining the usability of the method.
125 > Recently, Tenney and Maginn\cite{Maginn:2010} have discovered some
126 > problems with the original RNEMD swap technique.  Notably, large
127 > momentum fluxes (equivalent to frequent momentum swaps between the
128 > slabs) can result in ``notched'', ``peaked'' and generally non-thermal
129 > momentum distributions in the two slabs, as well as non-linear thermal
130 > and velocity distributions along the direction of the imposed flux
131 > ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
132 > and proposed self-adjusting metrics for retaining the usability of the
133 > method.
134  
135   In this paper, we develop and test a method for non-isotropic velocity
136 < scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
136 > scaling (NIVS) which retains the desirable features of RNEMD
137   (conservation of linear momentum and total energy, compatibility with
138   periodic boundary conditions) while establishing true thermal
139 < distributions in each of the two slabs.  In the next section, we
140 < develop the method for determining the scaling constraints.  We then
141 < test the method on both single component, multi-component, and
142 < non-isotropic mixtures and show that it is capable of providing
139 > distributions in each of the two slabs. In the next section, we
140 > present the method for determining the scaling constraints.  We then
141 > test the method on both liquids and solids as well as a non-isotropic
142 > liquid-solid interface and show that it is capable of providing
143   reasonable estimates of the thermal conductivity and shear viscosity
144 < in these cases.
144 > in all of these cases.
145  
146   \section{Methodology}
147 < We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
148 < system is partitioned into a series of thin slabs along a particular
147 > We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
148 > periodic system is partitioned into a series of thin slabs along one
149   axis ($z$).  One of the slabs at the end of the periodic box is
150   designated the ``hot'' slab, while the slab in the center of the box
151   is designated the ``cold'' slab.  The artificial momentum flux will be
# Line 121 | Line 153 | moves.  For molecules $\{i\}$ located within the hot s
153   hot slab.
154  
155   Rather than using momentum swaps, we use a series of velocity scaling
156 < moves.  For molecules $\{i\}$ located within the hot slab,
156 > moves.  For molecules $\{i\}$  located within the cold slab,
157   \begin{equation}
158 < \vec{v}_i \leftarrow \left( \begin{array}{c}
159 < x \\
160 < y \\
161 < z \\
158 > \vec{v}_i \leftarrow \left( \begin{array}{ccc}
159 > x & 0 & 0 \\
160 > 0 & y & 0 \\
161 > 0 & 0 & z \\
162   \end{array} \right) \cdot \vec{v}_i
163   \end{equation}
164 < where ${x, y, z}$ are a set of 3 scaling variables for each of the
165 < three directions in the system.  Likewise, the molecules $\{j\}$
166 < located in the cold bin will see a concomitant scaling of velocities,
164 > where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
165 > of the three directions in the system.  Likewise, the molecules
166 > $\{j\}$ located in the hot slab will see a concomitant scaling of
167 > velocities,
168   \begin{equation}
169 < \vec{v}_j \leftarrow \left( \begin{array}{c}
170 < x^\prime \\
171 < y^\prime \\
172 < z^\prime \\
169 > \vec{v}_j \leftarrow \left( \begin{array}{ccc}
170 > x^\prime & 0 & 0 \\
171 > 0 & y^\prime & 0 \\
172 > 0 & 0 & z^\prime \\
173   \end{array} \right) \cdot \vec{v}_j
174   \end{equation}
175  
176   Conservation of linear momentum in each of the three directions
177 < ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
177 > ($\alpha = x,y,z$) ties the values of the hot and cold scaling
178   parameters together:
179   \begin{equation}
180 < P_h^\alpha + P_c^\alpha = \alpha P_h^\alpha + \alpha^\prime P_c^\alpha
180 > P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
181   \end{equation}
182   where
183 < \begin{equation}
184 < \begin{array}{rcl}
185 < P_h^\alpha & = & \sum_{i = 1}^{N_h} m_i \left[\vec{v}_i\right]_\alpha \\
153 < P_c^\alpha & = & \sum_{j = 1}^{N_c} m_j \left[\vec{v}_j\right]_\alpha
154 < \\
155 < \end{array}
183 > \begin{eqnarray}
184 > P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i v_{i\alpha} \\
185 > P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j v_{j\alpha}
186   \label{eq:momentumdef}
187 < \end{equation}
188 < Therefore, for each of the three directions, the cold scaling
189 < parameters are a simple function of the hot scaling parameters and
190 < the instantaneous linear momentum in each of the two slabs.
187 > \end{eqnarray}
188 > Therefore, for each of the three directions, the hot scaling
189 > parameters are a simple function of the cold scaling parameters and
190 > the instantaneous linear momenta in each of the two slabs.
191   \begin{equation}
192 < \alpha^\prime = 1 + (1 - \alpha) \frac{P_h^\alpha}{P_c^\alpha}.
192 > \alpha^\prime = 1 + (1 - \alpha) p_\alpha
193   \label{eq:hotcoldscaling}
194   \end{equation}
195 + where
196 + \begin{equation}
197 + p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
198 + \end{equation}
199 + for convenience.
200  
201   Conservation of total energy also places constraints on the scaling:
202   \begin{equation}
203 < \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
204 < \alpha^2 K_h^\alpha + \left(\alpha^\prime\right)^2 K_c^\alpha.
203 > \sum_{\alpha = x,y,z} \left\{ K_h^\alpha + K_c^\alpha \right\} = \sum_{\alpha = x,y,z}
204 > \left\{ \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha \right\}
205   \end{equation}
206 < where the kinetic energies, $K_h^\alpha$ and $K_c^\alpha$, are computed
207 < for each of the three directions in a similar manner to the linear momenta
208 < (Eq. \ref{eq:momentumdef}).  Substituting in the expressions for the
209 < cold scaling parameters ($\alpha^\prime$) from
210 < Eq. (\ref{eq:hotcoldscaling}), we obtain the {\it constraint ellipsoid}:
206 > where the translational kinetic energies, $K_h^\alpha$ and
207 > $K_c^\alpha$, are computed for each of the three directions in a
208 > similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
209 > Substituting in the expressions for the hot scaling parameters
210 > ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
211 > {\it constraint ellipsoid}:
212   \begin{equation}
213 < C_x (1+x)^2 + C_y (1+y)^2 + C_z (1+z)^2 = 0,
213 > \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
214 >  c_\alpha \right) = 0
215   \label{eq:constraintEllipsoid}
216   \end{equation}
217   where the constants are obtained from the instantaneous values of the
218   linear momenta and kinetic energies for the hot and cold slabs,
219 < \begin{equation}
220 < C_\alpha = \left( K_h^\alpha + K_c^\alpha
221 <  \left(\frac{P_h^\alpha}{P_c^\alpha}\right)^2\right)
219 > \begin{eqnarray}
220 > a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
221 >  \left(p_\alpha\right)^2\right) \\
222 > b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
223 > c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
224   \label{eq:constraintEllipsoidConsts}
225 < \end{equation}
226 < This ellipsoid defines the set of hot slab scaling parameters which can be
227 < applied while preserving both linear momentum in all three directions
228 < as well as total energy.
225 > \end{eqnarray}
226 > This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
227 > cold slab scaling parameters which, when applied, preserve the linear
228 > momentum of the system in all three directions as well as total
229 > kinetic energy.
230  
231 < The goal of using velocity scaling variables is to transfer linear
232 < momentum or kinetic energy from the cold slab to the hot slab.  If the
233 < hot and cold slabs are separated along the z-axis, the energy flux is
234 < given simply by the increase in kinetic energy of the hot bin:
231 > The goal of using these velocity scaling variables is to transfer
232 > kinetic energy from the cold slab to the hot slab.  If the hot and
233 > cold slabs are separated along the z-axis, the energy flux is given
234 > simply by the decrease in kinetic energy of the cold bin:
235   \begin{equation}
236 < (x^2-1) K_h^x  + (y^2-1) K_h^y + (z^2-1) K_h^z = J_z
236 > (1-x^2) K_c^x  + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
237   \end{equation}
238   The expression for the energy flux can be re-written as another
239   ellipsoid centered on $(x,y,z) = 0$:
240   \begin{equation}
241 < x^2 K_h^x + y^2 K_h^y + z^2 K_h^z = (J_z + K_h^x + K_h^y + K_h^z)
241 > \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
242 > K_c^\alpha -J_z \Delta t
243   \label{eq:fluxEllipsoid}
244   \end{equation}
245 < The spatial extent of the {\it flux ellipsoid} is governed both by a
246 < targetted value, $J_z$ as well as the instantaneous values of the
247 < kinetic energy components in the hot bin.
245 > The spatial extent of the {\it thermal flux ellipsoid} is governed
246 > both by the target flux, $J_z$ as well as the instantaneous values of
247 > the kinetic energy components in the cold bin.
248  
249   To satisfy an energetic flux as well as the conservation constraints,
250 < it is sufficient to determine the points ${x,y,z}$ which lie on both
250 > we must determine the points ${x,y,z}$ that lie on both the constraint
251 > ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
252 > (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
253 > ellipsoids in 3-dimensional space.
254 >
255 > \begin{figure}
256 > \includegraphics[width=\linewidth]{ellipsoids}
257 > \caption{Velocity scaling coefficients which maintain both constant
258 >  energy and constant linear momentum of the system lie on the surface
259 >  of the {\it constraint ellipsoid} while points which generate the
260 >  target momentum flux lie on the surface of the {\it flux ellipsoid}.
261 >  The velocity distributions in the cold bin are scaled by only those
262 >  points which lie on both ellipsoids.}
263 > \label{ellipsoids}
264 > \end{figure}
265 >
266 > Since ellipsoids can be expressed as polynomials up to second order in
267 > each of the three coordinates, finding the the intersection points of
268 > two ellipsoids is isomorphic to finding the roots a polynomial of
269 > degree 16.  There are a number of polynomial root-finding methods in
270 > the literature,\cite{Hoffman:2001sf,384119} but numerically finding
271 > the roots of high-degree polynomials is generally an ill-conditioned
272 > problem.\cite{Hoffman:2001sf} One simplification is to maintain
273 > velocity scalings that are {\it as isotropic as possible}.  To do
274 > this, we impose $x=y$, and treat both the constraint and flux
275 > ellipsoids as 2-dimensional ellipses.  In reduced dimensionality, the
276 > intersecting-ellipse problem reduces to finding the roots of
277 > polynomials of degree 4.
278 >
279 > Depending on the target flux and current velocity distributions, the
280 > ellipsoids can have between 0 and 4 intersection points.  If there are
281 > no intersection points, it is not possible to satisfy the constraints
282 > while performing a non-equilibrium scaling move, and no change is made
283 > to the dynamics.  
284 >
285 > With multiple intersection points, any of the scaling points will
286 > conserve the linear momentum and kinetic energy of the system and will
287 > generate the correct target flux.  Although this method is inherently
288 > non-isotropic, the goal is still to maintain the system as close to an
289 > isotropic fluid as possible.  With this in mind, we would like the
290 > kinetic energies in the three different directions could become as
291 > close as each other as possible after each scaling.  Simultaneously,
292 > one would also like each scaling as gentle as possible, i.e. ${x,y,z
293 >  \rightarrow 1}$, in order to avoid large perturbation to the system.
294 > To do this, we pick the intersection point which maintains the three
295 > scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
296 > ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
297 >
298 > After the valid scaling parameters are arrived at by solving geometric
299 > intersection problems in $x, y, z$ space in order to obtain cold slab
300 > scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
301 > determine the conjugate hot slab scaling variables.
302 >
303 > \subsection{Introducing shear stress via velocity scaling}
304 > It is also possible to use this method to magnify the random
305 > fluctuations of the average momentum in each of the bins to induce a
306 > momentum flux.  Doing this repeatedly will create a shear stress on
307 > the system which will respond with an easily-measured strain.  The
308 > momentum flux (say along the $x$-direction) may be defined as:
309 > \begin{equation}
310 > (1-x) P_c^x = j_z(p_x)\Delta t
311 > \label{eq:fluxPlane}
312 > \end{equation}
313 > This {\it momentum flux plane} is perpendicular to the $x$-axis, with
314 > its position governed both by a target value, $j_z(p_x)$ as well as
315 > the instantaneous value of the momentum along the $x$-direction.
316 >
317 > In order to satisfy a momentum flux as well as the conservation
318 > constraints, we must determine the points ${x,y,z}$ which lie on both
319   the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
320 < flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
321 < the two ellipsoids in 3-space.
320 > flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
321 > ellipsoid and a plane in 3-dimensional space.
322  
323 + In the case of momentum flux transfer, we also impose another
324 + constraint to set the kinetic energy transfer as zero. In other
325 + words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$.  With
326 + one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
327 + set of quartic equations to the above kinetic energy transfer problem.
328  
329 < One may also define momentum flux (say along the x-direction) as:
329 > \section{Computational Details}
330 >
331 > We have implemented this methodology in our molecular dynamics code,
332 > OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
333 > with a variable frequency after the molecular dynamics (MD) steps.  We
334 > have tested the method in a variety of different systems, including:
335 > homogeneous fluids (Lennard-Jones and SPC/E water), crystalline
336 > solids, using both the embedded atom method
337 > (EAM)~\cite{PhysRevB.33.7983} and quantum Sutton-Chen
338 > (QSC)~\cite{PhysRevB.59.3527} models for Gold, and heterogeneous
339 > interfaces (QSC gold - SPC/E water). The last of these systems would
340 > have been difficult to study using previous RNEMD methods, but the
341 > current method can easily provide estimates of the interfacial thermal
342 > conductivity ($G$).
343 >
344 > \subsection{Simulation Cells}
345 >
346 > In each of the systems studied, the dynamics was carried out in a
347 > rectangular simulation cell using periodic boundary conditions in all
348 > three dimensions.  The cells were longer along the $z$ axis and the
349 > space was divided into $N$ slabs along this axis (typically $N=20$).
350 > The top slab ($n=1$) was designated the ``hot'' slab, while the
351 > central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
352 > cases, simulations were first thermalized in canonical ensemble (NVT)
353 > using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
354 > microcanonical ensemble (NVE) before introducing any non-equilibrium
355 > method.
356 >
357 > \subsection{RNEMD with M\"{u}ller-Plathe swaps}
358 >
359 > In order to compare our new methodology with the original
360 > M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
361 > first performed simulations using the original technique. At fixed
362 > intervals, kinetic energy or momentum exchange moves were performed
363 > between the hot and the cold slabs.  The interval between exchange
364 > moves governs the effective momentum flux ($j_z(p_x)$) or energy flux
365 > ($J_z$) between the two slabs so to vary this quantity, we performed
366 > simulations with a variety of delay intervals between the swapping moves.
367 >
368 > For thermal conductivity measurements, the particle with smallest
369 > speed in the hot slab and the one with largest speed in the cold slab
370 > had their entire momentum vectors swapped.  In the test cases run
371 > here, all particles had the same chemical identity and mass, so this
372 > move preserves both total linear momentum and total energy.  It is
373 > also possible to exchange energy by assuming an elastic collision
374 > between the two particles which are exchanging energy.
375 >
376 > For shear stress simulations, the particle with the most negative
377 > $p_x$ in the hot slab and the one with the most positive $p_x$ in the
378 > cold slab exchanged only this component of their momentum vectors.
379 >
380 > \subsection{RNEMD with NIVS scaling}
381 >
382 > For each simulation utilizing the swapping method, a corresponding
383 > NIVS-RNEMD simulation was carried out using a target momentum flux set
384 > to produce the same flux experienced in the swapping simulation.
385 >
386 > To test the temperature homogeneity, directional momentum and
387 > temperature distributions were accumulated for molecules in each of
388 > the slabs.  Transport coefficients were computed using the temperature
389 > (and momentum) gradients across the $z$-axis as well as the total
390 > momentum flux the system experienced during data collection portion of
391 > the simulation.
392 >
393 > \subsection{Shear viscosities}
394 >
395 > The momentum flux was calculated using the total non-physical momentum
396 > transferred (${P_x}$) and the data collection time ($t$):
397   \begin{equation}
398 < (x-1) P_h^x  = j_z(p_x)
398 > j_z(p_x) = \frac{P_x}{2 t L_x L_y}
399   \end{equation}
400 + where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
401 + box.  The factor of two in the denominator is present because physical
402 + momentum transfer between the slabs occurs in two directions ($+z$ and
403 + $-z$).  The velocity gradient ${\langle \partial v_x /\partial z
404 +  \rangle}$ was obtained using linear regression of the mean $x$
405 + component of the velocity, $\langle v_x \rangle$, in each of the bins.
406 + For Lennard-Jones simulations, shear viscosities are reported in
407 + reduced units (${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$).
408  
409 + \subsection{Thermal Conductivities}
410  
411 + The energy flux was calculated in a similar manner to the momentum
412 + flux, using the total non-physical energy transferred (${E_{total}}$)
413 + and the data collection time $t$:
414 + \begin{equation}
415 + J_z = \frac{E_{total}}{2 t L_x L_y}
416 + \end{equation}
417 + The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
418 + obtained by a linear regression of the temperature profile. For
419 + Lennard-Jones simulations, thermal conductivities are reported in
420 + reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
421 +  k_B^{-1}\varepsilon^{-1/2}}$).
422  
423 + \subsection{Interfacial Thermal Conductivities}
424  
425 + For interfaces with a relatively low interfacial conductance, the bulk
426 + regions on either side of an interface rapidly come to a state in
427 + which the two phases have relatively homogeneous (but distinct)
428 + temperatures.  The interfacial thermal conductivity $G$ can therefore
429 + be approximated as:
430  
431 < \section{Acknowledgments}
432 < Support for this project was provided by the National Science
433 < Foundation under grant CHE-0848243. Computational time was provided by
434 < the Center for Research Computing (CRC) at the University of Notre
435 < Dame.  \newpage
431 > \begin{equation}
432 > G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
433 >    \langle T_{water}\rangle \right)}
434 > \label{interfaceCalc}
435 > \end{equation}
436 > where ${E_{total}}$ is the imposed non-physical kinetic energy
437 > transfer and ${\langle T_{gold}\rangle}$ and ${\langle
438 >  T_{water}\rangle}$ are the average observed temperature of gold and
439 > water phases respectively.  If the interfacial conductance is {\it
440 >  not} small, it is also be possible to compute the interfacial
441 > thermal conductivity using this method utilizing the change in the
442 > slope of the thermal gradient ($\partial^2 \langle T \rangle / \partial
443 > z^2$) at the interface.
444  
445 < \bibliographystyle{jcp2}
445 > \section{Results}
446 >
447 > \subsection{Lennard-Jones Fluid}
448 > 2592 Lennard-Jones atoms were placed in an orthorhombic cell
449 > ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side.  The
450 > reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
451 > direct comparison between our results and previous methods.  These
452 > simulations were carried out with a reduced timestep ${\tau^* =
453 >  4.6\times10^{-4}}$.  For the shear viscosity calculations, the mean
454 > temperature was ${T^* = k_B T/\varepsilon = 0.72}$.  For thermal
455 > conductivity calculations, simulations were run under reduced
456 > temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
457 > ensemble. The simulations included $10^5$ steps of equilibration
458 > without any momentum flux, $10^5$ steps of stablization with an
459 > imposed momentum transfer to create a gradient, and $10^6$ steps of
460 > data collection under RNEMD.
461 >
462 > \subsubsection*{Thermal Conductivity}
463 >
464 > Our thermal conductivity calculations show that the NIVS method agrees
465 > well with the swapping method. Five different swap intervals were
466 > tested (Table \ref{LJ}). Similar thermal gradients were observed with
467 > similar thermal flux under the two different methods (Figure
468 > \ref{thermalGrad}). Furthermore, the 1-d temperature profiles showed
469 > no observable differences between the $x$, $y$ and $z$ axes (Figure
470 > \ref{thermalGrad} c), so even though we are using a non-isotropic
471 > scaling method, none of the three directions are experience
472 > disproportionate heating due to the velocity scaling.
473 >
474 > \begin{table*}
475 >  \begin{minipage}{\linewidth}
476 >    \begin{center}
477 >
478 >      \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
479 >        ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
480 >        ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
481 >        at various momentum fluxes.  The original swapping method and
482 >        the velocity scaling method give similar results.
483 >        Uncertainties are indicated in parentheses.}
484 >      
485 >      \begin{tabular}{|cc|cc|cc|}
486 >        \hline
487 >        \multicolumn{2}{|c}{Momentum Exchange} &
488 >        \multicolumn{2}{|c}{Swapping RNEMD} &
489 >        \multicolumn{2}{|c|}{NIVS-RNEMD} \\
490 >        \hline
491 >        \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
492 >        \multirow{2}{*}{$\lambda^*_{swap}$} &
493 >        \multirow{2}{*}{$\eta^*_{swap}$}  &
494 >        \multirow{2}{*}{$\lambda^*_{scale}$} &
495 >        \multirow{2}{*}{$\eta^*_{scale}$} \\
496 >        & $j_z^*(p_x)$ (reduced units) & & & & \\
497 >        \hline
498 >        250  & 0.16  & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
499 >        500  & 0.09  & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
500 >        1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
501 >        2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
502 >        2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
503 >        \hline
504 >      \end{tabular}
505 >      \label{LJ}
506 >    \end{center}
507 >  \end{minipage}
508 > \end{table*}
509 >
510 > \begin{figure}
511 >  \includegraphics[width=\linewidth]{thermalGrad}
512 >  \caption{The NIVS-RNEMD method creates similar temperature gradients
513 >    compared with the swapping method under a variety of imposed
514 >    kinetic energy flux values. Furthermore, the implementation of
515 >    Non-Isotropic Velocity Scaling does not cause temperature
516 >    anisotropy to develop in thermal conductivity calculations.}
517 >  \label{thermalGrad}
518 > \end{figure}
519 >
520 > \subsubsection*{Velocity Distributions}
521 >
522 > During these simulations, velocities were recorded every 1000 steps
523 > and were used to produce distributions of both velocity and speed in
524 > each of the slabs. From these distributions, we observed that under
525 > relatively high non-physical kinetic energy flux, the speed of
526 > molecules in slabs where swapping occured could deviate from the
527 > Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
528 > and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
529 > distributions deviate from an ideal distribution. In the ``hot'' slab,
530 > the probability density is notched at low speeds and has a substantial
531 > shoulder at higher speeds relative to the ideal MB distribution. In
532 > the cold slab, the opposite notching and shouldering occurs.  This
533 > phenomenon is more obvious at higher swapping rates.
534 >
535 > The peak of the velocity distribution is substantially flattened in
536 > the hot slab, and is overly sharp (with truncated wings) in the cold
537 > slab. This problem is rooted in the mechanism of the swapping method.
538 > Continually depleting low (high) speed particles in the high (low)
539 > temperature slab is not complemented by diffusions of low (high) speed
540 > particles from neighboring slabs, unless the swapping rate is
541 > sufficiently small. Simutaneously, surplus low speed particles in the
542 > low temperature slab do not have sufficient time to diffuse to
543 > neighboring slabs.  Since the thermal exchange rate must reach a
544 > minimum level to produce an observable thermal gradient, the
545 > swapping-method RNEMD has a relatively narrow choice of exchange times
546 > that can be utilized.
547 >
548 > For comparison, NIVS-RNEMD produces a speed distribution closer to the
549 > Maxwell-Boltzmann curve (Figure \ref{thermalHist}).  The reason for
550 > this is simple; upon velocity scaling, a Gaussian distribution remains
551 > Gaussian.  Although a single scaling move is non-isotropic in three
552 > dimensions, our criteria for choosing a set of scaling coefficients
553 > helps maintain the distributions as close to isotropic as possible.
554 > Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
555 > as the previous RNEMD methods but without large perturbations to the
556 > velocity distributions in the two slabs.
557 >
558 > \begin{figure}
559 > \includegraphics[width=\linewidth]{thermalHist}
560 > \caption{Velocity and speed distributions that develop under the
561 >  swapping and NIVS-RNEMD methods at high flux.  The distributions for
562 >  the hot bins (upper panels) and cold bins (lower panels) were
563 >  obtained from Lennard-Jones simulations with $\langle T^* \rangle =
564 >  4.5$ with a flux of $J_z^* \sim 5$ (equivalent to a swapping interval
565 >  of 10 time steps).  This is a relatively large flux which shows the
566 >  non-thermal distributions that develop under the swapping method.
567 >  NIVS does a better job of producing near-thermal distributions in
568 >  the bins.}
569 > \label{thermalHist}
570 > \end{figure}
571 >
572 >
573 > \subsubsection*{Shear Viscosity}
574 > Our calculations (Table \ref{LJ}) show that velocity-scaling RNEMD
575 > predicted comparable shear viscosities to swap RNEMD method. The
576 > average molecular momentum gradients of these samples are shown in
577 > Figure \ref{shear} (a) and (b).
578 >
579 > \begin{figure}
580 >  \includegraphics[width=\linewidth]{shear}
581 >  \caption{Average momentum gradients in shear viscosity simulations,
582 >    using ``swapping'' method (top panel) and NIVS-RNEMD method
583 >    (middle panel). NIVS-RNEMD produces a thermal anisotropy artifact
584 >    in the hot and cold bins when used for shear viscosity.  This
585 >    artifact does not appear in thermal conductivity calculations.}
586 >  \label{shear}
587 > \end{figure}
588 >
589 > Observations of the three one-dimensional temperatures in each of the
590 > slabs shows that NIVS-RNEMD does produce some thermal anisotropy,
591 > particularly in the hot and cold slabs.  Figure \ref{shear} (c)
592 > indicates that with a relatively large imposed momentum flux,
593 > $j_z(p_x)$, the $x$ direction approaches a different temperature from
594 > the $y$ and $z$ directions in both the hot and cold bins.  This is an
595 > artifact of the scaling constraints.  After the momentum gradient has
596 > been established, $P_c^x < 0$.  Consequently, the scaling factor $x$
597 > is nearly always greater than one in order to satisfy the constraints.
598 > This will continually increase the kinetic energy in $x$-dimension,
599 > $K_c^x$.  If there is not enough time for the kinetic energy to
600 > exchange among different directions and different slabs, the system
601 > will exhibit the observed thermal anisotropy in the hot and cold bins.
602 >
603 > Although results between scaling and swapping methods are comparable,
604 > the inherent temperature anisotropy does make NIVS-RNEMD method less
605 > attractive than swapping RNEMD for shear viscosity calculations.  We
606 > note that this problem appears only when momentum flux is applied, and
607 > does not appear in thermal flux calculations.
608 >
609 > \subsection{Bulk SPC/E water}
610 >
611 > We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
612 > to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
613 > the original swapping RNEMD method.  Bedrov {\it et
614 >  al.}\cite{Bedrov:2000} argued that exchange of the molecule
615 > center-of-mass velocities instead of single atom velocities in a
616 > molecule conserves the total kinetic energy and linear momentum.  This
617 > principle is also adopted Fin our simulations. Scaling was applied to
618 > the center-of-mass velocities of the rigid bodies of SPC/E model water
619 > molecules.
620 >
621 > To construct the simulations, a simulation box consisting of 1000
622 > molecules were first equilibrated under ambient pressure and
623 > temperature conditions using the isobaric-isothermal (NPT)
624 > ensemble.\cite{melchionna93} A fixed volume was chosen to match the
625 > average volume observed in the NPT simulations, and this was followed
626 > by equilibration, first in the canonical (NVT) ensemble, followed by a
627 > 100~ps period under constant-NVE conditions without any momentum flux.
628 > Another 100~ps was allowed to stabilize the system with an imposed
629 > momentum transfer to create a gradient, and 1~ns was allotted for data
630 > collection under RNEMD.
631 >
632 > In our simulations, the established temperature gradients were similar
633 > to the previous work.  Our simulation results at 318K are in good
634 > agreement with those from Bedrov {\it et al.} (Table
635 > \ref{spceThermal}). And both methods yield values in reasonable
636 > agreement with experimental values.  
637 >
638 > \begin{table*}
639 >  \begin{minipage}{\linewidth}
640 >    \begin{center}
641 >      
642 >      \caption{Thermal conductivity of SPC/E water under various
643 >        imposed thermal gradients. Uncertainties are indicated in
644 >        parentheses.}
645 >      
646 >      \begin{tabular}{|c|c|ccc|}
647 >        \hline
648 >        \multirow{2}{*}{$\langle T\rangle$(K)} &
649 >        \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
650 >        \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
651 >          \mathrm{K}^{-1})$} \\
652 >        & & This work & Previous simulations\cite{Bedrov:2000} &
653 >        Experiment\cite{WagnerKruse}\\
654 >        \hline
655 >        \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
656 >        \multirow{3}{*}{0.61}\\
657 >        & 0.81 & 0.770(0.008) & & \\
658 >        & 1.54 & 0.813(0.007) & & \\
659 >        \hline
660 >        \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
661 >        \multirow{2}{*}{0.64}\\
662 >        & 1.59 & 0.778(0.019) & 0.730 & \\
663 >        \hline
664 >      \end{tabular}
665 >      \label{spceThermal}
666 >    \end{center}
667 >  \end{minipage}
668 > \end{table*}
669 >
670 > \subsection{Crystalline Gold}
671 >
672 > To see how the method performed in a solid, we calculated thermal
673 > conductivities using two atomistic models for gold.  Several different
674 > potential models have been developed that reasonably describe
675 > interactions in transition metals. In particular, the Embedded Atom
676 > Model (EAM)~\cite{PhysRevB.33.7983} and Sutton-Chen (SC)~\cite{Chen90}
677 > potential have been used to study a wide range of phenomena in both
678 > bulk materials and
679 > nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
680 > Both potentials are based on a model of a metal which treats the
681 > nuclei and core electrons as pseudo-atoms embedded in the electron
682 > density due to the valence electrons on all of the other atoms in the
683 > system. The SC potential has a simple form that closely resembles the
684 > Lennard Jones potential,
685 > \begin{equation}
686 > \label{eq:SCP1}
687 > 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] ,
688 > \end{equation}
689 > where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
690 > \begin{equation}
691 > \label{eq:SCP2}
692 > 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}}.
693 > \end{equation}
694 > $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
695 > interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
696 > Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
697 > the interactions between the valence electrons and the cores of the
698 > pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
699 > scale, $c_i$ scales the attractive portion of the potential relative
700 > to the repulsive interaction and $\alpha_{ij}$ is a length parameter
701 > that assures a dimensionless form for $\rho$. These parameters are
702 > tuned to various experimental properties such as the density, cohesive
703 > energy, and elastic moduli for FCC transition metals. The quantum
704 > Sutton-Chen (QSC) formulation matches these properties while including
705 > zero-point quantum corrections for different transition
706 > metals.\cite{PhysRevB.59.3527} The EAM functional forms differ
707 > slightly from SC but the overall method is very similar.
708 >
709 > In this work, we have utilized both the EAM and the QSC potentials to
710 > test the behavior of scaling RNEMD.
711 >
712 > A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
713 > atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
714 > both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
715 > pre-equilibration at a target pressure of 1 atm. When equilibrated
716 > under NPT conditions, our simulation box expanded by approximately 1\%
717 > in volume. Following adjustment of the box volume, equilibrations in
718 > both the canonical and microcanonical ensembles were carried out. With
719 > the simulation cell divided evenly into 10 slabs, different thermal
720 > gradients were established by applying a set of target thermal
721 > transfer fluxes.
722 >
723 > The results for the thermal conductivity of gold are shown in Table
724 > \ref{AuThermal}.  In these calculations, the end and middle slabs were
725 > excluded in thermal gradient linear regession. EAM predicts slightly
726 > larger thermal conductivities than QSC.  However, both values are
727 > smaller than experimental value by a factor of more than 200. This
728 > behavior has been observed previously by Richardson and Clancy, and
729 > has been attributed to the lack of electronic contribution in these
730 > force fields.\cite{Clancy:1992} It should be noted that the density of
731 > the metal being simulated has an effect on thermal conductance. With
732 > an expanded lattice, lower thermal conductance is expected (and
733 > observed). We also observed a decrease in thermal conductance at
734 > higher temperatures, a trend that agrees with experimental
735 > measurements.\cite{AshcroftMermin}
736 >
737 > \begin{table*}
738 >  \begin{minipage}{\linewidth}
739 >    \begin{center}
740 >      
741 >      \caption{Calculated thermal conductivity of crystalline gold
742 >        using two related force fields. Calculations were done at both
743 >        experimental and equilibrated densities and at a range of
744 >        temperatures and thermal flux rates. Uncertainties are
745 >        indicated in parentheses. Richardson {\it et
746 >          al.}\cite{Clancy:1992} give an estimate of 1.74 $\mathrm{W
747 >          m}^{-1}\mathrm{K}^{-1}$ for EAM gold
748 >         at a density of 19.263 g / cm$^3$.}
749 >      
750 >      \begin{tabular}{|c|c|c|cc|}
751 >        \hline
752 >        Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
753 >        $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
754 >        \hline
755 >        \multirow{7}{*}{QSC} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
756 >        &        &     & 2.86 & 1.08(0.05)\\
757 >        &        &     & 5.14 & 1.15(0.07)\\\cline{2-5}
758 >        & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
759 >        &        &     & 3.02 & 1.26(0.05)\\\cline{3-5}
760 >        &        & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
761 >        &        &     & 4.84 & 0.92(0.05)\\
762 >        \hline
763 >        \multirow{8}{*}{EAM} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
764 >        &        &     & 2.06 & 1.37(0.04)\\
765 >        &        &     & 2.55 & 1.41(0.07)\\\cline{2-5}
766 >        & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
767 >        &        &     & 2.04 & 1.41(0.07)\\
768 >        &        &     & 2.41 & 1.53(0.10)\\\cline{3-5}
769 >        &        & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
770 >        &        &     & 4.14 & 1.08(0.05)\\
771 >        \hline
772 >      \end{tabular}
773 >      \label{AuThermal}
774 >    \end{center}
775 >  \end{minipage}
776 > \end{table*}
777 >
778 > \subsection{Thermal Conductance at the Au/H$_2$O interface}
779 > The most attractive aspect of the scaling approach for RNEMD is the
780 > ability to use the method in non-homogeneous systems, where molecules
781 > of different identities are segregated in different slabs.  To test
782 > this application, we simulated a Gold (111) / water interface.  To
783 > construct the interface, a box containing a lattice of 1188 Au atoms
784 > (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
785 > relax under ambient temperature and pressure.  A separate (but
786 > identically sized) box of SPC/E water was also equilibrated at ambient
787 > conditions.  The two boxes were combined by removing all water
788 > molecules within 3 \AA radius of any gold atom.  The final
789 > configuration contained 1862 SPC/E water molecules.
790 >
791 > The Spohr potential was adopted in depicting the interaction between
792 > metal atoms and water molecules.\cite{ISI:000167766600035} A similar
793 > protocol of equilibration to our water simulations was followed.  We
794 > observed that the two phases developed large temperature differences
795 > even under a relatively low thermal flux.
796 >
797 > The low interfacial conductance is probably due to an acoustic
798 > impedance mismatch between the solid and the liquid
799 > phase.\cite{Cahill:793,RevModPhys.61.605} Experiments on the thermal
800 > conductivity of gold nanoparticles and nanorods in solvent and in
801 > glass cages have predicted values for $G$ between 100 and 350
802 > (MW/m$^2$/K).  The experiments typically have multiple gold surfaces
803 > that have been protected by a capping agent (citrate or CTAB) or which
804 > are in direct contact with various glassy solids.  In these cases, the
805 > acoustic impedance mismatch would be substantially reduced, leading to
806 > much higher interfacial conductances.  It is also possible, however,
807 > that the lack of electronic effects that gives rise to the low thermal
808 > conductivity of EAM gold is also causing a low reading for this
809 > particular interface.
810 >
811 > Under this low thermal conductance, both gold and water phase have
812 > sufficient time to eliminate temperature difference inside
813 > respectively (Figure \ref{interface} b). With indistinguishable
814 > temperature difference within respective phase, it is valid to assume
815 > that the temperature difference between gold and water on surface
816 > would be approximately the same as the difference between the gold and
817 > water phase. This assumption enables convenient calculation of $G$
818 > using Eq.  \ref{interfaceCalc} instead of measuring temperatures of
819 > thin layer of water and gold close enough to surface, which would have
820 > greater fluctuation and lower accuracy. Reported results (Table
821 > \ref{interfaceRes}) are of two orders of magnitude smaller than our
822 > calculations on homogeneous systems, and thus have larger relative
823 > errors than our calculation results on homogeneous systems.
824 >
825 > \begin{figure}
826 > \includegraphics[width=\linewidth]{interface}
827 > \caption{Temperature profiles of the Gold / Water interface at four
828 >  different values for the thermal flux.  Temperatures for slabs
829 >  either in the gold or in the water show no significant differences,
830 >  although there is a large discontinuity between the materials
831 >  because of the relatively low interfacial thermal conductivity.}
832 > \label{interface}
833 > \end{figure}
834 >
835 > \begin{table*}
836 >  \begin{minipage}{\linewidth}
837 >    \begin{center}
838 >      
839 >      \caption{Computed interfacial thermal conductivity ($G$) values
840 >        for the Au(111) / water interface at ${\langle T\rangle \sim}$
841 >        300K using a range of energy fluxes. Uncertainties are
842 >        indicated in parentheses. }
843 >      
844 >      \begin{tabular}{|cccc| }
845 >        \hline
846 >        $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
847 >        T_{water} \rangle$ (K) & $G$
848 >        (MW/m$^2$/K)\\
849 >        \hline
850 >        98.0 & 355.2 & 295.8 & 1.65(0.21) \\
851 >        78.8 & 343.8 & 298.0 & 1.72(0.32) \\
852 >        73.6 & 344.3 & 298.0 & 1.59(0.24) \\
853 >        49.2 & 330.1 & 300.4 & 1.65(0.35) \\
854 >        \hline
855 >      \end{tabular}
856 >      \label{interfaceRes}
857 >    \end{center}
858 >  \end{minipage}
859 > \end{table*}
860 >
861 >
862 > \section{Conclusions}
863 > NIVS-RNEMD simulation method is developed and tested on various
864 > systems. Simulation results demonstrate its validity in thermal
865 > conductivity calculations, from Lennard-Jones fluid to multi-atom
866 > molecule like water and metal crystals. NIVS-RNEMD improves
867 > non-Boltzmann-Maxwell distributions, which exist inb previous RNEMD
868 > methods. Furthermore, it develops a valid means for unphysical thermal
869 > transfer between different species of molecules, and thus extends its
870 > applicability to interfacial systems. Our calculation of gold/water
871 > interfacial thermal conductivity demonstrates this advantage over
872 > previous RNEMD methods. NIVS-RNEMD has also limited application on
873 > shear viscosity calculations, but could cause temperature difference
874 > among different dimensions under high momentum flux. Modification is
875 > necessary to extend the applicability of NIVS-RNEMD in shear viscosity
876 > calculations.
877 >
878 > \section{Acknowledgments}
879 > The authors would like to thank Craig Tenney and Ed Maginn for many
880 > helpful discussions.  Support for this project was provided by the
881 > National Science Foundation under grant CHE-0848243. Computational
882 > time was provided by the Center for Research Computing (CRC) at the
883 > University of Notre Dame.  
884 > \newpage
885 >
886   \bibliography{nivsRnemd}
887 +
888   \end{doublespace}
889   \end{document}
890  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines