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 3583 by gezelter, Tue Apr 13 19:59:50 2010 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 39 | Line 44 | Notre Dame, Indiana 46556}
44  
45   \begin{abstract}
46    We present a new method for introducing stable non-equilibrium
47 <  velocity and temperature distributions in molecular dynamics
48 <  simulations of heterogeneous systems.  This method extends some
49 <  earlier Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods
50 <  which use momentum exchange swapping moves that can create
51 <  non-thermal velocity distributions (and which are difficult to use
52 <  for interfacial calculations).  By using non-isotropic velocity
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 and stable thermal and momentum gradients can then be
56 <  established.  The scaling method we have developed conserves the
57 <  total linear momentum and total energy of the system.  To test the
58 <  methods, we have computed the thermal conductivity of model liquid
59 <  and solid systems as well as the interfacial thermal conductivity of
60 <  a metal-water interface.  We find that the NIVS-RNEMD improves the
56 <  problematic velocity distributions that develop in other RNEMD
57 <  methods.
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 65 | Line 68 | Notre Dame, Indiana 46556}
68   %                          BODY OF TEXT
69   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70  
68
69
71   \section{Introduction}
72   The original formulation of Reverse Non-equilibrium Molecular Dynamics
73   (RNEMD) obtains transport coefficients (thermal conductivity and shear
# Line 86 | Line 87 | RNEMD has been widely used to provide computational es
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 has 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).\cite{ISI:000246190100032}  [MORE CITATIONS HERE]
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   \begin{figure}
97   \includegraphics[width=\linewidth]{thermalDemo}
# Line 102 | Line 104 | RNEMD is preferable in many ways to the forward NEMD m
104   \label{thermalDemo}
105   \end{figure}
106  
107 < RNEMD is preferable in many ways to the forward NEMD methods
108 < [CITATIONS NEEDED] because it imposes what is typically difficult to measure
109 < (a flux or stress) and it is typically much easier to compute momentum
110 < gradients or strains (the response).  For similar reasons, RNEMD is
111 < also preferable to slowly-converging equilibrium methods for measuring
112 < thermal conductivity and shear viscosity (using Green-Kubo relations
113 < [CITATIONS NEEDED] or the Helfand moment approach of Viscardy {\it et
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 118 | Line 122 | Recently, Tenney and Maginn\cite{ISI:000273472300004}
122   typically samples from the same manifold of states in the
123   microcanonical ensemble.
124  
125 < Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
126 < some problems with the original RNEMD swap technique.  Notably, large
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 self-adjusting metrics for retaining the usability of the method.
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
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 single component, multi-component, and
142 < non-isotropic mixtures and show that it is capable of providing
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 M\"{u}ller-Plathe's RNEMD method; the
# Line 156 | Line 161 | where ${x, y, z}$ are a set of 3 scaling variables for
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 hot slab 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}{ccc}
170   x^\prime & 0 & 0 \\
# Line 175 | Line 181 | P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_
181   \end{equation}
182   where
183   \begin{eqnarray}
184 < P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
185 < P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
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{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 momentum in each of the two slabs.
190 > the instantaneous linear momenta in each of the two slabs.
191   \begin{equation}
192   \alpha^\prime = 1 + (1 - \alpha) p_\alpha
193   \label{eq:hotcoldscaling}
# Line 194 | Line 200 | Conservation of total energy also places constraints o
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 < \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^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 translational kinetic energies, $K_h^\alpha$ and
207   $K_c^\alpha$, are computed for each of the three directions in a
# Line 204 | Line 210 | Substituting in the expressions for the hot scaling pa
210   ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
211   {\it constraint ellipsoid}:
212   \begin{equation}
213 < \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 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
# Line 217 | Line 224 | cold slab scaling parameters which can be applied whil
224   \label{eq:constraintEllipsoidConsts}
225   \end{eqnarray}
226   This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
227 < cold slab scaling parameters which can be applied while preserving
228 < both linear momentum in all three directions as well as total kinetic
229 < energy.
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 decrease in kinetic energy of the cold 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   (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_c^x + y^2 K_c^y + z^2 K_c^z = K_c^x + K_c^y + K_c^z - J_z\Delta t
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 thermal flux ellipsoid} is governed
246 < both by a targetted value, $J_z$ as well as the instantaneous values
247 < of the kinetic energy components in the cold bin.
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 < we must determine the points ${x,y,z}$ which lie on both the
251 < constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
252 < ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
253 < two ellipsoids in 3-dimensional space.
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{Scaling points which maintain both constant energy and
258 <  constant linear momentum of the system lie on the surface of the
259 <  {\it constraint ellipsoid} while points which generate the target
260 <  momentum flux lie on the surface of the {\it flux ellipsoid}.  The
261 <  velocity distributions in the cold bin are scaled by only those
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 < One may also define {\it momentum} flux (say along the $x$-direction) as:
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 < The above {\it momentum flux plane} is perpendicular to the $x$-axis,
314 < with its position governed both by a target value, $j_z(p_x)$ as well
315 < as the instantaneous value of the momentum along the $x$-direction.
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 plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
321 < ellipsoid and a plane in 3-dimensional space.  
321 > ellipsoid and a plane in 3-dimensional space.
322  
323 < In both the momentum and energy flux scenarios, valid scaling
324 < parameters are arrived at by solving geometric intersection problems
325 < in $x, y, z$ space in order to obtain cold slab scaling parameters.
326 < Once the scaling variables for the cold slab are known, the hot slab
327 < scaling has also been determined.
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 + \section{Computational Details}
330  
331 < The following problem will be choosing an optimal set of scaling
332 < variables among the possible sets. Although this method is inherently
333 < non-isotropic, the goal is still to maintain the system as isotropic
334 < as possible. Under this consideration, one would like the kinetic
335 < energies in different directions could become as close as each other
336 < after each scaling. Simultaneously, one would also like each scaling
337 < as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
338 < large perturbation to the system. Therefore, one approach to obtain
339 < the scaling variables would be constructing an criteria function, with
340 < constraints as above equation sets, and solving the function's minimum
341 < by method like Lagrange multipliers.
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 < In order to save computation time, we have a different approach to a
293 < relatively good set of scaling variables with much less calculation
294 < than above. Here is the detail of our simplification of the problem.
344 > \subsection{Simulation Cells}
345  
346 < In the case of kinetic energy transfer, we impose another constraint
347 < ${x = y}$, into the equation sets. Consequently, there are two
348 < variables left. And now one only needs to solve a set of two {\it
349 <  ellipses equations}. This problem would be transformed into solving
350 < one quartic equation for one of the two variables. There are known
351 < generic methods that solve real roots of quartic equations. Then one
352 < can determine the other variable and obtain sets of scaling
353 < variables. Among these sets, one can apply the above criteria to
354 < choose the best set, while much faster with only a few sets to choose.
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 < In the case of momentum flux transfer, we impose another constraint to
307 < set the kinetic energy transfer as zero. In another word, we apply
308 < Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
309 < variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
310 < of equations on the above kinetic energy transfer problem. Therefore,
311 < an approach similar to the above would be sufficient for this as well.
357 > \subsection{RNEMD with M\"{u}ller-Plathe swaps}
358  
359 < \section{Computational Details}
360 < \subsection{Lennard-Jones Fluid}
361 < Our simulation consists of a series of systems. All of these
362 < simulations were run with the OpenMD simulation software
363 < package\cite{Meineke:2005gd} integrated with RNEMD codes.
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 < A Lennard-Jones fluid system was built and tested first. In order to
369 < compare our method with swapping RNEMD, a series of simulations were
370 < performed to calculate the shear viscosity and thermal conductivity of
371 < argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
372 <  \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
373 < ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
374 < comparison between our results and others. These simulations used
326 < velocity Verlet algorithm with reduced timestep ${\tau^* =
327 <  4.6\times10^{-4}}$.
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 viscosity calculation, the reduced temperature was ${T^* =
377 <  k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
378 < ensemble (NVT), then equilibrated in microcanonical ensemble
332 < (NVE). Establishing and stablizing momentum gradient were followed
333 < also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
334 < adopted.\cite{ISI:000080382700030} The simulation box was under
335 < periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
336 < the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
337 < most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
338 < to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
339 < frequency were chosen. According to each result from swapping
340 < RNEMD, scaling RNEMD simulations were run with the target momentum
341 < flux set to produce a similar momentum flux, and consequently shear
342 < rate. Furthermore, various scaling frequencies can be tested for one
343 < single swapping rate. To test the temperature homogeneity in our
344 < system of swapping and scaling methods, temperatures of different
345 < dimensions in all the slabs were observed. Most of the simulations
346 < include $10^5$ steps of equilibration without imposing momentum flux,
347 < $10^5$ steps of stablization with imposing unphysical momentum
348 < transfer, and $10^6$ steps of data collection under RNEMD. For
349 < relatively high momentum flux simulations, ${5\times10^5}$ step data
350 < collection is sufficient. For some low momentum flux simulations,
351 < ${2\times10^6}$ steps were necessary.
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 < After each simulation, the shear viscosity was calculated in reduced
381 < unit. The momentum flux was calculated with total unphysical
382 < transferred momentum ${P_x}$ and data collection time $t$:
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   j_z(p_x) = \frac{P_x}{2 t L_x L_y}
399   \end{equation}
400 < where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
401 < box, and physical momentum transfer occurs in two ways due to our
402 < periodic boundary condition settings. And the velocity gradient
403 < ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
404 < linear regression of the velocity profile. From the shear viscosity
405 < $\eta$ calculated with the above parameters, one can further convert
406 < it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
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 < For thermal conductivity calculations, simulations were first run under
368 < reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
369 < ensemble. Muller-Plathe's algorithm was adopted in the swapping
370 < method. Under identical simulation box parameters with our shear
371 < viscosity calculations, in each swap, the top slab exchanges all three
372 < translational momentum components of the molecule with least kinetic
373 < energy with the same components of the molecule in the center slab
374 < with most kinetic energy, unless this ``coldest'' molecule in the
375 < ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
376 < ``cold'' slab. According to swapping RNEMD results, target energy flux
377 < for scaling RNEMD simulations can be set. Also, various scaling
378 < frequencies can be tested for one target energy flux. To compare the
379 < performance between swapping and scaling method, distributions of
380 < velocity and speed in different slabs were observed.
409 > \subsection{Thermal Conductivities}
410  
411 < For each swapping rate, thermal conductivity was calculated in reduced
412 < unit. The energy flux was calculated similarly to the momentum flux,
413 < with total unphysical transferred energy ${E_{total}}$ and data collection
385 < time $t$:
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 < And the temperature gradient ${\langle\partial T/\partial z\rangle}$
418 < can be obtained by a linear regression of the temperature
419 < profile. From the thermal conductivity $\lambda$ calculated, one can
420 < further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
421 <  m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
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{ Water / Metal Thermal Conductivity}
396 < Another series of our simulation is the calculation of interfacial
397 < thermal conductivity of a Au/H$_2$O system. Respective calculations of
398 < liquid water (Extended Simple Point Charge model) and crystal gold
399 < thermal conductivity were performed and compared with current results
400 < to ensure the validity of NIVS-RNEMD. After that, a mixture system was
401 < simulated.
423 > \subsection{Interfacial Thermal Conductivities}
424  
425 < For thermal conductivity calculation of bulk water, a simulation box
426 < consisting of 1000 molecules were first equilibrated under ambient
427 < pressure and temperature conditions using NPT ensemble, followed by
428 < equilibration in fixed volume (NVT). The system was then equilibrated in
429 < microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
408 < stable thermal gradient was followed. The simulation box was under
409 < periodic boundary condition and devided into 10 slabs. Data collection
410 < process was similar to Lennard-Jones fluid system.
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  
412 Thermal conductivity calculation of bulk crystal gold used a similar
413 protocol. Two types of force field parameters, Embedded Atom Method
414 (EAM) and Quantum Sutten-Chen (QSC) force field were used
415 respectively. The face-centered cubic crystal simulation box consists of
416 2880 Au atoms. The lattice was first allowed volume change to relax
417 under ambient temperature and pressure. Equilibrations in canonical and
418 microcanonical ensemble were followed in order. With the simulation
419 lattice devided evenly into 10 slabs, different thermal gradients were
420 established by applying a set of target thermal transfer flux. Data of
421 the series of thermal gradients was collected for calculation.
422
423 After simulations of bulk water and crystal gold, a mixture system was
424 constructed, consisting of 1188 Au atoms and 1862 H$_2$O
425 molecules. Spohr potential was adopted in depicting the interaction
426 between metal atom and water molecule.\cite{ISI:000167766600035} A
427 similar protocol of equilibration was followed. Several thermal
428 gradients was built under different target thermal flux. It was found
429 out that compared to our previous simulation systems, the two phases
430 could have large temperature difference even under a relatively low
431 thermal flux. Therefore, under our low flux conditions, it is assumed
432 that the metal and water phases have respectively homogeneous
433 temperature, excluding the surface regions. In calculating the
434 interfacial thermal conductivity $G$, this assumptioin was applied and
435 thus our formula becomes:
436
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 unphysical kinetic energy transfer
437 < and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
438 < average observed temperature of gold and water phases respectively.
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 < \section{Results And Discussions}
447 < \subsection{Thermal Conductivity}
448 < \subsubsection{Lennard-Jones Fluid}
449 < Our thermal conductivity calculations show that scaling method results
450 < agree with swapping method. Four different exchange intervals were
451 < tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
452 < 10fs exchange interval, target exchange kinetic energy was set to
453 < produce equivalent kinetic energy flux as in swapping method. And
454 < similar thermal gradients were observed with similar thermal flux in
455 < two simulation methods (Figure \ref{thermalGrad}).
445 > \section{Results}
446  
447 < \begin{table*}
448 < \begin{minipage}{\linewidth}
449 < \begin{center}
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 < \caption{Calculation results for thermal conductivity of Lennard-Jones
462 <  fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
463 <  swap and scale methods at various kinetic energy exchange rates. Results
464 <  in reduced unit. Errors of calculations in parentheses.}
462 > \subsubsection*{Thermal Conductivity}
463  
464 < \begin{tabular}{ccc}
465 < \hline
466 < (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
467 < $\lambda^*_{scale}$\\
468 < \hline
469 < 250  & 7.03(0.34) & 7.30(0.10)\\
470 < 500  & 7.03(0.14) & 6.95(0.09)\\
471 < 1000 & 6.91(0.42) & 7.19(0.07)\\
472 < 2000 & 7.52(0.15) & 7.19(0.28)\\
473 < \hline
474 < \end{tabular}
475 < \label{thermalLJRes}
476 < \end{center}
477 < \end{minipage}
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{Temperature gradients under various kinetic energy flux of
513 <  thermal conductivity simulations}
514 < \label{thermalGrad}
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 < During these simulations, molecule velocities were recorded in 1000 of
490 < all the snapshots of one single data collection process. These
491 < velocity data were used to produce histograms of velocity and speed
492 < distribution in different slabs. From these histograms, it is observed
493 < that under relatively high unphysical kinetic energy flux, speed and
494 < velocity distribution of molecules in slabs where swapping occured
495 < could deviate from Maxwell-Boltzmann distribution. Figure
496 < \ref{histSwap} illustrates how these distributions deviate from an
497 < ideal distribution. In high temperature slab, probability density in
498 < low speed is confidently smaller than ideal curve fit; in low
499 < temperature slab, probability density in high speed is smaller than
500 < ideal, while larger than ideal in low speed. This phenomenon is more
501 < obvious in our high swapping rate simulations. And this deviation
502 < could also leads to deviation of distribution of velocity in various
503 < dimensions. One feature of these deviated distribution is that in high
504 < temperature slab, the ideal Gaussian peak was changed into a
505 < relatively flat plateau; while in low temperature slab, that peak
506 < appears sharper. This problem is rooted in the mechanism of the
507 < swapping method. Continually depleting low (high) speed particles in
508 < the high (low) temperature slab could not be complemented by
509 < diffusions of low (high) speed particles from neighbor slabs, unless
510 < in suffciently low swapping rate. Simutaneously, surplus low speed
511 < particles in the low temperature slab do not have sufficient time to
512 < diffuse to neighbor slabs. However, thermal exchange rate should reach
513 < a minimum level to produce an observable thermal gradient under noise
514 < interference. Consequently, swapping RNEMD has a relatively narrow
515 < choice of swapping rate to satisfy these above restrictions.
520 > \subsubsection*{Velocity Distributions}
521  
522 < \begin{figure}
523 < \includegraphics[width=\linewidth]{histSwap}
524 < \caption{Speed distribution for thermal conductivity using swapping
525 <  RNEMD. Shown is from the simulation with 250 fs exchange interval.}
526 < \label{histSwap}
527 < \end{figure}
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 < Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
536 < curve fit (Figure \ref{histScale}). Essentially, after scaling, a
537 < Gaussian distribution function would remain Gaussian. Although a
538 < single scaling is non-isotropic in all three dimensions, our scaling
539 < coefficient criteria could help maintian the scaling region as
540 < isotropic as possible. On the other hand, scaling coefficients are
541 < preferred to be as close to 1 as possible, which also helps minimize
542 < the difference among different dimensions. This is possible if scaling
543 < interval and one-time thermal transfer energy are well
544 < chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
545 < thermal flux as the previous RNEMD method without large perturbation
546 < to the distribution of velocity and speed in the exchange regions.
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]{histScale}
560 < \caption{Speed distribution for thermal conductivity using scaling
561 <  RNEMD. Shown is from the simulation with an equilvalent thermal flux
562 <  as an 250 fs exchange interval swapping simulation.}
563 < \label{histScale}
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  
545 \subsubsection{SPC/E Water}
546 Our results of SPC/E water thermal conductivity are comparable to
547 Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
548 previous swapping RNEMD method for their calculation. Bedrov {\it et
549  al.}\cite{ISI:000090151400044} argued that exchange of the molecule
550 center-of-mass velocities instead of single atom velocities in a
551 molecule conserves the total kinetic energy and linear momentum. This
552 principle is adopted in our simulations. Scaling is applied to the
553 velocities of the rigid bodies of SPC/E model water molecules, instead
554 of each hydrogen and oxygen atoms in relevant water molecules. As
555 shown in Figure \ref{spceGrad}, temperature gradients were established
556 similar to their system. However, the average temperature of our
557 system is 300K, while theirs is 318K, which would be attributed for
558 part of the difference between the final calculation results (Table
559 \ref{spceThermal}). Both methods yields values in agreement with
560 experiment. And this shows the applicability of our method to
561 multi-atom molecular system.
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]{spceGrad}
581 < \caption{Temperature gradients for SPC/E water thermal conductivity.}
582 < \label{spceGrad}
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 < \begin{table*}
590 < \begin{minipage}{\linewidth}
591 < \begin{center}
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 < \caption{Calculation results for thermal conductivity of SPC/E water
604 <  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
605 <  calculations in parentheses. }
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 < \begin{tabular}{cccc}
578 < \hline
579 < $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
580 < & This work & Previous simulations\cite{ISI:000090151400044} &
581 < Experiment$^a$\\
582 < \hline
583 < 0.38 & 0.816(0.044) & & 0.64\\
584 < 0.81 & 0.770(0.008) & 0.784\\
585 < 1.54 & 0.813(0.007) & 0.730\\
586 < \hline
587 < \end{tabular}
588 < \label{spceThermal}
589 < \end{center}
590 < \end{minipage}
591 < \end{table*}
609 > \subsection{Bulk SPC/E water}
610  
611 < \subsubsection{Crystal Gold}
612 < Our results of gold thermal conductivity using two force fields are
613 < shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
614 < these calculations,the end and middle slabs were excluded in thermal
615 < gradient regession and only used as heat source and drain in the
616 < systems. Our yielded values using EAM force field are slightly larger
617 < than those using QSC force field. However, both series are
618 < significantly smaller than experimental value by an order of more than
619 < 100. It has been verified that this difference is mainly attributed to
602 < the lack of electron interaction representation in these force field
603 < parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
604 < force field parameters in their metal thermal conductivity
605 < calculations. The Non-Equilibrium MD method they employed in their
606 < simulations produced comparable results to ours. As Zhang {\it et
607 <  al.}\cite{ISI:000231042800044} stated, thermal conductivity values
608 < are influenced mainly by force field. Therefore, it is confident to
609 < conclude that NIVS-RNEMD is applicable to metal force field system.
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 < \begin{figure}
622 < \includegraphics[width=\linewidth]{AuGrad}
623 < \caption{Temperature gradients for thermal conductivity calculation of
624 <  crystal gold using QSC force field.}
625 < \label{AuGrad}
626 < \end{figure}
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 < \begin{table*}
633 < \begin{minipage}{\linewidth}
634 < \begin{center}
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 < \caption{Calculation results for thermal conductivity of crystal gold
639 <  using QSC force field at ${\langle T\rangle}$ = 300K at various
640 <  thermal exchange rates. Errors of calculations in parentheses. }
641 <
642 < \begin{tabular}{cc}
643 < \hline
644 < $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
645 < \hline
646 < 1.44 & 1.10(0.01)\\
647 < 2.86 & 1.08(0.02)\\
648 < 5.14 & 1.15(0.01)\\
649 < \hline
650 < \end{tabular}
651 < \label{qscThermal}
652 < \end{center}
653 < \end{minipage}
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 < \begin{figure}
641 < \includegraphics[width=\linewidth]{eamGrad}
642 < \caption{Temperature gradients for thermal conductivity calculation of
643 <  crystal gold using EAM force field.}
644 < \label{eamGrad}
645 < \end{figure}
670 > \subsection{Crystalline Gold}
671  
672 < \begin{table*}
673 < \begin{minipage}{\linewidth}
674 < \begin{center}
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 < \caption{Calculation results for thermal conductivity of crystal gold
710 <  using EAM force field at ${\langle T\rangle}$ = 300K at various
653 <  thermal exchange rates. Errors of calculations in parentheses. }
709 > In this work, we have utilized both the EAM and the QSC potentials to
710 > test the behavior of scaling RNEMD.
711  
712 < \begin{tabular}{cc}
713 < \hline
714 < $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
715 < \hline
716 < 1.24 & 1.24(0.06)\\
717 < 2.06 & 1.37(0.04)\\
718 < 2.55 & 1.41(0.03)\\
719 < \hline
720 < \end{tabular}
721 < \label{eamThermal}
722 < \end{center}
723 < \end{minipage}
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 < \subsection{Interfaciel Thermal Conductivity}
792 < After simulations of homogeneous water and gold systems using
793 < NIVS-RNEMD method were proved valid, calculation of gold/water
794 < interfacial thermal conductivity was followed. It is found out that
795 < the low interfacial conductance is probably due to the hydrophobic
796 < surface in our system. Figure \ref{interfaceDensity} demonstrates mass
797 < density change along $z$-axis, which is perpendicular to the
798 < gold/water interface. It is observed that water density significantly
799 < decreases when approaching the surface. Under this low thermal
800 < conductance, both gold and water phase have sufficient time to
801 < eliminate temperature difference inside respectively (Figure
802 < \ref{interfaceGrad}). With indistinguishable temperature difference
803 < within respective phase, it is valid to assume that the temperature
804 < difference between gold and water on surface would be approximately
805 < the same as the difference between the gold and water phase. This
806 < assumption enables convenient calculation of $G$ using
807 < Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
808 < layer of water and gold close enough to surface, which would have
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]{interfaceDensity}
827 < \caption{Density profile for interfacial thermal conductivity
828 <  simulation box. Significant water density decrease is observed on
829 <  gold surface.}
830 < \label{interfaceDensity}
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  
701 \begin{figure}
702 \includegraphics[width=\linewidth]{interfaceGrad}
703 \caption{Temperature profiles for interfacial thermal conductivity
704  simulation box. Temperatures of different slabs in the same phase
705  show no significant difference.}
706 \label{interfaceGrad}
707 \end{figure}
708
835   \begin{table*}
836 < \begin{minipage}{\linewidth}
837 < \begin{center}
838 <
839 < \caption{Calculation results for interfacial thermal conductivity
840 <  at ${\langle T\rangle \sim}$ 300K at various thermal exchange
841 <  rates. Errors of calculations in parentheses. }
842 <
843 < \begin{tabular}{cccc}
844 < \hline
845 < $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
846 < \hline
847 < 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
848 < 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
849 < 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
850 < 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
851 < \hline
852 < \end{tabular}
853 < \label{interfaceRes}
854 < \end{center}
855 < \end{minipage}
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  
732 \subsection{Shear Viscosity}
733 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
734 produced comparable shear viscosity to swap RNEMD method. In Table
735 \ref{shearRate}, the names of the calculated samples are devided into
736 two parts. The first number refers to total slabs in one simulation
737 box. The second number refers to the swapping interval in swap method, or
738 in scale method the equilvalent swapping interval that the same
739 momentum flux would theoretically result in swap method. All the scale
740 method results were from simulations that had a scaling interval of 10
741 time steps. The average molecular momentum gradients of these samples
742 are shown in Figure \ref{shearGrad}.
861  
744 \begin{table*}
745 \begin{minipage}{\linewidth}
746 \begin{center}
747
748 \caption{Calculation results for shear viscosity of Lennard-Jones
749  fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
750  methods at various momentum exchange rates. Results in reduced
751  unit. Errors of calculations in parentheses. }
752
753 \begin{tabular}{ccc}
754 \hline
755 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
756 \hline
757 20-500 & 3.64(0.05) & 3.76(0.09)\\
758 20-1000 & 3.52(0.16) & 3.66(0.06)\\
759 20-2000 & 3.72(0.05) & 3.32(0.18)\\
760 20-2500 & 3.42(0.06) & 3.43(0.08)\\
761 \hline
762 \end{tabular}
763 \label{shearRate}
764 \end{center}
765 \end{minipage}
766 \end{table*}
767
768 \begin{figure}
769 \includegraphics[width=\linewidth]{shearGrad}
770 \caption{Average momentum gradients of shear viscosity simulations}
771 \label{shearGrad}
772 \end{figure}
773
774 \begin{figure}
775 \includegraphics[width=\linewidth]{shearTempScale}
776 \caption{Temperature profile for scaling RNEMD simulation.}
777 \label{shearTempScale}
778 \end{figure}
779 However, observations of temperatures along three dimensions show that
780 inhomogeneity occurs in scaling RNEMD simulations, particularly in the
781 two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
782 relatively large imposed momentum flux, the temperature difference among $x$
783 and the other two dimensions was significant. This would result from the
784 algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
785 momentum gradient is set up, $P_c^x$ would be roughly stable
786 ($<0$). Consequently, scaling factor $x$ would most probably larger
787 than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
788 keep increase after most scaling steps. And if there is not enough time
789 for the kinetic energy to exchange among different dimensions and
790 different slabs, the system would finally build up temperature
791 (kinetic energy) difference among the three dimensions. Also, between
792 $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
793 are closer to neighbor slabs. This is due to momentum transfer along
794 $z$ dimension between slabs.
795
796 Although results between scaling and swapping methods are comparable,
797 the inherent temperature inhomogeneity even in relatively low imposed
798 exchange momentum flux simulations makes scaling RNEMD method less
799 attractive than swapping RNEMD in shear viscosity calculation.
800
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 in previous RNEMD
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
# Line 815 | Line 876 | Support for this project was provided by the National
876   calculations.
877  
878   \section{Acknowledgments}
879 < Support for this project was provided by the National Science
880 < Foundation under grant CHE-0848243. Computational time was provided by
881 < the Center for Research Computing (CRC) at the University of Notre
882 < Dame.  \newpage
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  
823 \bibliographystyle{aip}
886   \bibliography{nivsRnemd}
887  
888   \end{doublespace}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines