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 3570 by skuang, Wed Mar 10 02:00:06 2010 UTC vs.
Revision 3611 by gezelter, Wed Jul 14 15:20:31 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}
# Line 38 | Line 39 | Notre Dame, Indiana 46556}
39   \begin{doublespace}
40  
41   \begin{abstract}
42 <
42 >  We present a new method for introducing stable non-equilibrium
43 >  velocity and temperature distributions in molecular dynamics
44 >  simulations of heterogeneous systems.  This method extends earlier
45 >  Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
46 >  momentum exchange swapping moves that can create non-thermal
47 >  velocity distributions and are difficult to use for interfacial
48 >  calculations.  By using non-isotropic velocity scaling (NIVS) on the
49 >  molecules in specific regions of a system, it is possible to impose
50 >  momentum or thermal flux between regions of a simulation and stable
51 >  thermal and momentum gradients can then be established.  The scaling
52 >  method we have developed conserves the total linear momentum and
53 >  total energy of the system.  To test the methods, we have computed
54 >  the thermal conductivity of model liquid and solid systems as well
55 >  as the interfacial thermal conductivity of a metal-water interface.
56 >  We find that the NIVS-RNEMD improves the problematic velocity
57 >  distributions that develop in other RNEMD methods.
58   \end{abstract}
59  
60   \newpage
# Line 49 | Line 65 | Notre Dame, Indiana 46556}
65   %                          BODY OF TEXT
66   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67  
52
53
68   \section{Introduction}
69   The original formulation of Reverse Non-equilibrium Molecular Dynamics
70   (RNEMD) obtains transport coefficients (thermal conductivity and shear
71   viscosity) in a fluid by imposing an artificial momentum flux between
72   two thin parallel slabs of material that are spatially separated in
73   the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 < artificial flux is typically created by periodically ``swapping'' either
75 < the entire momentum vector $\vec{p}$ or single components of this
76 < vector ($p_x$) between molecules in each of the two slabs.  If the two
77 < slabs are separated along the z coordinate, the imposed flux is either
78 < directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
79 < simulated system to the imposed momentum flux will typically be a
80 < velocity or thermal gradient.  The transport coefficients (shear
81 < viscosity and thermal conductivity) are easily obtained by assuming
82 < linear response of the system,
74 > artificial flux is typically created by periodically ``swapping''
75 > either the entire momentum vector $\vec{p}$ or single components of
76 > this vector ($p_x$) between molecules in each of the two slabs.  If
77 > the two slabs are separated along the $z$ coordinate, the imposed flux
78 > is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
79 > response of a simulated system to the imposed momentum flux will
80 > typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
81 > The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
82 > easily obtained by assuming linear response of the system,
83   \begin{eqnarray}
84   j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85 < J & = & \lambda \frac{\partial T}{\partial z}
85 > J_z & = & \lambda \frac{\partial T}{\partial z}
86   \end{eqnarray}
87 < RNEMD has been widely used to provide computational estimates of thermal
88 < conductivities and shear viscosities in a wide range of materials,
89 < from liquid copper to monatomic liquids to molecular fluids
90 < (e.g. ionic liquids).\cite{ISI:000246190100032}
87 > RNEMD has been widely used to provide computational estimates of
88 > thermal conductivities and shear viscosities in a wide range of
89 > materials, from liquid copper to both monatomic and molecular fluids
90 > (e.g.  ionic
91 > 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}
92  
93 < RNEMD is preferable in many ways to the forward NEMD methods because
94 < it imposes what is typically difficult to measure (a flux or stress)
95 < and it is typically much easier to compute momentum gradients or
96 < strains (the response).  For similar reasons, RNEMD is also preferable
97 < to slowly-converging equilibrium methods for measuring thermal
98 < conductivity and shear viscosity (using Green-Kubo relations or the
99 < Helfand moment approach of Viscardy {\it et
93 > \begin{figure}
94 > \includegraphics[width=\linewidth]{thermalDemo}
95 > \caption{RNEMD methods impose an unphysical transfer of momentum or
96 >  kinetic energy between a ``hot'' slab and a ``cold'' slab in the
97 >  simulation box.  The molecular system responds to this imposed flux
98 >  by generating a momentum or temperature gradient.  The slope of the
99 >  gradient can then be used to compute transport properties (e.g.
100 >  shear viscosity and thermal conductivity).}
101 > \label{thermalDemo}
102 > \end{figure}
103 >
104 > RNEMD is preferable in many ways to the forward NEMD
105 > methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
106 > because it imposes what is typically difficult to measure (a flux or
107 > stress) and it is typically much easier to compute the response
108 > (momentum gradients or strains).  For similar reasons, RNEMD is also
109 > preferable to slowly-converging equilibrium methods for measuring
110 > thermal conductivity and shear viscosity (using Green-Kubo
111 > relations\cite{daivis:541,mondello:9327} or the Helfand moment
112 > approach of Viscardy {\it et
113    al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
114   computing difficult to measure quantities.
115  
# Line 91 | Line 119 | Recently, Tenney and Maginn\cite{ISI:000273472300004}
119   typically samples from the same manifold of states in the
120   microcanonical ensemble.
121  
122 < Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
122 > Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
123   some problems with the original RNEMD swap technique.  Notably, large
124   momentum fluxes (equivalent to frequent momentum swaps between the
125 < slabs) can result in ``notched'', ``peaked'' and generally non-thermal momentum
126 < distributions in the two slabs, as well as non-linear thermal and
127 < velocity distributions along the direction of the imposed flux ($z$).
128 < Tenney and Maginn obtained reasonable limits on imposed flux and
129 < self-adjusting metrics for retaining the usability of the method.
125 > slabs) can result in ``notched'', ``peaked'' and generally non-thermal
126 > momentum distributions in the two slabs, as well as non-linear thermal
127 > and velocity distributions along the direction of the imposed flux
128 > ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
129 > and self-adjusting metrics for retaining the usability of the method.
130  
131   In this paper, we develop and test a method for non-isotropic velocity
132 < scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
132 > scaling (NIVS) which retains the desirable features of RNEMD
133   (conservation of linear momentum and total energy, compatibility with
134   periodic boundary conditions) while establishing true thermal
135 < distributions in each of the two slabs.  In the next section, we
136 < develop the method for determining the scaling constraints.  We then
137 < test the method on both single component, multi-component, and
138 < non-isotropic mixtures and show that it is capable of providing
135 > distributions in each of the two slabs. In the next section, we
136 > present the method for determining the scaling constraints.  We then
137 > test the method on both liquids and solids as well as a non-isotropic
138 > liquid-solid interface and show that it is capable of providing
139   reasonable estimates of the thermal conductivity and shear viscosity
140 < in these cases.
140 > in all of these cases.
141  
142   \section{Methodology}
143 < We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
144 < system is partitioned into a series of thin slabs along a particular
143 > We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
144 > periodic system is partitioned into a series of thin slabs along one
145   axis ($z$).  One of the slabs at the end of the periodic box is
146   designated the ``hot'' slab, while the slab in the center of the box
147   is designated the ``cold'' slab.  The artificial momentum flux will be
# Line 121 | Line 149 | moves.  For molecules $\{i\}$ located within the cold
149   hot slab.
150  
151   Rather than using momentum swaps, we use a series of velocity scaling
152 < moves.  For molecules $\{i\}$ located within the cold slab,
152 > moves.  For molecules $\{i\}$  located within the cold slab,
153   \begin{equation}
154   \vec{v}_i \leftarrow \left( \begin{array}{ccc}
155   x & 0 & 0 \\
# Line 129 | Line 157 | where ${x, y, z}$ are a set of 3 scaling variables for
157   0 & 0 & z \\
158   \end{array} \right) \cdot \vec{v}_i
159   \end{equation}
160 < where ${x, y, z}$ are a set of 3 scaling variables for each of the
161 < three directions in the system.  Likewise, the molecules $\{j\}$
162 < located in the hot slab will see a concomitant scaling of velocities,
160 > where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
161 > of the three directions in the system.  Likewise, the molecules
162 > $\{j\}$ located in the hot slab will see a concomitant scaling of
163 > velocities,
164   \begin{equation}
165   \vec{v}_j \leftarrow \left( \begin{array}{ccc}
166   x^\prime & 0 & 0 \\
# Line 141 | Line 170 | Conservation of linear momentum in each of the three d
170   \end{equation}
171  
172   Conservation of linear momentum in each of the three directions
173 < ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
173 > ($\alpha = x,y,z$) ties the values of the hot and cold scaling
174   parameters together:
175   \begin{equation}
176   P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
# Line 170 | Line 199 | where the kinetic energies, $K_h^\alpha$ and $K_c^\alp
199   \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
200   \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
201   \end{equation}
202 < where the kinetic energies, $K_h^\alpha$ and $K_c^\alpha$, are computed
203 < for each of the three directions in a similar manner to the linear momenta
204 < (Eq. \ref{eq:momentumdef}).  Substituting in the expressions for the
205 < hot scaling parameters ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}),
206 < we obtain the {\it constraint ellipsoid equation}:
202 > where the translational kinetic energies, $K_h^\alpha$ and
203 > $K_c^\alpha$, are computed for each of the three directions in a
204 > similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
205 > Substituting in the expressions for the hot scaling parameters
206 > ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
207 > {\it constraint ellipsoid}:
208   \begin{equation}
209 < \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
209 > \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
210 >  c_\alpha \right) = 0
211   \label{eq:constraintEllipsoid}
212   \end{equation}
213   where the constants are obtained from the instantaneous values of the
# Line 188 | Line 219 | This ellipsoid equation defines the set of cold slab s
219   c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
220   \label{eq:constraintEllipsoidConsts}
221   \end{eqnarray}
222 < This ellipsoid equation defines the set of cold slab scaling
223 < parameters which can be applied while preserving both linear momentum
224 < in all three directions as well as kinetic energy.
222 > This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
223 > cold slab scaling parameters which, when applied, preserve the linear
224 > momentum of the system in all three directions as well as total
225 > kinetic energy.
226  
227 < The goal of using velocity scaling variables is to transfer linear
228 < momentum or kinetic energy from the cold slab to the hot slab.  If the
229 < hot and cold slabs are separated along the z-axis, the energy flux is
230 < given simply by the decrease in kinetic energy of the cold bin:
227 > The goal of using these velocity scaling variables is to transfer
228 > kinetic energy from the cold slab to the hot slab.  If the hot and
229 > cold slabs are separated along the z-axis, the energy flux is given
230 > simply by the decrease in kinetic energy of the cold bin:
231   \begin{equation}
232   (1-x^2) K_c^x  + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
233   \end{equation}
234   The expression for the energy flux can be re-written as another
235   ellipsoid centered on $(x,y,z) = 0$:
236   \begin{equation}
237 < 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
237 > \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
238 > K_c^\alpha -J_z \Delta t
239   \label{eq:fluxEllipsoid}
240   \end{equation}
241 < The spatial extent of the {\it flux ellipsoid equation} is governed
242 < both by a targetted value, $J_z$ as well as the instantaneous values of the
243 < kinetic energy components in the cold bin.
241 > The spatial extent of the {\it thermal flux ellipsoid} is governed
242 > both by the target flux, $J_z$ as well as the instantaneous values of
243 > the kinetic energy components in the cold bin.
244  
245   To satisfy an energetic flux as well as the conservation constraints,
246 < it is sufficient to determine the points ${x,y,z}$ which lie on both
247 < the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
248 < flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
249 < the two ellipsoids in 3-dimensional space.
246 > we must determine the points ${x,y,z}$ that lie on both the constraint
247 > ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
248 > (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
249 > ellipsoids in 3-dimensional space.
250  
251   \begin{figure}
252   \includegraphics[width=\linewidth]{ellipsoids}
253 < \caption{Scaling points which maintain both constant energy and
254 <  constant linear momentum of the system lie on the surface of the
255 <  {\it constraint ellipsoid} while points which generate the target
256 <  momentum flux lie on the surface of the {\it flux ellipsoid}.  The
257 <  velocity distributions in the hot bin are scaled by only those
253 > \caption{Velocity scaling coefficients which maintain both constant
254 >  energy and constant linear momentum of the system lie on the surface
255 >  of the {\it constraint ellipsoid} while points which generate the
256 >  target momentum flux lie on the surface of the {\it flux ellipsoid}.
257 >  The velocity distributions in the cold bin are scaled by only those
258    points which lie on both ellipsoids.}
259   \label{ellipsoids}
260   \end{figure}
261  
262 < One may also define momentum flux (say along the x-direction) as:
262 > Since ellipsoids can be expressed as polynomials up to second order in
263 > each of the three coordinates, finding the the intersection points of
264 > two ellipsoids is isomorphic to finding the roots a polynomial of
265 > degree 16.  There are a number of polynomial root-finding methods in
266 > the literature, [CITATIONS NEEDED] but numerically finding the roots
267 > of high-degree polynomials is generally an ill-conditioned
268 > problem.[CITATION NEEDED] One simplification is to maintain velocity
269 > scalings that are {\it as isotropic as possible}.  To do this, we
270 > impose $x=y$, and to treat both the constraint and flux ellipsoids as
271 > 2-dimensional ellipses.  In reduced dimensionality, the
272 > intersecting-ellipse problem reduces to finding the roots of
273 > polynomials of degree 4.
274 >
275 > Depending on the target flux and current velocity distributions, the
276 > ellipsoids can have between 0 and 4 intersection points.  If there are
277 > no intersection points, it is not possible to satisfy the constraints
278 > while performing a non-equilibrium scaling move, and no change is made
279 > to the dynamics.  
280 >
281 > With multiple intersection points, any of the scaling points will
282 > conserve the linear momentum and kinetic energy of the system and will
283 > generate the correct target flux.  Although this method is inherently
284 > non-isotropic, the goal is still to maintain the system as close to an
285 > isotropic fluid as possible.  With this in mind, we would like the
286 > kinetic energies in the three different directions could become as
287 > close as each other as possible after each scaling.  Simultaneously,
288 > one would also like each scaling as gentle as possible, i.e. ${x,y,z
289 >  \rightarrow 1}$, in order to avoid large perturbation to the system.
290 > To do this, we pick the intersection point which maintains the three
291 > scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
292 > ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
293 >
294 > After the valid scaling parameters are arrived at by solving geometric
295 > intersection problems in $x, y, z$ space in order to obtain cold slab
296 > scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
297 > determine the conjugate hot slab scaling variables.
298 >
299 > \subsection{Introducing shear stress via velocity scaling}
300 > It is also possible to use this method to magnify the random
301 > fluctuations of the average momentum in each of the bins to induce a
302 > momentum flux.  Doing this repeatedly will create a shear stress on
303 > the system which will respond with an easily-measured strain.  The
304 > momentum flux (say along the $x$-direction) may be defined as:
305   \begin{equation}
306   (1-x) P_c^x = j_z(p_x)\Delta t
307   \label{eq:fluxPlane}
308   \end{equation}
309 < The above {\it flux equation} is essentially a plane which is
310 < perpendicular to the x-axis, with its position governed both by a
311 < targetted value, $j_z(p_x)$ as well as the instantaneous value of the
237 < momentum along the x-direction.
309 > This {\it momentum flux plane} is perpendicular to the $x$-axis, with
310 > its position governed both by a target value, $j_z(p_x)$ as well as
311 > the instantaneous value of the momentum along the $x$-direction.
312  
313 < Similarly, to satisfy a momentum flux as well as the conservation
314 < constraints, it is sufficient to determine the points ${x,y,z}$ which
315 < lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
316 < and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
317 < an ellipsoid and a plane in 3-dimensional space.
313 > In order to satisfy a momentum flux as well as the conservation
314 > constraints, we must determine the points ${x,y,z}$ which lie on both
315 > the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
316 > flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
317 > ellipsoid and a plane in 3-dimensional space.
318  
319 < To summarize, by solving respective equation sets, one can determine
320 < possible sets of scaling variables for cold slab. And corresponding
321 < sets of scaling variables for hot slab can be determine as well.
319 > In the case of momentum flux transfer, we also impose another
320 > constraint to set the kinetic energy transfer as zero. In other
321 > words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$.  With
322 > one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
323 > set of quartic equations to the above kinetic energy transfer problem.
324  
325 < The following problem will be choosing an optimal set of scaling
250 < variables among the possible sets. Although this method is inherently
251 < non-isotropic, the goal is still to maintain the system as isotropic
252 < as possible. Under this consideration, one would like the kinetic
253 < energies in different directions could become as close as each other
254 < after each scaling. Simultaneously, one would also like each scaling
255 < as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
256 < large perturbation to the system. Therefore, one approach to obtain the
257 < scaling variables would be constructing an criteria function, with
258 < constraints as above equation sets, and solving the function's minimum
259 < by method like Lagrange multipliers.
260 <
261 < In order to save computation time, we have a different approach to a
262 < relatively good set of scaling variables with much less calculation
263 < than above. Here is the detail of our simplification of the problem.
325 > \section{Computational Details}
326  
327 < In the case of kinetic energy transfer, we impose another constraint
328 < ${x = y}$, into the equation sets. Consequently, there are two
329 < variables left. And now one only needs to solve a set of two {\it
330 <  ellipses equations}. This problem would be transformed into solving
331 < one quartic equation for one of the two variables. There are known
332 < generic methods that solve real roots of quartic equations. Then one
333 < can determine the other variable and obtain sets of scaling
334 < variables. Among these sets, one can apply the above criteria to
335 < choose the best set, while much faster with only a few sets to choose.
327 > We have implemented this methodology in our molecular dynamics code,
328 > OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
329 > after each MD step.  We have tested the method in a variety of
330 > different systems, including homogeneous fluids (Lennard-Jones and
331 > SPC/E water), crystalline solids ({\sc eam}~\cite{PhysRevB.33.7983} and
332 > quantum Sutton-Chen ({\sc q-sc})~\cite{PhysRevB.59.3527}
333 > models for Gold), and heterogeneous interfaces (QSC gold - SPC/E
334 > water). The last of these systems would have been difficult to study
335 > using previous RNEMD methods, but using velocity scaling moves, we can
336 > even obtain estimates of the interfacial thermal conductivities ($G$).
337  
338 < In the case of momentum flux transfer, we impose another constraint to
276 < set the kinetic energy transfer as zero. In another word, we apply
277 < Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
278 < variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
279 < of equations on the above kinetic energy transfer problem. Therefore,
280 < an approach similar to the above would be sufficient for this as well.
338 > \subsection{Simulation Cells}
339  
340 < \section{Computational Details}
341 < Our simulation consists of a series of systems. All of these
342 < simulations were run with the OpenMD simulation software
343 < package\cite{Meineke:2005gd} integrated with RNEMD methods.
340 > In each of the systems studied, the dynamics was carried out in a
341 > rectangular simulation cell using periodic boundary conditions in all
342 > three dimensions.  The cells were longer along the $z$ axis and the
343 > space was divided into $N$ slabs along this axis (typically $N=20$).
344 > The top slab ($n=1$) was designated the ``cold'' slab, while the
345 > central slab ($n= N/2 + 1$) was designated the ``hot'' slab.  In all
346 > cases, simulations were first thermalized in canonical ensemble (NVT)
347 > using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
348 > microcanonical ensemble (NVE) before introducing any non-equilibrium
349 > method.
350  
351 < A Lennard-Jones fluid system was built and tested first. In order to
288 < compare our method with swapping RNEMD, a series of simulations were
289 < performed to calculate the shear viscosity and thermal conductivity of
290 < argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
291 <  \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
292 < ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
293 < comparison between our results and others. These simulations used
294 < velocity Verlet algorithm with reduced timestep ${\tau^* =
295 <  4.6\times10^{-4}}$.
351 > \subsection{RNEMD with M\"{u}ller-Plathe swaps}
352  
353 < For shear viscosity calculation, the reduced temperature was ${T^* =
354 <  k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
355 < ensemble (NVT), then equilibrated in microcanonical ensemble
300 < (NVE). Establishing and stablizing momentum gradient were followed
301 < also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
302 < adopted.\cite{ISI:000080382700030} The simulation box was under
303 < periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
304 < the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
305 < most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
306 < to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
307 < frequency were chosen. According to each result from swapping
308 < RNEMD, scaling RNEMD simulations were run with the target momentum
309 < flux set to produce a similar momentum flux and shear
310 < rate. Furthermore, various scaling frequencies can be tested for one
311 < single swapping rate. To compare the performance between swapping and
312 < scaling method, temperatures of different dimensions in all the slabs
313 < were observed. Most of the simulations include $10^5$ steps of
314 < equilibration without imposing momentum flux, $10^5$ steps of
315 < stablization with imposing momentum transfer, and $10^6$ steps of data
316 < collection under RNEMD. For relatively high momentum flux simulations,
317 < ${5\times10^5}$ step data collection is sufficient. For some low momentum
318 < flux simulations, ${2\times10^6}$ steps were necessary.
353 > In order to compare our new methodology with the original
354 > M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
355 > first performed simulations using the original technique.
356  
357 < After each simulation, the shear viscosity was calculated in reduced
358 < unit. The momentum flux was calculated with total unphysical
359 < transferred momentum ${P_x}$ and data collection time $t$:
357 > \subsection{RNEMD with NIVS scaling}
358 >
359 > For each simulation utilizing the swapping method, a corresponding
360 > NIVS-RNEMD simulation was carried out using a target momentum flux set
361 > to produce a the same momentum or energy flux exhibited in the
362 > swapping simulation.
363 >
364 > To test the temperature homogeneity (and to compute transport
365 > coefficients), directional momentum and temperature distributions were
366 > accumulated for molecules in each of the slabs.
367 >
368 > \subsection{Shear viscosities}
369 >
370 > The momentum flux was calculated using the total non-physical momentum
371 > transferred (${P_x}$) and the data collection time ($t$):
372   \begin{equation}
373   j_z(p_x) = \frac{P_x}{2 t L_x L_y}
374   \end{equation}
375 < And the velocity gradient ${\langle \partial v_x /\partial z \rangle}$
376 < can be obtained by a linear regression of the velocity profile. From
377 < the shear viscosity $\eta$ calculated with the above parameters, one
378 < can further convert it into reduced unit ${\eta^* = \eta \sigma^2
379 <  (\varepsilon  m)^{-1/2}}$.
375 > where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
376 > box.  The factor of two in the denominator is present because physical
377 > momentum transfer occurs in two directions due to our periodic
378 > boundary conditions.  The velocity gradient ${\langle \partial v_x
379 >  /\partial z \rangle}$ was obtained using linear regression of the
380 > velocity profiles in the bins.  For Lennard-Jones simulations, shear
381 > viscosities are reporte in reduced units (${\eta^* = \eta \sigma^2
382 >  (\varepsilon m)^{-1/2}}$).
383  
384 < For thermal conductivity calculation, simulations were first run under
333 < reduced temperature ${T^* = 0.72}$ in NVE ensemble. Muller-Plathe's
334 < algorithm was adopted in the swapping method. Under identical
335 < simulation box parameters, in each swap, the top slab exchange the
336 < molecule with least kinetic energy with the molecule in the center
337 < slab with most kinetic energy, unless this ``coldest'' molecule in the
338 < ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the ``cold''
339 < slab. According to swapping RNEMD results, target energy flux for
340 < scaling RNEMD simulations can be set. Also, various scaling
341 < frequencies can be tested for one target energy flux. To compare the
342 < performance between swapping and scaling method, distributions of
343 < velocity and speed in different slabs were observed.
384 > \subsection{Thermal Conductivities}
385  
386 < For each swapping rate, thermal conductivity was calculated in reduced
387 < unit. The energy flux was calculated similarly to the momentum flux,
388 < with total unphysical transferred energy ${E_{total}}$ and data collection
348 < time $t$:
386 > The energy flux was calculated similarly to the momentum flux, using
387 > the total non-physical energy transferred (${E_{total}}$) and the data
388 > collection time $t$:
389   \begin{equation}
390   J_z = \frac{E_{total}}{2 t L_x L_y}
391   \end{equation}
392 < And the temperature gradient ${\langle\partial T/\partial z\rangle}$
393 < can be obtained by a linear regression of the temperature
394 < profile. From the thermal conductivity $\lambda$ calculated, one can
395 < further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
396 <  m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
392 > The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
393 > obtained by a linear regression of the temperature profile. For
394 > Lennard-Jones simulations, thermal conductivities are reported in
395 > reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
396 >  k_B^{-1}\varepsilon^{-1/2}}$).
397  
398 < Another series of our simulation is to calculate the interfacial
359 < thermal conductivity of a Au/H${_2}$O system. Respective calculations of
360 < water (SPC/E) and gold (QSC) thermal conductivity were performed and
361 < compared with current results to ensure the validity of
362 < NIVS-RNEMD. After that, the mixture system was simulated.
398 > \subsection{Interfacial Thermal Conductivities}
399  
400 < \section{Results And Discussion}
401 < \subsection{Shear Viscosity}
402 < Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
403 < produced comparable shear viscosity to swap RNEMD method. In Table
404 < \ref{shearRate}, the names of the calculated samples are devided into
405 < two parts. The first number refers to total slabs in one simulation
370 < box. The second number refers to the swapping interval in swap method, or
371 < in scale method the equilvalent swapping interval that the same
372 < momentum flux would theoretically result in swap method. All the scale
373 < method results were from simulations that had a scaling interval of 10
374 < time steps. The average molecular momentum gradients of these samples
375 < are shown in Figure \ref{shearGrad}.
400 > For materials with a relatively low interfacial conductance, and in
401 > cases where the flux between the materials is small, the bulk regions
402 > on either side of an interface rapidly come to a state in which the
403 > two phases have relatively homogeneous (but distinct) temperatures.
404 > In calculating the interfacial thermal conductivity $G$, this
405 > assumption was made, and the conductance can be approximated as:
406  
407 + \begin{equation}
408 + G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
409 +    \langle T_{water}\rangle \right)}
410 + \label{interfaceCalc}
411 + \end{equation}
412 + where ${E_{total}}$ is the imposed non-physical kinetic energy
413 + transfer and ${\langle T_{gold}\rangle}$ and ${\langle
414 +  T_{water}\rangle}$ are the average observed temperature of gold and
415 + water phases respectively.
416 +
417 + \section{Results}
418 +
419 + \subsection{Lennard-Jones Fluid}
420 + 2592 Lennard-Jones atoms were placed in an orthorhombic cell
421 + ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side.  The
422 + reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
423 + direct comparison between our results and previous methods.  These
424 + simulations were carried out with a reduced timestep ${\tau^* =
425 +  4.6\times10^{-4}}$.  For the shear viscosity calculations, the mean
426 + temperature was ${T^* = k_B T/\varepsilon = 0.72}$.  For thermal
427 + conductivity calculations, simulations were first run under reduced
428 + temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
429 + ensemble, but other temperatures ([XXX, YYY, and ZZZ]) were also
430 + sampled.  The simulations included $10^5$ steps of equilibration
431 + without any momentum flux, $10^5$ steps of stablization with an
432 + imposed momentum transfer to create a gradient, and $10^6$ steps of
433 + data collection under RNEMD.
434 +
435 + \subsubsection*{Thermal Conductivity}
436 +
437 + Our thermal conductivity calculations show that the NIVS method agrees
438 + well with the swapping method. Four different swap intervals were
439 + tested (Table \ref{thermalLJRes}). With a fixed 10 fs [WHY NOT REDUCED
440 + UNITS???]  scaling interval, the target exchange kinetic energy
441 + produced equivalent kinetic energy flux as in the swapping method.
442 + Similar thermal gradients were observed with similar thermal flux
443 + under the two different methods (Figure \ref{thermalGrad}).
444 +
445   \begin{table*}
446 < \begin{minipage}{\linewidth}
447 < \begin{center}
446 >  \begin{minipage}{\linewidth}
447 >    \begin{center}
448  
449 < \caption{Calculation results for shear viscosity of Lennard-Jones
450 <  fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
451 <  methods at various momentum exchange rates. Results in reduced
452 <  unit. Errors of calculations in parentheses. }
449 >      \caption{Thermal conductivity (in reduced units) of a
450 >        Lennard-Jones fluid at ${\langle T^* \rangle = 0.72}$ and
451 >        ${\rho^* = 0.85}$ for the swapping and scaling methods at
452 >        various kinetic energy exchange rates. Uncertainties are
453 >        indicated in parentheses.}
454 >      
455 >      \begin{tabular}{|cc|cc|}
456 >        \hline
457 >        \multicolumn{2}{|c|}{Swapping RNEMD} &
458 >        \multicolumn{2}{|c|}{NIVS-RNEMD} \\
459 >        \hline
460 >        Swap Interval (fs) & $\lambda^*_{swap}$ & Equilvalent $J_z^*$ & $\lambda^*_{scale}$\\
461 >        \hline
462 >        250  & 7.03(0.34) & 0.16  & 7.30(0.10)\\
463 >        500  & 7.03(0.14) & 0.09  & 6.95(0.09)\\
464 >        1000 & 6.91(0.42) & 0.047 & 7.19(0.07)\\
465 >        2000 & 7.52(0.15) & 0.024 & 7.19(0.28)\\
466 >        \hline
467 >      \end{tabular}
468 >      \label{thermalLJRes}
469 >    \end{center}
470 >  \end{minipage}
471 > \end{table*}
472  
473 < \begin{tabular}{ccc}
474 < \hline
475 < Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
476 < \hline
477 < 20-500 & 3.64(0.05) & 3.76(0.09)\\
478 < 20-1000 & 3.52(0.16) & 3.66(0.06)\\
479 < 20-2000 & 3.72(0.05) & 3.32(0.18)\\
393 < 20-2500 & 3.42(0.06) & 3.43(0.08)\\
394 < \hline
395 < \end{tabular}
396 < \label{shearRate}
397 < \end{center}
398 < \end{minipage}
399 < \end{table*}
473 > \begin{figure}
474 > \includegraphics[width=\linewidth]{thermalGrad}
475 > \caption{NIVS-RNEMD method creates similar temperature gradients
476 >  compared with the swapping method under a variety of imposed kinetic
477 >  energy flux values.}
478 > \label{thermalGrad}
479 > \end{figure}
480  
481 + During these simulations, velocities were recorded every 1000 steps
482 + and was used to produce distributions of both velocity and speed in
483 + each of the slabs. From these distributions, we observed that under
484 + relatively high non-physical kinetic energy flux, the spee of
485 + molecules in slabs where swapping occured could deviate from the
486 + Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
487 + and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
488 + distributions deviate from an ideal distribution. In the ``hot'' slab,
489 + the probability density is notched at low speeds and has a substantial
490 + shoulder at higher speeds relative to the ideal MB distribution.  In
491 + the cold slab, the opposite notching and shouldering occurs.  This
492 + phenomenon is more obvious at higher swapping rates.  
493 +
494 + In the velocity distributions, the ideal Gaussian peak is
495 + substantially flattened in the hot slab, and is overly sharp (with
496 + truncated wings) in the cold slab. This problem is rooted in the
497 + mechanism of the swapping method. Continually depleting low (high)
498 + speed particles in the high (low) temperature slab is not complemented
499 + by diffusions of low (high) speed particles from neighboring slabs,
500 + unless the swapping rate is sufficiently small. Simutaneously, surplus
501 + low speed particles in the low temperature slab do not have sufficient
502 + time to diffuse to neighboring slabs.  Since the thermal exchange rate
503 + must reach a minimum level to produce an observable thermal gradient,
504 + the swapping-method RNEMD has a relatively narrow choice of exchange
505 + times that can be utilized.
506 +
507 + For comparison, NIVS-RNEMD produces a speed distribution closer to the
508 + Maxwell-Boltzmann curve (Figure \ref{thermalHist}).  The reason for
509 + this is simple; upon velocity scaling, a Gaussian distribution remains
510 + Gaussian.  Although a single scaling move is non-isotropic in three
511 + dimensions, our criteria for choosing a set of scaling coefficients
512 + helps maintain the distributions as close to isotropic as possible.
513 + Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
514 + as the previous RNEMD methods but without large perturbations to the
515 + velocity distributions in the two slabs.
516 +
517   \begin{figure}
518 < \includegraphics[width=\linewidth]{shearGrad}
519 < \caption{Average momentum gradients of shear viscosity simulations}
520 < \label{shearGrad}
518 > \includegraphics[width=\linewidth]{thermalHist}
519 > \caption{Speed distribution for thermal conductivity using a)
520 >  ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
521 >  simulations with an exchange or equilvalent exchange interval of 250
522 >  fs. In circled areas, distributions from ``swapping'' RNEMD
523 >  simulation have deviation from ideal Maxwell-Boltzmann distribution
524 >  (curves fit for each distribution).}
525 > \label{thermalHist}
526   \end{figure}
527  
528 +
529 + \subsubsection*{Shear Viscosity}
530 + Our calculations (Table \ref{shearRate}) show that velocity-scaling
531 + RNEMD predicted comparable shear viscosities to swap RNEMD method.  All
532 + the scale method results were from simulations that had a scaling
533 + interval of 10 time steps. The average molecular momentum gradients of
534 + these samples are shown in Figure \ref{shear} (a) and (b).
535 +
536 + \begin{table*}
537 +  \begin{minipage}{\linewidth}
538 +    \begin{center}
539 +      
540 +      \caption{Shear viscosities of Lennard-Jones fluid at ${T^* =
541 +          0.72}$ and ${\rho^* = 0.85}$ using swapping and NIVS methods
542 +        at various momentum exchange rates.  Uncertainties are
543 +        indicated in parentheses. }
544 +      
545 +      \begin{tabular}{ccccc}
546 +        Swapping method & & & NIVS-RNEMD & \\
547 +        \hline
548 +        Swap Interval (fs) & $\eta^*_{swap}$ & & Equilvalent $j_p^*(v_x)$ &
549 +        $\eta^*_{scale}$\\
550 +        \hline
551 +        500  & 3.64(0.05) & & 0.09  & 3.76(0.09)\\
552 +        1000 & 3.52(0.16) & & 0.046 & 3.66(0.06)\\
553 +        2000 & 3.72(0.05) & & 0.024 & 3.32(0.18)\\
554 +        2500 & 3.42(0.06) & & 0.019 & 3.43(0.08)\\
555 +        \hline
556 +      \end{tabular}
557 +      \label{shearRate}
558 +    \end{center}
559 +  \end{minipage}
560 + \end{table*}
561 +
562   \begin{figure}
563 < \includegraphics[width=\linewidth]{shearTempScale}
564 < \caption{Temperature profile for scaling RNEMD simulation.}
565 < \label{shearTempScale}
563 >  \includegraphics[width=\linewidth]{shear}
564 >  \caption{Average momentum gradients in shear viscosity simulations,
565 >    using (a) ``swapping'' method and (b) NIVS-RNEMD method
566 >    respectively. (c) Temperature difference among x and y, z dimensions
567 >    observed when using NIVS-RNEMD with equivalent exchange interval of
568 >    500 fs.}
569 >  \label{shear}
570   \end{figure}
571 +
572   However, observations of temperatures along three dimensions show that
573   inhomogeneity occurs in scaling RNEMD simulations, particularly in the
574 < two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
574 > two slabs which were scaled. Figure \ref{shear} (c) indicate that with
575   relatively large imposed momentum flux, the temperature difference among $x$
576   and the other two dimensions was significant. This would result from the
577   algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
# Line 431 | Line 591 | attractive than swapping RNEMD in shear viscosity calc
591   exchange momentum flux simulations makes scaling RNEMD method less
592   attractive than swapping RNEMD in shear viscosity calculation.
593  
434 \subsection{Thermal Conductivity}
594  
595 < Our thermal conductivity calculations also show that scaling method results
437 < agree with swapping method. Table \ref{thermal} lists our simulation
438 < results with similar manner we used in shear viscosity
439 < calculation. All the data reported from scaling method were obtained
440 < by simulations of 10-step exchange frequency, and the target exchange
441 < kinetic energy were set to produce equivalent kinetic energy flux as
442 < in swapping method. Figure \ref{thermalGrad} exhibits similar thermal
443 < gradients of respective similar kinetic energy flux.
595 > \subsection{Bulk SPC/E water}
596  
597 < \begin{table*}
598 < \begin{minipage}{\linewidth}
599 < \begin{center}
597 > We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
598 > to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
599 > the original swapping RNEMD method.  Bedrov {\it et
600 >  al.}\cite{Bedrov:2000} argued that exchange of the molecule
601 > center-of-mass velocities instead of single atom velocities in a
602 > molecule conserves the total kinetic energy and linear momentum.  This
603 > principle is also adopted in our simulations. Scaling was applied to
604 > the center-of-mass velocities of the rigid bodies of SPC/E model water
605 > molecules.
606  
607 < \caption{Calculation results for thermal conductivity of Lennard-Jones
608 <  fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
609 <  swap and scale methods at various kinetic energy exchange rates. Results
610 <  in reduced unit. Errors of calculations in parentheses.}
607 > To construct the simulations, a simulation box consisting of 1000
608 > molecules were first equilibrated under ambient pressure and
609 > temperature conditions using the isobaric-isothermal (NPT)
610 > ensemble.\cite{melchionna93} A fixed volume was chosen to match the
611 > average volume observed in the NPT simulations, and this was followed
612 > by equilibration, first in the canonical (NVT) ensemble, followed by a
613 > [XXX ps] period under constant-NVE conditions without any momentum
614 > flux.  [YYY ps] was allowed to stabilize the system with an imposed
615 > momentum transfer to create a gradient, and [ZZZ ps] was alotted for
616 > data collection under RNEMD.
617  
618 < \begin{tabular}{ccc}
619 < \hline
620 < Series & $\lambda^*_{swap}$ & $\lambda^*_{scale}$\\
621 < \hline
622 < 20-250  & 7.03(0.34) & 7.30(0.10)\\
623 < 20-500  & 7.03(0.14) & 6.95(0.09)\\
624 < 20-1000 & 6.91(0.42) & 7.19(0.07)\\
461 < 20-2000 & 7.52(0.15) & 7.19(0.28)\\
462 < \hline
463 < \end{tabular}
464 < \label{thermal}
465 < \end{center}
466 < \end{minipage}
467 < \end{table*}
618 > As shown in Figure \ref{spceGrad}, temperature gradients were
619 > established similar to the previous work.  However, the average
620 > temperature of our system is 300K, while that in Bedrov {\it et al.}
621 > is 318K, which would be attributed for part of the difference between
622 > the final calculation results (Table \ref{spceThermal}). [WHY DIDN'T
623 > WE DO 318 K?]  Both methods yield values in reasonable agreement with
624 > experiment [DONE AT WHAT TEMPERATURE?]
625  
626   \begin{figure}
627 < \includegraphics[width=\linewidth]{thermalGrad}
628 < \caption{Temperature gradients of thermal conductivity simulations}
629 < \label{thermalGrad}
627 >  \includegraphics[width=\linewidth]{spceGrad}
628 >  \caption{Temperature gradients in SPC/E water thermal conductivity
629 >    simulations.}
630 >  \label{spceGrad}
631   \end{figure}
632  
633 < During these simulations, molecule velocities were recorded in 1000 of
634 < all the snapshots. These velocity data were used to produce histograms
635 < of velocity and speed distribution in different slabs. From these
636 < histograms, it is observed that with increasing unphysical kinetic
637 < energy flux, speed and velocity distribution of molecules in slabs
638 < where swapping occured could deviate from Maxwell-Boltzmann
639 < distribution. Figure \ref{histSwap} indicates how these distributions
640 < deviate from ideal condition. In high temperature slabs, probability
641 < density in low speed is confidently smaller than ideal distribution;
642 < in low temperature slabs, probability density in high speed is smaller
643 < than ideal. This phenomenon is observable even in our relatively low
644 < swapping rate simulations. And this deviation could also leads to
645 < deviation of distribution of velocity in various dimensions. One
646 < feature of these deviated distribution is that in high temperature
647 < slab, the ideal Gaussian peak was changed into a relatively flat
648 < plateau; while in low temperature slab, that peak appears sharper.
633 > \begin{table*}
634 >  \begin{minipage}{\linewidth}
635 >    \begin{center}
636 >      
637 >      \caption{Thermal conductivity of SPC/E water under various
638 >        imposed thermal gradients. Uncertainties are indicated in
639 >        parentheses.}
640 >      
641 >      \begin{tabular}{|c|ccc|}
642 >        \hline
643 >        $\langle dT/dz\rangle$(K/\AA) & \multicolumn{3}{|c|}{$\lambda
644 >          (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
645 >        & This work (300K) & Previous simulations (318K)\cite{Bedrov:2000} &
646 >        Experiment\cite{WagnerKruse}\\
647 >        \hline
648 >        0.38 & 0.816(0.044) & & 0.64\\
649 >        0.81 & 0.770(0.008) & 0.784 & \\
650 >        1.54 & 0.813(0.007) & 0.730 & \\
651 >        \hline
652 >      \end{tabular}
653 >      \label{spceThermal}
654 >    \end{center}
655 >  \end{minipage}
656 > \end{table*}
657  
658 < \begin{figure}
493 < \includegraphics[width=\linewidth]{histSwap}
494 < \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
495 < \label{histSwap}
496 < \end{figure}
658 > \subsection{Crystalline Gold}
659  
660 < \begin{figure}
661 < \includegraphics[width=\linewidth]{histScale}
662 < \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
663 < \label{histScale}
664 < \end{figure}
660 > To see how the method performed in a solid, we calculated thermal
661 > conductivities using two atomistic models for gold.  Several different
662 > potential models have been developed that reasonably describe
663 > interactions in transition metals. In particular, the Embedded Atom
664 > Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc
665 >  sc})~\cite{Chen90} potential have been used to study a wide range of
666 > phenomena in both bulk materials and
667 > nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
668 > Both potentials are based on a model of a metal which treats the
669 > nuclei and core electrons as pseudo-atoms embedded in the electron
670 > density due to the valence electrons on all of the other atoms in the
671 > system. The {\sc sc} potential has a simple form that closely
672 > resembles the Lennard Jones potential,
673 > \begin{equation}
674 > \label{eq:SCP1}
675 > 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] ,
676 > \end{equation}
677 > where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
678 > \begin{equation}
679 > \label{eq:SCP2}
680 > 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}}.
681 > \end{equation}
682 > $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
683 > interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
684 > Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
685 > the interactions between the valence electrons and the cores of the
686 > pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
687 > scale, $c_i$ scales the attractive portion of the potential relative
688 > to the repulsive interaction and $\alpha_{ij}$ is a length parameter
689 > that assures a dimensionless form for $\rho$. These parameters are
690 > tuned to various experimental properties such as the density, cohesive
691 > energy, and elastic moduli for FCC transition metals. The quantum
692 > Sutton-Chen ({\sc q-sc}) formulation matches these properties while
693 > including zero-point quantum corrections for different transition
694 > metals.\cite{PhysRevB.59.3527}  The {\sc eam} functional forms differ
695 > slightly from {\sc sc} but the overall method is very similar.
696  
697 < \subsection{Interfaciel Thermal Conductivity}
697 > In this work, we have utilized both the {\sc eam} and the {\sc q-sc}
698 > potentials to test the behavior of scaling RNEMD.
699  
700 < \begin{figure}
701 < \includegraphics[width=\linewidth]{spceGrad}
702 < \caption{Temperature gradients for SPC/E water thermal conductivity.}
703 < \label{spceGrad}
704 < \end{figure}
700 > A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
701 > atoms.  [LxMxN UNIT CELLS].  Simulations were run both with and
702 > without isobaric-isothermal (NPT)~\cite{melchionna93}
703 > pre-equilibration at a target pressure of 1 atm. When equilibrated
704 > under NPT conditions, our simulation box expanded by approximately 1\%
705 > [IN VOLUME OR LINEAR DIMENSIONS ?].  Following adjustment of the box
706 > volume, equilibrations in both the canonical and microcanonical
707 > ensembles were carried out. With the simulation cell divided evenly
708 > into 10 slabs, different thermal gradients were established by
709 > applying a set of target thermal transfer fluxes.
710  
711 + The results for the thermal conductivity of gold are shown in Table
712 + \ref{AuThermal}.  In these calculations, the end and middle slabs were
713 + excluded in thermal gradient linear regession. {\sc eam} predicts
714 + slightly larger thermal conductivities than {\sc q-sc}.  However, both
715 + values are smaller than experimental value by a factor of more than
716 + 200. This behavior has been observed previously by Richardson and
717 + Clancy, and has been attributed to the lack of electronic effects in
718 + these force fields.\cite{Clancy:1992} The non-equilibrium MD method
719 + employed in their simulations produced comparable results to ours.  It
720 + should be noted that the density of the metal being simulated also
721 + greatly affects the thermal conductivity.  With an expanded lattice,
722 + lower thermal conductance is expected (and observed). We also observed
723 + a decrease in thermal conductance at higher temperatures, a trend that
724 + agrees with experimental measurements [PAGE
725 + NUMBERS?].\cite{AshcroftMermin}
726 +
727   \begin{table*}
728 < \begin{minipage}{\linewidth}
729 < \begin{center}
728 >  \begin{minipage}{\linewidth}
729 >    \begin{center}
730 >      
731 >      \caption{Calculated thermal conductivity of crystalline gold
732 >        using two related force fields. Calculations were done at both
733 >        experimental and equilibrated densities and at a range of
734 >        temperatures and thermal flux rates.  Uncertainties are
735 >        indicated in parentheses. [CLANCY COMPARISON? SWAPPING
736 >        COMPARISON?]}
737 >      
738 >      \begin{tabular}{|c|c|c|cc|}
739 >        \hline
740 >        Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
741 >        $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
742 >        \hline
743 >        \multirow{7}{*}{\sc q-sc} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
744 >        &        &     & 2.86 & 1.08(0.05)\\
745 >        &        &     & 5.14 & 1.15(0.07)\\\cline{2-5}
746 >        & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
747 >        &        &     & 3.02 & 1.26(0.05)\\\cline{3-5}
748 >        &        & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
749 >        &        &     & 4.84 & 0.92(0.05)\\
750 >        \hline
751 >        \multirow{8}{*}{\sc eam} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
752 >        &        &     & 2.06 & 1.37(0.04)\\
753 >        &        &     & 2.55 & 1.41(0.07)\\\cline{2-5}
754 >        & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
755 >        &        &     & 2.04 & 1.41(0.07)\\
756 >        &        &     & 2.41 & 1.53(0.10)\\\cline{3-5}
757 >        &        & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
758 >        &        &     & 4.14 & 1.08(0.05)\\
759 >        \hline
760 >      \end{tabular}
761 >      \label{AuThermal}
762 >    \end{center}
763 >  \end{minipage}
764 > \end{table*}
765  
766 < \caption{Calculation results for thermal conductivity of SPC/E water
767 <  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
768 <  calculations in parentheses. }
766 > \subsection{Thermal Conductance at the Au/H$_2$O interface}
767 > The most attractive aspect of the scaling approach for RNEMD is the
768 > ability to use the method in non-homogeneous systems, where molecules
769 > of different identities are segregated in different slabs.  To test
770 > this application, we simulated a Gold (111) / water interface.  To
771 > construct the interface, a box containing a lattice of 1188 Au atoms
772 > (with the 111 surface in the +z and -z directions) was allowed to
773 > relax under ambient temperature and pressure.  A separate (but
774 > identically sized) box of SPC/E water was also equilibrated at ambient
775 > conditions.  The two boxes were combined by removing all water
776 > molecules withing 3 \AA radius of any gold atom.  The final
777 > configuration contained 1862 SPC/E water molecules.
778  
779 < \begin{tabular}{cccc}
780 < \hline
781 < $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
782 < & This work & Previous simulations$^a$ & Experiment$^b$\\
783 < \hline
784 < 0.3 & 0.82()  & 0.784 & 0.64\\
785 < 0.8 & 0.770() & 0.730\\
786 < 1.5 & 0.813() & \\
787 < \hline
529 < \end{tabular}
530 < \label{spceThermal}
531 < \end{center}
532 < \end{minipage}
533 < \end{table*}
779 > After simulations of bulk water and crystal gold, a mixture system was
780 > constructed, consisting of 1188 Au atoms and 1862 H$_2$O
781 > molecules. Spohr potential was adopted in depicting the interaction
782 > between metal atom and water molecule.\cite{ISI:000167766600035} A
783 > similar protocol of equilibration was followed. Several thermal
784 > gradients was built under different target thermal flux. It was found
785 > out that compared to our previous simulation systems, the two phases
786 > could have large temperature difference even under a relatively low
787 > thermal flux.
788  
789  
790 + After simulations of homogeneous water and gold systems using
791 + NIVS-RNEMD method were proved valid, calculation of gold/water
792 + interfacial thermal conductivity was followed. It is found out that
793 + the low interfacial conductance is probably due to the hydrophobic
794 + surface in our system. Figure \ref{interface} (a) demonstrates mass
795 + density change along $z$-axis, which is perpendicular to the
796 + gold/water interface. It is observed that water density significantly
797 + decreases when approaching the surface. Under this low thermal
798 + conductance, both gold and water phase have sufficient time to
799 + eliminate temperature difference inside respectively (Figure
800 + \ref{interface} b). With indistinguishable temperature difference
801 + within respective phase, it is valid to assume that the temperature
802 + difference between gold and water on surface would be approximately
803 + the same as the difference between the gold and water phase. This
804 + assumption enables convenient calculation of $G$ using
805 + Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
806 + layer of water and gold close enough to surface, which would have
807 + greater fluctuation and lower accuracy. Reported results (Table
808 + \ref{interfaceRes}) are of two orders of magnitude smaller than our
809 + calculations on homogeneous systems, and thus have larger relative
810 + errors than our calculation results on homogeneous systems.
811 +
812   \begin{figure}
813 < \includegraphics[width=\linewidth]{AuGrad}
814 < \caption{Temperature gradients for crystal gold thermal conductivity.}
815 < \label{AuGrad}
813 > \includegraphics[width=\linewidth]{interface}
814 > \caption{Simulation results for Gold/Water interfacial thermal
815 >  conductivity: (a) Significant water density decrease is observed on
816 >  crystalline gold surface, which indicates low surface contact and
817 >  leads to low thermal conductance. (b) Temperature profiles for a
818 >  series of simulations. Temperatures of different slabs in the same
819 >  phase show no significant differences.}
820 > \label{interface}
821   \end{figure}
822  
823   \begin{table*}
824   \begin{minipage}{\linewidth}
825   \begin{center}
826  
827 < \caption{Calculation results for thermal conductivity of crystal gold
828 <  at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
829 <  calculations in parentheses. }
827 > \caption{Calculation results for interfacial thermal conductivity
828 >  at ${\langle T\rangle \sim}$ 300K at various thermal exchange
829 >  rates. Errors of calculations in parentheses. }
830  
831 < \begin{tabular}{ccc}
831 > \begin{tabular}{cccc}
832   \hline
833 < $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
834 < & This work & Previous simulations$^a$ \\
833 > $J_z$ (MW/m$^2$) & $T_{gold}$ (K) & $T_{water}$ (K) & $G$
834 > (MW/m$^2$/K)\\
835   \hline
836 < 1.4 & 1.10() & \\
837 < 2.8 & 1.08() & \\
838 < 5.1 & 1.15() & \\
836 > 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
837 > 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
838 > 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
839 > 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
840   \hline
841   \end{tabular}
842 < \label{AuThermal}
842 > \label{interfaceRes}
843   \end{center}
844   \end{minipage}
845   \end{table*}
846  
847  
848 + \section{Conclusions}
849 + NIVS-RNEMD simulation method is developed and tested on various
850 + systems. Simulation results demonstrate its validity in thermal
851 + conductivity calculations, from Lennard-Jones fluid to multi-atom
852 + molecule like water and metal crystals. NIVS-RNEMD improves
853 + non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
854 + methods. Furthermore, it develops a valid means for unphysical thermal
855 + transfer between different species of molecules, and thus extends its
856 + applicability to interfacial systems. Our calculation of gold/water
857 + interfacial thermal conductivity demonstrates this advantage over
858 + previous RNEMD methods. NIVS-RNEMD has also limited application on
859 + shear viscosity calculations, but could cause temperature difference
860 + among different dimensions under high momentum flux. Modification is
861 + necessary to extend the applicability of NIVS-RNEMD in shear viscosity
862 + calculations.
863 +
864   \section{Acknowledgments}
865   Support for this project was provided by the National Science
866   Foundation under grant CHE-0848243. Computational time was provided by
867   the Center for Research Computing (CRC) at the University of Notre
868   Dame.  \newpage
869  
870 < \bibliographystyle{jcp2}
870 > \bibliographystyle{aip}
871   \bibliography{nivsRnemd}
872 +
873   \end{doublespace}
874   \end{document}
875  

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines