ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3588
Committed: Wed Apr 14 22:02:41 2010 UTC (14 years, 5 months ago) by skuang
Content type: application/x-tex
File size: 37292 byte(s)
Log Message:
more citations.

File Contents

# User Rev Content
1 gezelter 3524 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5     \usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{tabularx}
8     \usepackage{graphicx}
9     %\usepackage{booktabs}
10     %\usepackage{bibentry}
11     %\usepackage{mathrsfs}
12     \usepackage[ref]{overcite}
13     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
14     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
15     9.0in \textwidth 6.5in \brokenpenalty=10000
16    
17     % double space list of tables and figures
18     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
19     \setlength{\abovecaptionskip}{20 pt}
20     \setlength{\belowcaptionskip}{30 pt}
21    
22     \renewcommand\citemid{\ } % no comma in optional referenc note
23    
24     \begin{document}
25    
26     \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
27    
28     \author{Shenyu Kuang and J. Daniel
29     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
30     Department of Chemistry and Biochemistry,\\
31     University of Notre Dame\\
32     Notre Dame, Indiana 46556}
33    
34     \date{\today}
35    
36     \maketitle
37    
38     \begin{doublespace}
39    
40     \begin{abstract}
41 gezelter 3583 We present a new method for introducing stable non-equilibrium
42     velocity and temperature distributions in molecular dynamics
43     simulations of heterogeneous systems. This method extends some
44     earlier Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods
45     which use momentum exchange swapping moves that can create
46     non-thermal velocity distributions (and which are difficult to use
47     for interfacial calculations). By using non-isotropic velocity
48     scaling (NIVS) on the molecules in specific regions of a system, it
49     is possible to impose momentum or thermal flux between regions of a
50     simulation and stable thermal and momentum gradients can then be
51     established. The scaling method we have developed conserves the
52     total linear momentum and total energy of the system. To test the
53     methods, we have computed the thermal conductivity of model liquid
54     and solid systems as well as the interfacial thermal conductivity of
55     a metal-water interface. We find that the NIVS-RNEMD improves the
56     problematic velocity distributions that develop in other RNEMD
57     methods.
58 gezelter 3524 \end{abstract}
59    
60     \newpage
61    
62     %\narrowtext
63    
64     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65     % BODY OF TEXT
66     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67    
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 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 gezelter 3583 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 gezelter 3524 \begin{eqnarray}
84 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
86 gezelter 3524 \end{eqnarray}
87 skuang 3528 RNEMD has been widely used to provide computational estimates of thermal
88 gezelter 3524 conductivities and shear viscosities in a wide range of materials,
89     from liquid copper to monatomic liquids to molecular fluids
90 skuang 3587 (e.g. ionic 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}
91 gezelter 3524
92 skuang 3574 \begin{figure}
93     \includegraphics[width=\linewidth]{thermalDemo}
94 gezelter 3583 \caption{RNEMD methods impose an unphysical transfer of momentum or
95     kinetic energy between a ``hot'' slab and a ``cold'' slab in the
96     simulation box. The molecular system responds to this imposed flux
97     by generating a momentum or temperature gradient. The slope of the
98     gradient can then be used to compute transport properties (e.g.
99     shear viscosity and thermal conductivity).}
100 skuang 3574 \label{thermalDemo}
101     \end{figure}
102    
103 skuang 3588 RNEMD is preferable in many ways to the forward NEMD methods\cite{ISI:A1988Q205300014} because it imposes what is typically difficult to measure
104 gezelter 3583 (a flux or stress) and it is typically much easier to compute momentum
105     gradients or strains (the response). For similar reasons, RNEMD is
106     also preferable to slowly-converging equilibrium methods for measuring
107     thermal conductivity and shear viscosity (using Green-Kubo relations
108     [CITATIONS NEEDED] or the Helfand moment approach of Viscardy {\it et
109 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
110 gezelter 3524 computing difficult to measure quantities.
111    
112     Another attractive feature of RNEMD is that it conserves both total
113     linear momentum and total energy during the swaps (as long as the two
114     molecules have the same identity), so the swapped configurations are
115     typically samples from the same manifold of states in the
116     microcanonical ensemble.
117    
118 skuang 3588 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
119 skuang 3565 some problems with the original RNEMD swap technique. Notably, large
120     momentum fluxes (equivalent to frequent momentum swaps between the
121 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
122     momentum distributions in the two slabs, as well as non-linear thermal
123     and velocity distributions along the direction of the imposed flux
124     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
125     and self-adjusting metrics for retaining the usability of the method.
126 gezelter 3524
127     In this paper, we develop and test a method for non-isotropic velocity
128     scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
129     (conservation of linear momentum and total energy, compatibility with
130     periodic boundary conditions) while establishing true thermal
131     distributions in each of the two slabs. In the next section, we
132 gezelter 3583 present the method for determining the scaling constraints. We then
133 gezelter 3524 test the method on both single component, multi-component, and
134     non-isotropic mixtures and show that it is capable of providing
135     reasonable estimates of the thermal conductivity and shear viscosity
136     in these cases.
137    
138     \section{Methodology}
139 gezelter 3583 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
140     periodic system is partitioned into a series of thin slabs along one
141 gezelter 3524 axis ($z$). One of the slabs at the end of the periodic box is
142     designated the ``hot'' slab, while the slab in the center of the box
143     is designated the ``cold'' slab. The artificial momentum flux will be
144     established by transferring momentum from the cold slab and into the
145     hot slab.
146    
147     Rather than using momentum swaps, we use a series of velocity scaling
148 gezelter 3583 moves. For molecules $\{i\}$ located within the cold slab,
149 gezelter 3524 \begin{equation}
150 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
151     x & 0 & 0 \\
152     0 & y & 0 \\
153     0 & 0 & z \\
154 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
155     \end{equation}
156     where ${x, y, z}$ are a set of 3 scaling variables for each of the
157     three directions in the system. Likewise, the molecules $\{j\}$
158 skuang 3528 located in the hot slab will see a concomitant scaling of velocities,
159 gezelter 3524 \begin{equation}
160 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
161     x^\prime & 0 & 0 \\
162     0 & y^\prime & 0 \\
163     0 & 0 & z^\prime \\
164 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
165     \end{equation}
166    
167     Conservation of linear momentum in each of the three directions
168 gezelter 3583 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
169 gezelter 3524 parameters together:
170     \begin{equation}
171 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
172 gezelter 3524 \end{equation}
173     where
174 skuang 3565 \begin{eqnarray}
175 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
176 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
177 gezelter 3524 \label{eq:momentumdef}
178 skuang 3565 \end{eqnarray}
179 skuang 3528 Therefore, for each of the three directions, the hot scaling
180     parameters are a simple function of the cold scaling parameters and
181 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
182     \begin{equation}
183 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
184 gezelter 3524 \label{eq:hotcoldscaling}
185     \end{equation}
186 skuang 3528 where
187     \begin{equation}
188     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
189     \end{equation}
190     for convenience.
191 gezelter 3524
192     Conservation of total energy also places constraints on the scaling:
193     \begin{equation}
194     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
195 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
196 gezelter 3524 \end{equation}
197 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
198     $K_c^\alpha$, are computed for each of the three directions in a
199     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
200     Substituting in the expressions for the hot scaling parameters
201     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
202 gezelter 3583 {\it constraint ellipsoid}:
203 gezelter 3524 \begin{equation}
204 skuang 3565 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
205 gezelter 3524 \label{eq:constraintEllipsoid}
206     \end{equation}
207     where the constants are obtained from the instantaneous values of the
208     linear momenta and kinetic energies for the hot and cold slabs,
209 skuang 3565 \begin{eqnarray}
210 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
211     \left(p_\alpha\right)^2\right) \\
212     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
213 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
214 gezelter 3524 \label{eq:constraintEllipsoidConsts}
215 skuang 3565 \end{eqnarray}
216 gezelter 3583 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
217     cold slab scaling parameters which can be applied while preserving
218     both linear momentum in all three directions as well as total kinetic
219     energy.
220 gezelter 3524
221     The goal of using velocity scaling variables is to transfer linear
222     momentum or kinetic energy from the cold slab to the hot slab. If the
223     hot and cold slabs are separated along the z-axis, the energy flux is
224 skuang 3528 given simply by the decrease in kinetic energy of the cold bin:
225 gezelter 3524 \begin{equation}
226 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
227 gezelter 3524 \end{equation}
228     The expression for the energy flux can be re-written as another
229     ellipsoid centered on $(x,y,z) = 0$:
230     \begin{equation}
231 skuang 3534 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
232 gezelter 3524 \label{eq:fluxEllipsoid}
233     \end{equation}
234 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
235     both by a targetted value, $J_z$ as well as the instantaneous values
236     of the kinetic energy components in the cold bin.
237 gezelter 3524
238     To satisfy an energetic flux as well as the conservation constraints,
239 gezelter 3583 we must determine the points ${x,y,z}$ which lie on both the
240     constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
241     ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
242     two ellipsoids in 3-dimensional space.
243 gezelter 3524
244 gezelter 3569 \begin{figure}
245     \includegraphics[width=\linewidth]{ellipsoids}
246     \caption{Scaling points which maintain both constant energy and
247     constant linear momentum of the system lie on the surface of the
248     {\it constraint ellipsoid} while points which generate the target
249     momentum flux lie on the surface of the {\it flux ellipsoid}. The
250 skuang 3575 velocity distributions in the cold bin are scaled by only those
251 gezelter 3569 points which lie on both ellipsoids.}
252     \label{ellipsoids}
253     \end{figure}
254    
255 gezelter 3583 One may also define {\it momentum} flux (say along the $x$-direction) as:
256 gezelter 3524 \begin{equation}
257 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
258 skuang 3531 \label{eq:fluxPlane}
259 gezelter 3524 \end{equation}
260 gezelter 3583 The above {\it momentum flux plane} is perpendicular to the $x$-axis,
261     with its position governed both by a target value, $j_z(p_x)$ as well
262     as the instantaneous value of the momentum along the $x$-direction.
263 gezelter 3524
264 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
265     constraints, we must determine the points ${x,y,z}$ which lie on both
266     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
267     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
268     ellipsoid and a plane in 3-dimensional space.
269 gezelter 3524
270 gezelter 3583 In both the momentum and energy flux scenarios, valid scaling
271     parameters are arrived at by solving geometric intersection problems
272     in $x, y, z$ space in order to obtain cold slab scaling parameters.
273     Once the scaling variables for the cold slab are known, the hot slab
274     scaling has also been determined.
275 gezelter 3524
276 gezelter 3583
277 skuang 3531 The following problem will be choosing an optimal set of scaling
278     variables among the possible sets. Although this method is inherently
279     non-isotropic, the goal is still to maintain the system as isotropic
280     as possible. Under this consideration, one would like the kinetic
281     energies in different directions could become as close as each other
282     after each scaling. Simultaneously, one would also like each scaling
283     as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
284 gezelter 3583 large perturbation to the system. Therefore, one approach to obtain
285     the scaling variables would be constructing an criteria function, with
286 skuang 3531 constraints as above equation sets, and solving the function's minimum
287     by method like Lagrange multipliers.
288 gezelter 3524
289 skuang 3531 In order to save computation time, we have a different approach to a
290     relatively good set of scaling variables with much less calculation
291     than above. Here is the detail of our simplification of the problem.
292 gezelter 3524
293 skuang 3531 In the case of kinetic energy transfer, we impose another constraint
294     ${x = y}$, into the equation sets. Consequently, there are two
295     variables left. And now one only needs to solve a set of two {\it
296     ellipses equations}. This problem would be transformed into solving
297     one quartic equation for one of the two variables. There are known
298     generic methods that solve real roots of quartic equations. Then one
299     can determine the other variable and obtain sets of scaling
300     variables. Among these sets, one can apply the above criteria to
301     choose the best set, while much faster with only a few sets to choose.
302    
303     In the case of momentum flux transfer, we impose another constraint to
304     set the kinetic energy transfer as zero. In another word, we apply
305     Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
306     variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
307     of equations on the above kinetic energy transfer problem. Therefore,
308     an approach similar to the above would be sufficient for this as well.
309    
310     \section{Computational Details}
311 skuang 3576 \subsection{Lennard-Jones Fluid}
312 skuang 3534 Our simulation consists of a series of systems. All of these
313 skuang 3565 simulations were run with the OpenMD simulation software
314 skuang 3576 package\cite{Meineke:2005gd} integrated with RNEMD codes.
315 skuang 3531
316 skuang 3532 A Lennard-Jones fluid system was built and tested first. In order to
317     compare our method with swapping RNEMD, a series of simulations were
318     performed to calculate the shear viscosity and thermal conductivity of
319 skuang 3534 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
320     \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
321     ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
322     comparison between our results and others. These simulations used
323 skuang 3565 velocity Verlet algorithm with reduced timestep ${\tau^* =
324 skuang 3534 4.6\times10^{-4}}$.
325 skuang 3532
326     For shear viscosity calculation, the reduced temperature was ${T^* =
327 skuang 3565 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
328     ensemble (NVT), then equilibrated in microcanonical ensemble
329     (NVE). Establishing and stablizing momentum gradient were followed
330     also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
331 skuang 3532 adopted.\cite{ISI:000080382700030} The simulation box was under
332 skuang 3534 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
333     the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
334     most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
335 skuang 3565 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
336 skuang 3534 frequency were chosen. According to each result from swapping
337 skuang 3532 RNEMD, scaling RNEMD simulations were run with the target momentum
338 skuang 3576 flux set to produce a similar momentum flux, and consequently shear
339 skuang 3534 rate. Furthermore, various scaling frequencies can be tested for one
340 skuang 3576 single swapping rate. To test the temperature homogeneity in our
341     system of swapping and scaling methods, temperatures of different
342     dimensions in all the slabs were observed. Most of the simulations
343     include $10^5$ steps of equilibration without imposing momentum flux,
344     $10^5$ steps of stablization with imposing unphysical momentum
345     transfer, and $10^6$ steps of data collection under RNEMD. For
346     relatively high momentum flux simulations, ${5\times10^5}$ step data
347     collection is sufficient. For some low momentum flux simulations,
348     ${2\times10^6}$ steps were necessary.
349 skuang 3532
350 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
351     unit. The momentum flux was calculated with total unphysical
352 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
353 skuang 3534 \begin{equation}
354     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
355     \end{equation}
356 skuang 3576 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
357     box, and physical momentum transfer occurs in two ways due to our
358     periodic boundary condition settings. And the velocity gradient
359     ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
360     linear regression of the velocity profile. From the shear viscosity
361     $\eta$ calculated with the above parameters, one can further convert
362     it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
363 skuang 3532
364 skuang 3576 For thermal conductivity calculations, simulations were first run under
365     reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
366     ensemble. Muller-Plathe's algorithm was adopted in the swapping
367     method. Under identical simulation box parameters with our shear
368     viscosity calculations, in each swap, the top slab exchanges all three
369     translational momentum components of the molecule with least kinetic
370     energy with the same components of the molecule in the center slab
371     with most kinetic energy, unless this ``coldest'' molecule in the
372     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
373     ``cold'' slab. According to swapping RNEMD results, target energy flux
374     for scaling RNEMD simulations can be set. Also, various scaling
375 skuang 3534 frequencies can be tested for one target energy flux. To compare the
376     performance between swapping and scaling method, distributions of
377     velocity and speed in different slabs were observed.
378    
379     For each swapping rate, thermal conductivity was calculated in reduced
380     unit. The energy flux was calculated similarly to the momentum flux,
381 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
382 skuang 3534 time $t$:
383     \begin{equation}
384     J_z = \frac{E_{total}}{2 t L_x L_y}
385     \end{equation}
386     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
387     can be obtained by a linear regression of the temperature
388     profile. From the thermal conductivity $\lambda$ calculated, one can
389     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
390     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
391    
392 skuang 3576 \subsection{ Water / Metal Thermal Conductivity}
393     Another series of our simulation is the calculation of interfacial
394 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
395 skuang 3579 liquid water (Extended Simple Point Charge model) and crystal gold
396 skuang 3580 thermal conductivity were performed and compared with current results
397     to ensure the validity of NIVS-RNEMD. After that, a mixture system was
398     simulated.
399 skuang 3563
400 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
401     consisting of 1000 molecules were first equilibrated under ambient
402 skuang 3576 pressure and temperature conditions using NPT ensemble, followed by
403     equilibration in fixed volume (NVT). The system was then equilibrated in
404     microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
405 skuang 3573 stable thermal gradient was followed. The simulation box was under
406     periodic boundary condition and devided into 10 slabs. Data collection
407 skuang 3576 process was similar to Lennard-Jones fluid system.
408 skuang 3573
409 skuang 3576 Thermal conductivity calculation of bulk crystal gold used a similar
410 skuang 3580 protocol. Two types of force field parameters, Embedded Atom Method
411     (EAM) and Quantum Sutten-Chen (QSC) force field were used
412     respectively. The face-centered cubic crystal simulation box consists of
413 skuang 3576 2880 Au atoms. The lattice was first allowed volume change to relax
414     under ambient temperature and pressure. Equilibrations in canonical and
415     microcanonical ensemble were followed in order. With the simulation
416     lattice devided evenly into 10 slabs, different thermal gradients were
417     established by applying a set of target thermal transfer flux. Data of
418     the series of thermal gradients was collected for calculation.
419    
420 skuang 3573 After simulations of bulk water and crystal gold, a mixture system was
421     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
422     molecules. Spohr potential was adopted in depicting the interaction
423     between metal atom and water molecule.\cite{ISI:000167766600035} A
424 skuang 3576 similar protocol of equilibration was followed. Several thermal
425     gradients was built under different target thermal flux. It was found
426     out that compared to our previous simulation systems, the two phases
427     could have large temperature difference even under a relatively low
428     thermal flux. Therefore, under our low flux conditions, it is assumed
429 skuang 3573 that the metal and water phases have respectively homogeneous
430 skuang 3576 temperature, excluding the surface regions. In calculating the
431     interfacial thermal conductivity $G$, this assumptioin was applied and
432     thus our formula becomes:
433 skuang 3573
434     \begin{equation}
435     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
436     \langle T_{water}\rangle \right)}
437     \label{interfaceCalc}
438     \end{equation}
439     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
440     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
441     average observed temperature of gold and water phases respectively.
442    
443 skuang 3577 \section{Results And Discussions}
444 skuang 3538 \subsection{Thermal Conductivity}
445 skuang 3573 \subsubsection{Lennard-Jones Fluid}
446 skuang 3577 Our thermal conductivity calculations show that scaling method results
447     agree with swapping method. Four different exchange intervals were
448     tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
449     10fs exchange interval, target exchange kinetic energy was set to
450     produce equivalent kinetic energy flux as in swapping method. And
451     similar thermal gradients were observed with similar thermal flux in
452     two simulation methods (Figure \ref{thermalGrad}).
453 skuang 3538
454 skuang 3563 \begin{table*}
455     \begin{minipage}{\linewidth}
456     \begin{center}
457 skuang 3538
458 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
459 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
460 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
461     in reduced unit. Errors of calculations in parentheses.}
462    
463 skuang 3565 \begin{tabular}{ccc}
464 skuang 3563 \hline
465 skuang 3577 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
466     $\lambda^*_{scale}$\\
467 skuang 3565 \hline
468 skuang 3577 250 & 7.03(0.34) & 7.30(0.10)\\
469     500 & 7.03(0.14) & 6.95(0.09)\\
470     1000 & 6.91(0.42) & 7.19(0.07)\\
471     2000 & 7.52(0.15) & 7.19(0.28)\\
472 skuang 3566 \hline
473 skuang 3563 \end{tabular}
474 skuang 3577 \label{thermalLJRes}
475 skuang 3563 \end{center}
476     \end{minipage}
477     \end{table*}
478    
479     \begin{figure}
480 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
481 skuang 3577 \caption{Temperature gradients under various kinetic energy flux of
482     thermal conductivity simulations}
483 skuang 3567 \label{thermalGrad}
484 skuang 3563 \end{figure}
485    
486     During these simulations, molecule velocities were recorded in 1000 of
487 skuang 3578 all the snapshots of one single data collection process. These
488     velocity data were used to produce histograms of velocity and speed
489     distribution in different slabs. From these histograms, it is observed
490     that under relatively high unphysical kinetic energy flux, speed and
491     velocity distribution of molecules in slabs where swapping occured
492     could deviate from Maxwell-Boltzmann distribution. Figure
493     \ref{histSwap} illustrates how these distributions deviate from an
494     ideal distribution. In high temperature slab, probability density in
495     low speed is confidently smaller than ideal curve fit; in low
496     temperature slab, probability density in high speed is smaller than
497     ideal, while larger than ideal in low speed. This phenomenon is more
498     obvious in our high swapping rate simulations. And this deviation
499     could also leads to deviation of distribution of velocity in various
500     dimensions. One feature of these deviated distribution is that in high
501     temperature slab, the ideal Gaussian peak was changed into a
502     relatively flat plateau; while in low temperature slab, that peak
503     appears sharper. This problem is rooted in the mechanism of the
504     swapping method. Continually depleting low (high) speed particles in
505     the high (low) temperature slab could not be complemented by
506     diffusions of low (high) speed particles from neighbor slabs, unless
507     in suffciently low swapping rate. Simutaneously, surplus low speed
508     particles in the low temperature slab do not have sufficient time to
509     diffuse to neighbor slabs. However, thermal exchange rate should reach
510     a minimum level to produce an observable thermal gradient under noise
511     interference. Consequently, swapping RNEMD has a relatively narrow
512     choice of swapping rate to satisfy these above restrictions.
513 skuang 3563
514     \begin{figure}
515 skuang 3565 \includegraphics[width=\linewidth]{histSwap}
516 skuang 3578 \caption{Speed distribution for thermal conductivity using swapping
517     RNEMD. Shown is from the simulation with 250 fs exchange interval.}
518 skuang 3563 \label{histSwap}
519     \end{figure}
520    
521 skuang 3578 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
522     curve fit (Figure \ref{histScale}). Essentially, after scaling, a
523     Gaussian distribution function would remain Gaussian. Although a
524     single scaling is non-isotropic in all three dimensions, our scaling
525     coefficient criteria could help maintian the scaling region as
526     isotropic as possible. On the other hand, scaling coefficients are
527     preferred to be as close to 1 as possible, which also helps minimize
528     the difference among different dimensions. This is possible if scaling
529     interval and one-time thermal transfer energy are well
530     chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
531     thermal flux as the previous RNEMD method without large perturbation
532     to the distribution of velocity and speed in the exchange regions.
533    
534 skuang 3568 \begin{figure}
535     \includegraphics[width=\linewidth]{histScale}
536 skuang 3578 \caption{Speed distribution for thermal conductivity using scaling
537     RNEMD. Shown is from the simulation with an equilvalent thermal flux
538     as an 250 fs exchange interval swapping simulation.}
539 skuang 3568 \label{histScale}
540     \end{figure}
541    
542 skuang 3573 \subsubsection{SPC/E Water}
543     Our results of SPC/E water thermal conductivity are comparable to
544     Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
545 skuang 3579 previous swapping RNEMD method for their calculation. Bedrov {\it et
546     al.}\cite{ISI:000090151400044} argued that exchange of the molecule
547     center-of-mass velocities instead of single atom velocities in a
548     molecule conserves the total kinetic energy and linear momentum. This
549     principle is adopted in our simulations. Scaling is applied to the
550     velocities of the rigid bodies of SPC/E model water molecules, instead
551     of each hydrogen and oxygen atoms in relevant water molecules. As
552     shown in Figure \ref{spceGrad}, temperature gradients were established
553     similar to their system. However, the average temperature of our
554     system is 300K, while theirs is 318K, which would be attributed for
555     part of the difference between the final calculation results (Table
556     \ref{spceThermal}). Both methods yields values in agreement with
557     experiment. And this shows the applicability of our method to
558     multi-atom molecular system.
559 skuang 3563
560 skuang 3570 \begin{figure}
561     \includegraphics[width=\linewidth]{spceGrad}
562     \caption{Temperature gradients for SPC/E water thermal conductivity.}
563     \label{spceGrad}
564     \end{figure}
565    
566     \begin{table*}
567     \begin{minipage}{\linewidth}
568     \begin{center}
569    
570     \caption{Calculation results for thermal conductivity of SPC/E water
571     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
572     calculations in parentheses. }
573    
574     \begin{tabular}{cccc}
575     \hline
576     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
577 skuang 3573 & This work & Previous simulations\cite{ISI:000090151400044} &
578     Experiment$^a$\\
579 skuang 3570 \hline
580 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
581     0.81 & 0.770(0.008) & 0.784\\
582     1.54 & 0.813(0.007) & 0.730\\
583 skuang 3570 \hline
584     \end{tabular}
585     \label{spceThermal}
586     \end{center}
587     \end{minipage}
588     \end{table*}
589    
590 skuang 3573 \subsubsection{Crystal Gold}
591 skuang 3580 Our results of gold thermal conductivity using two force fields are
592     shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
593     these calculations,the end and middle slabs were excluded in thermal
594     gradient regession and only used as heat source and drain in the
595     systems. Our yielded values using EAM force field are slightly larger
596     than those using QSC force field. However, both series are
597     significantly smaller than experimental value by an order of more than
598     100. It has been verified that this difference is mainly attributed to
599     the lack of electron interaction representation in these force field
600 skuang 3582 parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
601 skuang 3580 force field parameters in their metal thermal conductivity
602     calculations. The Non-Equilibrium MD method they employed in their
603     simulations produced comparable results to ours. As Zhang {\it et
604     al.}\cite{ISI:000231042800044} stated, thermal conductivity values
605     are influenced mainly by force field. Therefore, it is confident to
606     conclude that NIVS-RNEMD is applicable to metal force field system.
607 skuang 3570
608     \begin{figure}
609     \includegraphics[width=\linewidth]{AuGrad}
610 skuang 3580 \caption{Temperature gradients for thermal conductivity calculation of
611     crystal gold using QSC force field.}
612 skuang 3570 \label{AuGrad}
613     \end{figure}
614    
615     \begin{table*}
616     \begin{minipage}{\linewidth}
617     \begin{center}
618    
619     \caption{Calculation results for thermal conductivity of crystal gold
620 skuang 3580 using QSC force field at ${\langle T\rangle}$ = 300K at various
621     thermal exchange rates. Errors of calculations in parentheses. }
622 skuang 3570
623 skuang 3579 \begin{tabular}{cc}
624 skuang 3570 \hline
625     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
626     \hline
627 skuang 3579 1.44 & 1.10(0.01)\\
628     2.86 & 1.08(0.02)\\
629     5.14 & 1.15(0.01)\\
630 skuang 3570 \hline
631     \end{tabular}
632 skuang 3580 \label{qscThermal}
633 skuang 3570 \end{center}
634     \end{minipage}
635     \end{table*}
636    
637 skuang 3580 \begin{figure}
638     \includegraphics[width=\linewidth]{eamGrad}
639     \caption{Temperature gradients for thermal conductivity calculation of
640     crystal gold using EAM force field.}
641     \label{eamGrad}
642     \end{figure}
643    
644     \begin{table*}
645     \begin{minipage}{\linewidth}
646     \begin{center}
647    
648     \caption{Calculation results for thermal conductivity of crystal gold
649     using EAM force field at ${\langle T\rangle}$ = 300K at various
650     thermal exchange rates. Errors of calculations in parentheses. }
651    
652     \begin{tabular}{cc}
653     \hline
654     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
655     \hline
656     1.24 & 1.24(0.06)\\
657     2.06 & 1.37(0.04)\\
658     2.55 & 1.41(0.03)\\
659     \hline
660     \end{tabular}
661     \label{eamThermal}
662     \end{center}
663     \end{minipage}
664     \end{table*}
665    
666    
667 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
668 skuang 3581 After simulations of homogeneous water and gold systems using
669     NIVS-RNEMD method were proved valid, calculation of gold/water
670     interfacial thermal conductivity was followed. It is found out that
671     the low interfacial conductance is probably due to the hydrophobic
672     surface in our system. Figure \ref{interfaceDensity} demonstrates mass
673     density change along $z$-axis, which is perpendicular to the
674     gold/water interface. It is observed that water density significantly
675     decreases when approaching the surface. Under this low thermal
676     conductance, both gold and water phase have sufficient time to
677     eliminate temperature difference inside respectively (Figure
678     \ref{interfaceGrad}). With indistinguishable temperature difference
679     within respective phase, it is valid to assume that the temperature
680     difference between gold and water on surface would be approximately
681     the same as the difference between the gold and water phase. This
682     assumption enables convenient calculation of $G$ using
683     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
684     layer of water and gold close enough to surface, which would have
685     greater fluctuation and lower accuracy. Reported results (Table
686     \ref{interfaceRes}) are of two orders of magnitude smaller than our
687     calculations on homogeneous systems, and thus have larger relative
688     errors than our calculation results on homogeneous systems.
689 skuang 3573
690 skuang 3571 \begin{figure}
691     \includegraphics[width=\linewidth]{interfaceDensity}
692     \caption{Density profile for interfacial thermal conductivity
693 skuang 3581 simulation box. Significant water density decrease is observed on
694     gold surface.}
695 skuang 3571 \label{interfaceDensity}
696     \end{figure}
697    
698 skuang 3572 \begin{figure}
699     \includegraphics[width=\linewidth]{interfaceGrad}
700     \caption{Temperature profiles for interfacial thermal conductivity
701 skuang 3581 simulation box. Temperatures of different slabs in the same phase
702     show no significant difference.}
703 skuang 3572 \label{interfaceGrad}
704     \end{figure}
705    
706     \begin{table*}
707     \begin{minipage}{\linewidth}
708     \begin{center}
709    
710     \caption{Calculation results for interfacial thermal conductivity
711     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
712     rates. Errors of calculations in parentheses. }
713    
714     \begin{tabular}{cccc}
715     \hline
716     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
717     \hline
718 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
719     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
720     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
721     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
722 skuang 3572 \hline
723     \end{tabular}
724 skuang 3574 \label{interfaceRes}
725 skuang 3572 \end{center}
726     \end{minipage}
727     \end{table*}
728    
729 skuang 3576 \subsection{Shear Viscosity}
730     Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
731     produced comparable shear viscosity to swap RNEMD method. In Table
732     \ref{shearRate}, the names of the calculated samples are devided into
733     two parts. The first number refers to total slabs in one simulation
734     box. The second number refers to the swapping interval in swap method, or
735     in scale method the equilvalent swapping interval that the same
736     momentum flux would theoretically result in swap method. All the scale
737     method results were from simulations that had a scaling interval of 10
738     time steps. The average molecular momentum gradients of these samples
739     are shown in Figure \ref{shearGrad}.
740    
741     \begin{table*}
742     \begin{minipage}{\linewidth}
743     \begin{center}
744    
745     \caption{Calculation results for shear viscosity of Lennard-Jones
746     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
747     methods at various momentum exchange rates. Results in reduced
748     unit. Errors of calculations in parentheses. }
749    
750     \begin{tabular}{ccc}
751     \hline
752     Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
753     \hline
754     20-500 & 3.64(0.05) & 3.76(0.09)\\
755     20-1000 & 3.52(0.16) & 3.66(0.06)\\
756     20-2000 & 3.72(0.05) & 3.32(0.18)\\
757     20-2500 & 3.42(0.06) & 3.43(0.08)\\
758     \hline
759     \end{tabular}
760     \label{shearRate}
761     \end{center}
762     \end{minipage}
763     \end{table*}
764    
765     \begin{figure}
766     \includegraphics[width=\linewidth]{shearGrad}
767     \caption{Average momentum gradients of shear viscosity simulations}
768     \label{shearGrad}
769     \end{figure}
770    
771     \begin{figure}
772     \includegraphics[width=\linewidth]{shearTempScale}
773     \caption{Temperature profile for scaling RNEMD simulation.}
774     \label{shearTempScale}
775     \end{figure}
776     However, observations of temperatures along three dimensions show that
777     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
778     two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
779     relatively large imposed momentum flux, the temperature difference among $x$
780     and the other two dimensions was significant. This would result from the
781     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
782     momentum gradient is set up, $P_c^x$ would be roughly stable
783     ($<0$). Consequently, scaling factor $x$ would most probably larger
784     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
785     keep increase after most scaling steps. And if there is not enough time
786     for the kinetic energy to exchange among different dimensions and
787     different slabs, the system would finally build up temperature
788     (kinetic energy) difference among the three dimensions. Also, between
789     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
790     are closer to neighbor slabs. This is due to momentum transfer along
791     $z$ dimension between slabs.
792    
793     Although results between scaling and swapping methods are comparable,
794     the inherent temperature inhomogeneity even in relatively low imposed
795     exchange momentum flux simulations makes scaling RNEMD method less
796     attractive than swapping RNEMD in shear viscosity calculation.
797    
798 skuang 3574 \section{Conclusions}
799     NIVS-RNEMD simulation method is developed and tested on various
800 skuang 3581 systems. Simulation results demonstrate its validity in thermal
801     conductivity calculations, from Lennard-Jones fluid to multi-atom
802     molecule like water and metal crystals. NIVS-RNEMD improves
803     non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
804     methods. Furthermore, it develops a valid means for unphysical thermal
805     transfer between different species of molecules, and thus extends its
806     applicability to interfacial systems. Our calculation of gold/water
807     interfacial thermal conductivity demonstrates this advantage over
808     previous RNEMD methods. NIVS-RNEMD has also limited application on
809     shear viscosity calculations, but could cause temperature difference
810     among different dimensions under high momentum flux. Modification is
811     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
812     calculations.
813 skuang 3572
814 gezelter 3524 \section{Acknowledgments}
815     Support for this project was provided by the National Science
816     Foundation under grant CHE-0848243. Computational time was provided by
817     the Center for Research Computing (CRC) at the University of Notre
818     Dame. \newpage
819    
820 gezelter 3583 \bibliographystyle{aip}
821 gezelter 3524 \bibliography{nivsRnemd}
822 gezelter 3583
823 gezelter 3524 \end{doublespace}
824     \end{document}
825