ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3579
Committed: Tue Mar 23 22:46:17 2010 UTC (14 years, 5 months ago) by skuang
Content type: application/x-tex
File size: 33173 byte(s)
Log Message:
modified spce thermal conductivity discussions

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    
42     \end{abstract}
43    
44     \newpage
45    
46     %\narrowtext
47    
48     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
49     % BODY OF TEXT
50     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51    
52    
53    
54     \section{Introduction}
55     The original formulation of Reverse Non-equilibrium Molecular Dynamics
56     (RNEMD) obtains transport coefficients (thermal conductivity and shear
57     viscosity) in a fluid by imposing an artificial momentum flux between
58     two thin parallel slabs of material that are spatially separated in
59 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
60 skuang 3531 artificial flux is typically created by periodically ``swapping'' either
61 gezelter 3524 the entire momentum vector $\vec{p}$ or single components of this
62     vector ($p_x$) between molecules in each of the two slabs. If the two
63 skuang 3575 slabs are separated along the $z$ coordinate, the imposed flux is either
64 skuang 3532 directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
65 gezelter 3524 simulated system to the imposed momentum flux will typically be a
66 skuang 3575 velocity or thermal gradient (Fig. \ref{thermalDemo}). The transport
67     coefficients (shear viscosity and thermal conductivity) are easily
68     obtained by assuming linear response of the system,
69 gezelter 3524 \begin{eqnarray}
70 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
71 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
72 gezelter 3524 \end{eqnarray}
73 skuang 3528 RNEMD has been widely used to provide computational estimates of thermal
74 gezelter 3524 conductivities and shear viscosities in a wide range of materials,
75     from liquid copper to monatomic liquids to molecular fluids
76 skuang 3528 (e.g. ionic liquids).\cite{ISI:000246190100032}
77 gezelter 3524
78 skuang 3574 \begin{figure}
79     \includegraphics[width=\linewidth]{thermalDemo}
80 skuang 3576 \caption{Demostration of thermal gradient estalished by RNEMD
81     method. Physical thermal flow directs from high temperature region
82     to low temperature region. Unphysical thermal transfer counteracts
83     it and maintains a steady thermal gradient.}
84 skuang 3574 \label{thermalDemo}
85     \end{figure}
86    
87 gezelter 3524 RNEMD is preferable in many ways to the forward NEMD methods because
88     it imposes what is typically difficult to measure (a flux or stress)
89     and it is typically much easier to compute momentum gradients or
90     strains (the response). For similar reasons, RNEMD is also preferable
91     to slowly-converging equilibrium methods for measuring thermal
92     conductivity and shear viscosity (using Green-Kubo relations or the
93     Helfand moment approach of Viscardy {\it et
94 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
95 gezelter 3524 computing difficult to measure quantities.
96    
97     Another attractive feature of RNEMD is that it conserves both total
98     linear momentum and total energy during the swaps (as long as the two
99     molecules have the same identity), so the swapped configurations are
100     typically samples from the same manifold of states in the
101     microcanonical ensemble.
102    
103 skuang 3565 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
104     some problems with the original RNEMD swap technique. Notably, large
105     momentum fluxes (equivalent to frequent momentum swaps between the
106 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
107     momentum distributions in the two slabs, as well as non-linear thermal
108     and velocity distributions along the direction of the imposed flux
109     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
110     and self-adjusting metrics for retaining the usability of the method.
111 gezelter 3524
112     In this paper, we develop and test a method for non-isotropic velocity
113     scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
114     (conservation of linear momentum and total energy, compatibility with
115     periodic boundary conditions) while establishing true thermal
116     distributions in each of the two slabs. In the next section, we
117     develop the method for determining the scaling constraints. We then
118     test the method on both single component, multi-component, and
119     non-isotropic mixtures and show that it is capable of providing
120     reasonable estimates of the thermal conductivity and shear viscosity
121     in these cases.
122    
123     \section{Methodology}
124     We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
125     system is partitioned into a series of thin slabs along a particular
126     axis ($z$). One of the slabs at the end of the periodic box is
127     designated the ``hot'' slab, while the slab in the center of the box
128     is designated the ``cold'' slab. The artificial momentum flux will be
129     established by transferring momentum from the cold slab and into the
130     hot slab.
131    
132     Rather than using momentum swaps, we use a series of velocity scaling
133 skuang 3528 moves. For molecules $\{i\}$ located within the cold slab,
134 gezelter 3524 \begin{equation}
135 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
136     x & 0 & 0 \\
137     0 & y & 0 \\
138     0 & 0 & z \\
139 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
140     \end{equation}
141     where ${x, y, z}$ are a set of 3 scaling variables for each of the
142     three directions in the system. Likewise, the molecules $\{j\}$
143 skuang 3528 located in the hot slab will see a concomitant scaling of velocities,
144 gezelter 3524 \begin{equation}
145 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
146     x^\prime & 0 & 0 \\
147     0 & y^\prime & 0 \\
148     0 & 0 & z^\prime \\
149 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
150     \end{equation}
151    
152     Conservation of linear momentum in each of the three directions
153     ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
154     parameters together:
155     \begin{equation}
156 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
157 gezelter 3524 \end{equation}
158     where
159 skuang 3565 \begin{eqnarray}
160 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
161 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
162 gezelter 3524 \label{eq:momentumdef}
163 skuang 3565 \end{eqnarray}
164 skuang 3528 Therefore, for each of the three directions, the hot scaling
165     parameters are a simple function of the cold scaling parameters and
166 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
167     \begin{equation}
168 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
169 gezelter 3524 \label{eq:hotcoldscaling}
170     \end{equation}
171 skuang 3528 where
172     \begin{equation}
173     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
174     \end{equation}
175     for convenience.
176 gezelter 3524
177     Conservation of total energy also places constraints on the scaling:
178     \begin{equation}
179     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
180 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
181 gezelter 3524 \end{equation}
182 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
183     $K_c^\alpha$, are computed for each of the three directions in a
184     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
185     Substituting in the expressions for the hot scaling parameters
186     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
187     {\it constraint ellipsoid equation}:
188 gezelter 3524 \begin{equation}
189 skuang 3565 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
190 gezelter 3524 \label{eq:constraintEllipsoid}
191     \end{equation}
192     where the constants are obtained from the instantaneous values of the
193     linear momenta and kinetic energies for the hot and cold slabs,
194 skuang 3565 \begin{eqnarray}
195 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
196     \left(p_\alpha\right)^2\right) \\
197     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
198 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
199 gezelter 3524 \label{eq:constraintEllipsoidConsts}
200 skuang 3565 \end{eqnarray}
201 skuang 3528 This ellipsoid equation defines the set of cold slab scaling
202     parameters which can be applied while preserving both linear momentum
203 skuang 3530 in all three directions as well as kinetic energy.
204 gezelter 3524
205     The goal of using velocity scaling variables is to transfer linear
206     momentum or kinetic energy from the cold slab to the hot slab. If the
207     hot and cold slabs are separated along the z-axis, the energy flux is
208 skuang 3528 given simply by the decrease in kinetic energy of the cold bin:
209 gezelter 3524 \begin{equation}
210 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
211 gezelter 3524 \end{equation}
212     The expression for the energy flux can be re-written as another
213     ellipsoid centered on $(x,y,z) = 0$:
214     \begin{equation}
215 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
216 gezelter 3524 \label{eq:fluxEllipsoid}
217     \end{equation}
218 skuang 3575 The spatial extent of the {\it thermal flux ellipsoid equation} is
219     governed both by a targetted value, $J_z$ as well as the instantaneous
220     values of the kinetic energy components in the cold bin.
221 gezelter 3524
222     To satisfy an energetic flux as well as the conservation constraints,
223     it is sufficient to determine the points ${x,y,z}$ which lie on both
224     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
225     flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
226 skuang 3528 the two ellipsoids in 3-dimensional space.
227 gezelter 3524
228 gezelter 3569 \begin{figure}
229     \includegraphics[width=\linewidth]{ellipsoids}
230     \caption{Scaling points which maintain both constant energy and
231     constant linear momentum of the system lie on the surface of the
232     {\it constraint ellipsoid} while points which generate the target
233     momentum flux lie on the surface of the {\it flux ellipsoid}. The
234 skuang 3575 velocity distributions in the cold bin are scaled by only those
235 gezelter 3569 points which lie on both ellipsoids.}
236     \label{ellipsoids}
237     \end{figure}
238    
239 skuang 3575 One may also define momentum flux (say along the $x$-direction) as:
240 gezelter 3524 \begin{equation}
241 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
242 skuang 3531 \label{eq:fluxPlane}
243 gezelter 3524 \end{equation}
244 skuang 3575 The above {\it momentum flux equation} is essentially a plane which is
245     perpendicular to the $x$-axis, with its position governed both by a
246     target value, $j_z(p_x)$ as well as the instantaneous value of the
247     momentum along the $x$-direction.
248 gezelter 3524
249 skuang 3531 Similarly, to satisfy a momentum flux as well as the conservation
250     constraints, it is sufficient to determine the points ${x,y,z}$ which
251     lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
252     and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
253     an ellipsoid and a plane in 3-dimensional space.
254 gezelter 3524
255 skuang 3531 To summarize, by solving respective equation sets, one can determine
256     possible sets of scaling variables for cold slab. And corresponding
257     sets of scaling variables for hot slab can be determine as well.
258 gezelter 3524
259 skuang 3531 The following problem will be choosing an optimal set of scaling
260     variables among the possible sets. Although this method is inherently
261     non-isotropic, the goal is still to maintain the system as isotropic
262     as possible. Under this consideration, one would like the kinetic
263     energies in different directions could become as close as each other
264     after each scaling. Simultaneously, one would also like each scaling
265     as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
266     large perturbation to the system. Therefore, one approach to obtain the
267     scaling variables would be constructing an criteria function, with
268     constraints as above equation sets, and solving the function's minimum
269     by method like Lagrange multipliers.
270 gezelter 3524
271 skuang 3531 In order to save computation time, we have a different approach to a
272     relatively good set of scaling variables with much less calculation
273     than above. Here is the detail of our simplification of the problem.
274 gezelter 3524
275 skuang 3531 In the case of kinetic energy transfer, we impose another constraint
276     ${x = y}$, into the equation sets. Consequently, there are two
277     variables left. And now one only needs to solve a set of two {\it
278     ellipses equations}. This problem would be transformed into solving
279     one quartic equation for one of the two variables. There are known
280     generic methods that solve real roots of quartic equations. Then one
281     can determine the other variable and obtain sets of scaling
282     variables. Among these sets, one can apply the above criteria to
283     choose the best set, while much faster with only a few sets to choose.
284    
285     In the case of momentum flux transfer, we impose another constraint to
286     set the kinetic energy transfer as zero. In another word, we apply
287     Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
288     variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
289     of equations on the above kinetic energy transfer problem. Therefore,
290     an approach similar to the above would be sufficient for this as well.
291    
292     \section{Computational Details}
293 skuang 3576 \subsection{Lennard-Jones Fluid}
294 skuang 3534 Our simulation consists of a series of systems. All of these
295 skuang 3565 simulations were run with the OpenMD simulation software
296 skuang 3576 package\cite{Meineke:2005gd} integrated with RNEMD codes.
297 skuang 3531
298 skuang 3532 A Lennard-Jones fluid system was built and tested first. In order to
299     compare our method with swapping RNEMD, a series of simulations were
300     performed to calculate the shear viscosity and thermal conductivity of
301 skuang 3534 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
302     \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
303     ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
304     comparison between our results and others. These simulations used
305 skuang 3565 velocity Verlet algorithm with reduced timestep ${\tau^* =
306 skuang 3534 4.6\times10^{-4}}$.
307 skuang 3532
308     For shear viscosity calculation, the reduced temperature was ${T^* =
309 skuang 3565 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
310     ensemble (NVT), then equilibrated in microcanonical ensemble
311     (NVE). Establishing and stablizing momentum gradient were followed
312     also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
313 skuang 3532 adopted.\cite{ISI:000080382700030} The simulation box was under
314 skuang 3534 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
315     the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
316     most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
317 skuang 3565 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
318 skuang 3534 frequency were chosen. According to each result from swapping
319 skuang 3532 RNEMD, scaling RNEMD simulations were run with the target momentum
320 skuang 3576 flux set to produce a similar momentum flux, and consequently shear
321 skuang 3534 rate. Furthermore, various scaling frequencies can be tested for one
322 skuang 3576 single swapping rate. To test the temperature homogeneity in our
323     system of swapping and scaling methods, temperatures of different
324     dimensions in all the slabs were observed. Most of the simulations
325     include $10^5$ steps of equilibration without imposing momentum flux,
326     $10^5$ steps of stablization with imposing unphysical momentum
327     transfer, and $10^6$ steps of data collection under RNEMD. For
328     relatively high momentum flux simulations, ${5\times10^5}$ step data
329     collection is sufficient. For some low momentum flux simulations,
330     ${2\times10^6}$ steps were necessary.
331 skuang 3532
332 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
333     unit. The momentum flux was calculated with total unphysical
334 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
335 skuang 3534 \begin{equation}
336     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
337     \end{equation}
338 skuang 3576 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
339     box, and physical momentum transfer occurs in two ways due to our
340     periodic boundary condition settings. And the velocity gradient
341     ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
342     linear regression of the velocity profile. From the shear viscosity
343     $\eta$ calculated with the above parameters, one can further convert
344     it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
345 skuang 3532
346 skuang 3576 For thermal conductivity calculations, simulations were first run under
347     reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
348     ensemble. Muller-Plathe's algorithm was adopted in the swapping
349     method. Under identical simulation box parameters with our shear
350     viscosity calculations, in each swap, the top slab exchanges all three
351     translational momentum components of the molecule with least kinetic
352     energy with the same components of the molecule in the center slab
353     with most kinetic energy, unless this ``coldest'' molecule in the
354     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
355     ``cold'' slab. According to swapping RNEMD results, target energy flux
356     for scaling RNEMD simulations can be set. Also, various scaling
357 skuang 3534 frequencies can be tested for one target energy flux. To compare the
358     performance between swapping and scaling method, distributions of
359     velocity and speed in different slabs were observed.
360    
361     For each swapping rate, thermal conductivity was calculated in reduced
362     unit. The energy flux was calculated similarly to the momentum flux,
363 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
364 skuang 3534 time $t$:
365     \begin{equation}
366     J_z = \frac{E_{total}}{2 t L_x L_y}
367     \end{equation}
368     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
369     can be obtained by a linear regression of the temperature
370     profile. From the thermal conductivity $\lambda$ calculated, one can
371     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
372     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
373    
374 skuang 3576 \subsection{ Water / Metal Thermal Conductivity}
375     Another series of our simulation is the calculation of interfacial
376 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
377 skuang 3579 liquid water (Extended Simple Point Charge model) and crystal gold
378     (Quantum Sutton-Chen potential) thermal conductivity were performed
379     and compared with current results to ensure the validity of
380     NIVS-RNEMD. After that, a mixture system was simulated.
381 skuang 3563
382 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
383     consisting of 1000 molecules were first equilibrated under ambient
384 skuang 3576 pressure and temperature conditions using NPT ensemble, followed by
385     equilibration in fixed volume (NVT). The system was then equilibrated in
386     microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
387 skuang 3573 stable thermal gradient was followed. The simulation box was under
388     periodic boundary condition and devided into 10 slabs. Data collection
389 skuang 3576 process was similar to Lennard-Jones fluid system.
390 skuang 3573
391 skuang 3576 Thermal conductivity calculation of bulk crystal gold used a similar
392 skuang 3579 protocol. The face-centered cubic crystal simulation box consists of
393 skuang 3576 2880 Au atoms. The lattice was first allowed volume change to relax
394     under ambient temperature and pressure. Equilibrations in canonical and
395     microcanonical ensemble were followed in order. With the simulation
396     lattice devided evenly into 10 slabs, different thermal gradients were
397     established by applying a set of target thermal transfer flux. Data of
398     the series of thermal gradients was collected for calculation.
399    
400 skuang 3573 After simulations of bulk water and crystal gold, a mixture system was
401     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
402     molecules. Spohr potential was adopted in depicting the interaction
403     between metal atom and water molecule.\cite{ISI:000167766600035} A
404 skuang 3576 similar protocol of equilibration was followed. Several thermal
405     gradients was built under different target thermal flux. It was found
406     out that compared to our previous simulation systems, the two phases
407     could have large temperature difference even under a relatively low
408     thermal flux. Therefore, under our low flux conditions, it is assumed
409 skuang 3573 that the metal and water phases have respectively homogeneous
410 skuang 3576 temperature, excluding the surface regions. In calculating the
411     interfacial thermal conductivity $G$, this assumptioin was applied and
412     thus our formula becomes:
413 skuang 3573
414     \begin{equation}
415     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
416     \langle T_{water}\rangle \right)}
417     \label{interfaceCalc}
418     \end{equation}
419     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
420     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
421     average observed temperature of gold and water phases respectively.
422    
423 skuang 3577 \section{Results And Discussions}
424 skuang 3538 \subsection{Thermal Conductivity}
425 skuang 3573 \subsubsection{Lennard-Jones Fluid}
426 skuang 3577 Our thermal conductivity calculations show that scaling method results
427     agree with swapping method. Four different exchange intervals were
428     tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
429     10fs exchange interval, target exchange kinetic energy was set to
430     produce equivalent kinetic energy flux as in swapping method. And
431     similar thermal gradients were observed with similar thermal flux in
432     two simulation methods (Figure \ref{thermalGrad}).
433 skuang 3538
434 skuang 3563 \begin{table*}
435     \begin{minipage}{\linewidth}
436     \begin{center}
437 skuang 3538
438 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
439 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
440 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
441     in reduced unit. Errors of calculations in parentheses.}
442    
443 skuang 3565 \begin{tabular}{ccc}
444 skuang 3563 \hline
445 skuang 3577 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
446     $\lambda^*_{scale}$\\
447 skuang 3565 \hline
448 skuang 3577 250 & 7.03(0.34) & 7.30(0.10)\\
449     500 & 7.03(0.14) & 6.95(0.09)\\
450     1000 & 6.91(0.42) & 7.19(0.07)\\
451     2000 & 7.52(0.15) & 7.19(0.28)\\
452 skuang 3566 \hline
453 skuang 3563 \end{tabular}
454 skuang 3577 \label{thermalLJRes}
455 skuang 3563 \end{center}
456     \end{minipage}
457     \end{table*}
458    
459     \begin{figure}
460 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
461 skuang 3577 \caption{Temperature gradients under various kinetic energy flux of
462     thermal conductivity simulations}
463 skuang 3567 \label{thermalGrad}
464 skuang 3563 \end{figure}
465    
466     During these simulations, molecule velocities were recorded in 1000 of
467 skuang 3578 all the snapshots of one single data collection process. These
468     velocity data were used to produce histograms of velocity and speed
469     distribution in different slabs. From these histograms, it is observed
470     that under relatively high unphysical kinetic energy flux, speed and
471     velocity distribution of molecules in slabs where swapping occured
472     could deviate from Maxwell-Boltzmann distribution. Figure
473     \ref{histSwap} illustrates how these distributions deviate from an
474     ideal distribution. In high temperature slab, probability density in
475     low speed is confidently smaller than ideal curve fit; in low
476     temperature slab, probability density in high speed is smaller than
477     ideal, while larger than ideal in low speed. This phenomenon is more
478     obvious in our high swapping rate simulations. And this deviation
479     could also leads to deviation of distribution of velocity in various
480     dimensions. One feature of these deviated distribution is that in high
481     temperature slab, the ideal Gaussian peak was changed into a
482     relatively flat plateau; while in low temperature slab, that peak
483     appears sharper. This problem is rooted in the mechanism of the
484     swapping method. Continually depleting low (high) speed particles in
485     the high (low) temperature slab could not be complemented by
486     diffusions of low (high) speed particles from neighbor slabs, unless
487     in suffciently low swapping rate. Simutaneously, surplus low speed
488     particles in the low temperature slab do not have sufficient time to
489     diffuse to neighbor slabs. However, thermal exchange rate should reach
490     a minimum level to produce an observable thermal gradient under noise
491     interference. Consequently, swapping RNEMD has a relatively narrow
492     choice of swapping rate to satisfy these above restrictions.
493 skuang 3563
494     \begin{figure}
495 skuang 3565 \includegraphics[width=\linewidth]{histSwap}
496 skuang 3578 \caption{Speed distribution for thermal conductivity using swapping
497     RNEMD. Shown is from the simulation with 250 fs exchange interval.}
498 skuang 3563 \label{histSwap}
499     \end{figure}
500    
501 skuang 3578 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
502     curve fit (Figure \ref{histScale}). Essentially, after scaling, a
503     Gaussian distribution function would remain Gaussian. Although a
504     single scaling is non-isotropic in all three dimensions, our scaling
505     coefficient criteria could help maintian the scaling region as
506     isotropic as possible. On the other hand, scaling coefficients are
507     preferred to be as close to 1 as possible, which also helps minimize
508     the difference among different dimensions. This is possible if scaling
509     interval and one-time thermal transfer energy are well
510     chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
511     thermal flux as the previous RNEMD method without large perturbation
512     to the distribution of velocity and speed in the exchange regions.
513    
514 skuang 3568 \begin{figure}
515     \includegraphics[width=\linewidth]{histScale}
516 skuang 3578 \caption{Speed distribution for thermal conductivity using scaling
517     RNEMD. Shown is from the simulation with an equilvalent thermal flux
518     as an 250 fs exchange interval swapping simulation.}
519 skuang 3568 \label{histScale}
520     \end{figure}
521    
522 skuang 3573 \subsubsection{SPC/E Water}
523     Our results of SPC/E water thermal conductivity are comparable to
524     Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
525 skuang 3579 previous swapping RNEMD method for their calculation. Bedrov {\it et
526     al.}\cite{ISI:000090151400044} argued that exchange of the molecule
527     center-of-mass velocities instead of single atom velocities in a
528     molecule conserves the total kinetic energy and linear momentum. This
529     principle is adopted in our simulations. Scaling is applied to the
530     velocities of the rigid bodies of SPC/E model water molecules, instead
531     of each hydrogen and oxygen atoms in relevant water molecules. As
532     shown in Figure \ref{spceGrad}, temperature gradients were established
533     similar to their system. However, the average temperature of our
534     system is 300K, while theirs is 318K, which would be attributed for
535     part of the difference between the final calculation results (Table
536     \ref{spceThermal}). Both methods yields values in agreement with
537     experiment. And this shows the applicability of our method to
538     multi-atom molecular system.
539 skuang 3563
540 skuang 3570 \begin{figure}
541     \includegraphics[width=\linewidth]{spceGrad}
542     \caption{Temperature gradients for SPC/E water thermal conductivity.}
543     \label{spceGrad}
544     \end{figure}
545    
546     \begin{table*}
547     \begin{minipage}{\linewidth}
548     \begin{center}
549    
550     \caption{Calculation results for thermal conductivity of SPC/E water
551     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
552     calculations in parentheses. }
553    
554     \begin{tabular}{cccc}
555     \hline
556     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
557 skuang 3573 & This work & Previous simulations\cite{ISI:000090151400044} &
558     Experiment$^a$\\
559 skuang 3570 \hline
560 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
561     0.81 & 0.770(0.008) & 0.784\\
562     1.54 & 0.813(0.007) & 0.730\\
563 skuang 3570 \hline
564     \end{tabular}
565     \label{spceThermal}
566     \end{center}
567     \end{minipage}
568     \end{table*}
569    
570 skuang 3573 \subsubsection{Crystal Gold}
571 skuang 3579 Our results of gold thermal conductivity using QSC force field are
572 skuang 3574 shown in Table \ref{AuThermal}. Although our calculation is smaller
573     than experimental value by an order of more than 100, this difference
574     is mainly attributed to the lack of electron interaction
575     representation in our force field parameters. Richardson {\it et
576 skuang 3579 al.}\cite{ISI:A1992HX37800010} using similar force field parameters
577 skuang 3574 in their metal thermal conductivity calculations. The EMD method they
578     employed in their simulations produced comparable results to
579     ours. Therefore, it is confident to conclude that NIVS-RNEMD is
580     applicable to metal force field system.
581 skuang 3570
582     \begin{figure}
583     \includegraphics[width=\linewidth]{AuGrad}
584     \caption{Temperature gradients for crystal gold thermal conductivity.}
585     \label{AuGrad}
586     \end{figure}
587    
588     \begin{table*}
589     \begin{minipage}{\linewidth}
590     \begin{center}
591    
592     \caption{Calculation results for thermal conductivity of crystal gold
593     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
594     calculations in parentheses. }
595    
596 skuang 3579 \begin{tabular}{cc}
597 skuang 3570 \hline
598     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
599     \hline
600 skuang 3579 1.44 & 1.10(0.01)\\
601     2.86 & 1.08(0.02)\\
602     5.14 & 1.15(0.01)\\
603 skuang 3570 \hline
604     \end{tabular}
605     \label{AuThermal}
606     \end{center}
607     \end{minipage}
608     \end{table*}
609    
610 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
611 skuang 3574 After valid simulations of homogeneous water and gold systems using
612     NIVS-RNEMD method, calculation of gold/water interfacial thermal
613     conductivity was followed. It is found out that the interfacial
614     conductance is low due to a hydrophobic surface in our system. Figure
615     \ref{interfaceDensity} demonstrates this observance. Consequently, our
616     reported results (Table \ref{interfaceRes}) are of two orders of
617     magnitude smaller than our calculations on homogeneous systems.
618 skuang 3573
619 skuang 3571 \begin{figure}
620     \includegraphics[width=\linewidth]{interfaceDensity}
621     \caption{Density profile for interfacial thermal conductivity
622     simulation box.}
623     \label{interfaceDensity}
624     \end{figure}
625    
626 skuang 3572 \begin{figure}
627     \includegraphics[width=\linewidth]{interfaceGrad}
628     \caption{Temperature profiles for interfacial thermal conductivity
629     simulation box.}
630     \label{interfaceGrad}
631     \end{figure}
632    
633     \begin{table*}
634     \begin{minipage}{\linewidth}
635     \begin{center}
636    
637     \caption{Calculation results for interfacial thermal conductivity
638     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
639     rates. Errors of calculations in parentheses. }
640    
641     \begin{tabular}{cccc}
642     \hline
643     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
644     \hline
645 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
646     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
647     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
648     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
649 skuang 3572 \hline
650     \end{tabular}
651 skuang 3574 \label{interfaceRes}
652 skuang 3572 \end{center}
653     \end{minipage}
654     \end{table*}
655    
656 skuang 3576 \subsection{Shear Viscosity}
657     Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
658     produced comparable shear viscosity to swap RNEMD method. In Table
659     \ref{shearRate}, the names of the calculated samples are devided into
660     two parts. The first number refers to total slabs in one simulation
661     box. The second number refers to the swapping interval in swap method, or
662     in scale method the equilvalent swapping interval that the same
663     momentum flux would theoretically result in swap method. All the scale
664     method results were from simulations that had a scaling interval of 10
665     time steps. The average molecular momentum gradients of these samples
666     are shown in Figure \ref{shearGrad}.
667    
668     \begin{table*}
669     \begin{minipage}{\linewidth}
670     \begin{center}
671    
672     \caption{Calculation results for shear viscosity of Lennard-Jones
673     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
674     methods at various momentum exchange rates. Results in reduced
675     unit. Errors of calculations in parentheses. }
676    
677     \begin{tabular}{ccc}
678     \hline
679     Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
680     \hline
681     20-500 & 3.64(0.05) & 3.76(0.09)\\
682     20-1000 & 3.52(0.16) & 3.66(0.06)\\
683     20-2000 & 3.72(0.05) & 3.32(0.18)\\
684     20-2500 & 3.42(0.06) & 3.43(0.08)\\
685     \hline
686     \end{tabular}
687     \label{shearRate}
688     \end{center}
689     \end{minipage}
690     \end{table*}
691    
692     \begin{figure}
693     \includegraphics[width=\linewidth]{shearGrad}
694     \caption{Average momentum gradients of shear viscosity simulations}
695     \label{shearGrad}
696     \end{figure}
697    
698     \begin{figure}
699     \includegraphics[width=\linewidth]{shearTempScale}
700     \caption{Temperature profile for scaling RNEMD simulation.}
701     \label{shearTempScale}
702     \end{figure}
703     However, observations of temperatures along three dimensions show that
704     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
705     two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
706     relatively large imposed momentum flux, the temperature difference among $x$
707     and the other two dimensions was significant. This would result from the
708     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
709     momentum gradient is set up, $P_c^x$ would be roughly stable
710     ($<0$). Consequently, scaling factor $x$ would most probably larger
711     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
712     keep increase after most scaling steps. And if there is not enough time
713     for the kinetic energy to exchange among different dimensions and
714     different slabs, the system would finally build up temperature
715     (kinetic energy) difference among the three dimensions. Also, between
716     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
717     are closer to neighbor slabs. This is due to momentum transfer along
718     $z$ dimension between slabs.
719    
720     Although results between scaling and swapping methods are comparable,
721     the inherent temperature inhomogeneity even in relatively low imposed
722     exchange momentum flux simulations makes scaling RNEMD method less
723     attractive than swapping RNEMD in shear viscosity calculation.
724    
725 skuang 3574 \section{Conclusions}
726     NIVS-RNEMD simulation method is developed and tested on various
727     systems. Simulation results demonstrate its validity of thermal
728     conductivity calculations. NIVS-RNEMD improves non-Boltzmann-Maxwell
729     distributions existing in previous RNEMD methods, and extends its
730     applicability to interfacial systems. NIVS-RNEMD has also limited
731     application on shear viscosity calculations, but under high momentum
732     flux, it could cause temperature difference among different
733     dimensions. Modification is necessary to extend the applicability of
734     NIVS-RNEMD in shear viscosity calculations.
735 skuang 3572
736 gezelter 3524 \section{Acknowledgments}
737     Support for this project was provided by the National Science
738     Foundation under grant CHE-0848243. Computational time was provided by
739     the Center for Research Computing (CRC) at the University of Notre
740     Dame. \newpage
741    
742     \bibliographystyle{jcp2}
743     \bibliography{nivsRnemd}
744     \end{doublespace}
745     \end{document}
746