ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3582
Committed: Thu Apr 8 21:23:55 2010 UTC (14 years, 4 months ago) by skuang
Content type: application/x-tex
File size: 36618 byte(s)
Log Message:
edited one citation.

File Contents

# Content
1 \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 A molecular simulation method is developed based on the previous RNEMD
42 algorithm. With scaling the velocities of molecules in specific
43 regions of a system, unphysical thermal transfer can be achieved and
44 thermal / momentum gradient can be established, as well as
45 conservation of linear momentum and translational kinetic
46 energy. Thermal conductivity calculations of Lennard-Jones fluid water
47 (SPC/E model) and crystal gold (QSC and EAM model) are performed and
48 demostrate the validity of the method. Furthermore, NIVS-RNEMD
49 improves the non-Maxwell-Boltzmann velocity distributions in previous
50 RNEMD methods, where unphysical momentum transfer occurs. NIVS-RNEMD
51 also extends its application to interfacial thermal conductivity
52 calculations, thanks to its novel means in kinetic energy transfer.
53 \end{abstract}
54
55 \newpage
56
57 %\narrowtext
58
59 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60 % BODY OF TEXT
61 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62
63
64
65 \section{Introduction}
66 The original formulation of Reverse Non-equilibrium Molecular Dynamics
67 (RNEMD) obtains transport coefficients (thermal conductivity and shear
68 viscosity) in a fluid by imposing an artificial momentum flux between
69 two thin parallel slabs of material that are spatially separated in
70 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
71 artificial flux is typically created by periodically ``swapping'' either
72 the entire momentum vector $\vec{p}$ or single components of this
73 vector ($p_x$) between molecules in each of the two slabs. If the two
74 slabs are separated along the $z$ coordinate, the imposed flux is either
75 directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
76 simulated system to the imposed momentum flux will typically be a
77 velocity or thermal gradient (Fig. \ref{thermalDemo}). The transport
78 coefficients (shear viscosity and thermal conductivity) are easily
79 obtained by assuming linear response of the system,
80 \begin{eqnarray}
81 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
82 J_z & = & \lambda \frac{\partial T}{\partial z}
83 \end{eqnarray}
84 RNEMD has been widely used to provide computational estimates of thermal
85 conductivities and shear viscosities in a wide range of materials,
86 from liquid copper to monatomic liquids to molecular fluids
87 (e.g. ionic liquids).\cite{ISI:000246190100032}
88
89 \begin{figure}
90 \includegraphics[width=\linewidth]{thermalDemo}
91 \caption{Demostration of thermal gradient estalished by RNEMD
92 method. Physical thermal flow directs from high temperature region
93 to low temperature region. Unphysical thermal transfer counteracts
94 it and maintains a steady thermal gradient.}
95 \label{thermalDemo}
96 \end{figure}
97
98 RNEMD is preferable in many ways to the forward NEMD methods because
99 it imposes what is typically difficult to measure (a flux or stress)
100 and it is typically much easier to compute momentum gradients or
101 strains (the response). For similar reasons, RNEMD is also preferable
102 to slowly-converging equilibrium methods for measuring thermal
103 conductivity and shear viscosity (using Green-Kubo relations or the
104 Helfand moment approach of Viscardy {\it et
105 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
106 computing difficult to measure quantities.
107
108 Another attractive feature of RNEMD is that it conserves both total
109 linear momentum and total energy during the swaps (as long as the two
110 molecules have the same identity), so the swapped configurations are
111 typically samples from the same manifold of states in the
112 microcanonical ensemble.
113
114 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
115 some problems with the original RNEMD swap technique. Notably, large
116 momentum fluxes (equivalent to frequent momentum swaps between the
117 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
118 momentum distributions in the two slabs, as well as non-linear thermal
119 and velocity distributions along the direction of the imposed flux
120 ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
121 and self-adjusting metrics for retaining the usability of the method.
122
123 In this paper, we develop and test a method for non-isotropic velocity
124 scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
125 (conservation of linear momentum and total energy, compatibility with
126 periodic boundary conditions) while establishing true thermal
127 distributions in each of the two slabs. In the next section, we
128 develop the method for determining the scaling constraints. We then
129 test the method on both single component, multi-component, and
130 non-isotropic mixtures and show that it is capable of providing
131 reasonable estimates of the thermal conductivity and shear viscosity
132 in these cases.
133
134 \section{Methodology}
135 We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
136 system is partitioned into a series of thin slabs along a particular
137 axis ($z$). One of the slabs at the end of the periodic box is
138 designated the ``hot'' slab, while the slab in the center of the box
139 is designated the ``cold'' slab. The artificial momentum flux will be
140 established by transferring momentum from the cold slab and into the
141 hot slab.
142
143 Rather than using momentum swaps, we use a series of velocity scaling
144 moves. For molecules $\{i\}$ located within the cold slab,
145 \begin{equation}
146 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
147 x & 0 & 0 \\
148 0 & y & 0 \\
149 0 & 0 & z \\
150 \end{array} \right) \cdot \vec{v}_i
151 \end{equation}
152 where ${x, y, z}$ are a set of 3 scaling variables for each of the
153 three directions in the system. Likewise, the molecules $\{j\}$
154 located in the hot slab will see a concomitant scaling of velocities,
155 \begin{equation}
156 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
157 x^\prime & 0 & 0 \\
158 0 & y^\prime & 0 \\
159 0 & 0 & z^\prime \\
160 \end{array} \right) \cdot \vec{v}_j
161 \end{equation}
162
163 Conservation of linear momentum in each of the three directions
164 ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
165 parameters together:
166 \begin{equation}
167 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
168 \end{equation}
169 where
170 \begin{eqnarray}
171 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
172 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
173 \label{eq:momentumdef}
174 \end{eqnarray}
175 Therefore, for each of the three directions, the hot scaling
176 parameters are a simple function of the cold scaling parameters and
177 the instantaneous linear momentum in each of the two slabs.
178 \begin{equation}
179 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
180 \label{eq:hotcoldscaling}
181 \end{equation}
182 where
183 \begin{equation}
184 p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
185 \end{equation}
186 for convenience.
187
188 Conservation of total energy also places constraints on the scaling:
189 \begin{equation}
190 \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
191 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
192 \end{equation}
193 where the translational kinetic energies, $K_h^\alpha$ and
194 $K_c^\alpha$, are computed for each of the three directions in a
195 similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
196 Substituting in the expressions for the hot scaling parameters
197 ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
198 {\it constraint ellipsoid equation}:
199 \begin{equation}
200 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
201 \label{eq:constraintEllipsoid}
202 \end{equation}
203 where the constants are obtained from the instantaneous values of the
204 linear momenta and kinetic energies for the hot and cold slabs,
205 \begin{eqnarray}
206 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
207 \left(p_\alpha\right)^2\right) \\
208 b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
209 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
210 \label{eq:constraintEllipsoidConsts}
211 \end{eqnarray}
212 This ellipsoid equation defines the set of cold slab scaling
213 parameters which can be applied while preserving both linear momentum
214 in all three directions as well as kinetic energy.
215
216 The goal of using velocity scaling variables is to transfer linear
217 momentum or kinetic energy from the cold slab to the hot slab. If the
218 hot and cold slabs are separated along the z-axis, the energy flux is
219 given simply by the decrease in kinetic energy of the cold bin:
220 \begin{equation}
221 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
222 \end{equation}
223 The expression for the energy flux can be re-written as another
224 ellipsoid centered on $(x,y,z) = 0$:
225 \begin{equation}
226 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
227 \label{eq:fluxEllipsoid}
228 \end{equation}
229 The spatial extent of the {\it thermal flux ellipsoid equation} is
230 governed both by a targetted value, $J_z$ as well as the instantaneous
231 values of the kinetic energy components in the cold bin.
232
233 To satisfy an energetic flux as well as the conservation constraints,
234 it is sufficient to determine the points ${x,y,z}$ which lie on both
235 the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
236 flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
237 the two ellipsoids in 3-dimensional space.
238
239 \begin{figure}
240 \includegraphics[width=\linewidth]{ellipsoids}
241 \caption{Scaling points which maintain both constant energy and
242 constant linear momentum of the system lie on the surface of the
243 {\it constraint ellipsoid} while points which generate the target
244 momentum flux lie on the surface of the {\it flux ellipsoid}. The
245 velocity distributions in the cold bin are scaled by only those
246 points which lie on both ellipsoids.}
247 \label{ellipsoids}
248 \end{figure}
249
250 One may also define momentum flux (say along the $x$-direction) as:
251 \begin{equation}
252 (1-x) P_c^x = j_z(p_x)\Delta t
253 \label{eq:fluxPlane}
254 \end{equation}
255 The above {\it momentum flux equation} is essentially a plane which is
256 perpendicular to the $x$-axis, with its position governed both by a
257 target value, $j_z(p_x)$ as well as the instantaneous value of the
258 momentum along the $x$-direction.
259
260 Similarly, to satisfy a momentum flux as well as the conservation
261 constraints, it is sufficient to determine the points ${x,y,z}$ which
262 lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
263 and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
264 an ellipsoid and a plane in 3-dimensional space.
265
266 To summarize, by solving respective equation sets, one can determine
267 possible sets of scaling variables for cold slab. And corresponding
268 sets of scaling variables for hot slab can be determine as well.
269
270 The following problem will be choosing an optimal set of scaling
271 variables among the possible sets. Although this method is inherently
272 non-isotropic, the goal is still to maintain the system as isotropic
273 as possible. Under this consideration, one would like the kinetic
274 energies in different directions could become as close as each other
275 after each scaling. Simultaneously, one would also like each scaling
276 as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
277 large perturbation to the system. Therefore, one approach to obtain the
278 scaling variables would be constructing an criteria function, with
279 constraints as above equation sets, and solving the function's minimum
280 by method like Lagrange multipliers.
281
282 In order to save computation time, we have a different approach to a
283 relatively good set of scaling variables with much less calculation
284 than above. Here is the detail of our simplification of the problem.
285
286 In the case of kinetic energy transfer, we impose another constraint
287 ${x = y}$, into the equation sets. Consequently, there are two
288 variables left. And now one only needs to solve a set of two {\it
289 ellipses equations}. This problem would be transformed into solving
290 one quartic equation for one of the two variables. There are known
291 generic methods that solve real roots of quartic equations. Then one
292 can determine the other variable and obtain sets of scaling
293 variables. Among these sets, one can apply the above criteria to
294 choose the best set, while much faster with only a few sets to choose.
295
296 In the case of momentum flux transfer, we impose another constraint to
297 set the kinetic energy transfer as zero. In another word, we apply
298 Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
299 variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
300 of equations on the above kinetic energy transfer problem. Therefore,
301 an approach similar to the above would be sufficient for this as well.
302
303 \section{Computational Details}
304 \subsection{Lennard-Jones Fluid}
305 Our simulation consists of a series of systems. All of these
306 simulations were run with the OpenMD simulation software
307 package\cite{Meineke:2005gd} integrated with RNEMD codes.
308
309 A Lennard-Jones fluid system was built and tested first. In order to
310 compare our method with swapping RNEMD, a series of simulations were
311 performed to calculate the shear viscosity and thermal conductivity of
312 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
313 \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
314 ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
315 comparison between our results and others. These simulations used
316 velocity Verlet algorithm with reduced timestep ${\tau^* =
317 4.6\times10^{-4}}$.
318
319 For shear viscosity calculation, the reduced temperature was ${T^* =
320 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
321 ensemble (NVT), then equilibrated in microcanonical ensemble
322 (NVE). Establishing and stablizing momentum gradient were followed
323 also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
324 adopted.\cite{ISI:000080382700030} The simulation box was under
325 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
326 the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
327 most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
328 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
329 frequency were chosen. According to each result from swapping
330 RNEMD, scaling RNEMD simulations were run with the target momentum
331 flux set to produce a similar momentum flux, and consequently shear
332 rate. Furthermore, various scaling frequencies can be tested for one
333 single swapping rate. To test the temperature homogeneity in our
334 system of swapping and scaling methods, temperatures of different
335 dimensions in all the slabs were observed. Most of the simulations
336 include $10^5$ steps of equilibration without imposing momentum flux,
337 $10^5$ steps of stablization with imposing unphysical momentum
338 transfer, and $10^6$ steps of data collection under RNEMD. For
339 relatively high momentum flux simulations, ${5\times10^5}$ step data
340 collection is sufficient. For some low momentum flux simulations,
341 ${2\times10^6}$ steps were necessary.
342
343 After each simulation, the shear viscosity was calculated in reduced
344 unit. The momentum flux was calculated with total unphysical
345 transferred momentum ${P_x}$ and data collection time $t$:
346 \begin{equation}
347 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
348 \end{equation}
349 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
350 box, and physical momentum transfer occurs in two ways due to our
351 periodic boundary condition settings. And the velocity gradient
352 ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
353 linear regression of the velocity profile. From the shear viscosity
354 $\eta$ calculated with the above parameters, one can further convert
355 it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
356
357 For thermal conductivity calculations, simulations were first run under
358 reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
359 ensemble. Muller-Plathe's algorithm was adopted in the swapping
360 method. Under identical simulation box parameters with our shear
361 viscosity calculations, in each swap, the top slab exchanges all three
362 translational momentum components of the molecule with least kinetic
363 energy with the same components of the molecule in the center slab
364 with most kinetic energy, unless this ``coldest'' molecule in the
365 ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
366 ``cold'' slab. According to swapping RNEMD results, target energy flux
367 for scaling RNEMD simulations can be set. Also, various scaling
368 frequencies can be tested for one target energy flux. To compare the
369 performance between swapping and scaling method, distributions of
370 velocity and speed in different slabs were observed.
371
372 For each swapping rate, thermal conductivity was calculated in reduced
373 unit. The energy flux was calculated similarly to the momentum flux,
374 with total unphysical transferred energy ${E_{total}}$ and data collection
375 time $t$:
376 \begin{equation}
377 J_z = \frac{E_{total}}{2 t L_x L_y}
378 \end{equation}
379 And the temperature gradient ${\langle\partial T/\partial z\rangle}$
380 can be obtained by a linear regression of the temperature
381 profile. From the thermal conductivity $\lambda$ calculated, one can
382 further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
383 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
384
385 \subsection{ Water / Metal Thermal Conductivity}
386 Another series of our simulation is the calculation of interfacial
387 thermal conductivity of a Au/H$_2$O system. Respective calculations of
388 liquid water (Extended Simple Point Charge model) and crystal gold
389 thermal conductivity were performed and compared with current results
390 to ensure the validity of NIVS-RNEMD. After that, a mixture system was
391 simulated.
392
393 For thermal conductivity calculation of bulk water, a simulation box
394 consisting of 1000 molecules were first equilibrated under ambient
395 pressure and temperature conditions using NPT ensemble, followed by
396 equilibration in fixed volume (NVT). The system was then equilibrated in
397 microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
398 stable thermal gradient was followed. The simulation box was under
399 periodic boundary condition and devided into 10 slabs. Data collection
400 process was similar to Lennard-Jones fluid system.
401
402 Thermal conductivity calculation of bulk crystal gold used a similar
403 protocol. Two types of force field parameters, Embedded Atom Method
404 (EAM) and Quantum Sutten-Chen (QSC) force field were used
405 respectively. The face-centered cubic crystal simulation box consists of
406 2880 Au atoms. The lattice was first allowed volume change to relax
407 under ambient temperature and pressure. Equilibrations in canonical and
408 microcanonical ensemble were followed in order. With the simulation
409 lattice devided evenly into 10 slabs, different thermal gradients were
410 established by applying a set of target thermal transfer flux. Data of
411 the series of thermal gradients was collected for calculation.
412
413 After simulations of bulk water and crystal gold, a mixture system was
414 constructed, consisting of 1188 Au atoms and 1862 H$_2$O
415 molecules. Spohr potential was adopted in depicting the interaction
416 between metal atom and water molecule.\cite{ISI:000167766600035} A
417 similar protocol of equilibration was followed. Several thermal
418 gradients was built under different target thermal flux. It was found
419 out that compared to our previous simulation systems, the two phases
420 could have large temperature difference even under a relatively low
421 thermal flux. Therefore, under our low flux conditions, it is assumed
422 that the metal and water phases have respectively homogeneous
423 temperature, excluding the surface regions. In calculating the
424 interfacial thermal conductivity $G$, this assumptioin was applied and
425 thus our formula becomes:
426
427 \begin{equation}
428 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
429 \langle T_{water}\rangle \right)}
430 \label{interfaceCalc}
431 \end{equation}
432 where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
433 and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
434 average observed temperature of gold and water phases respectively.
435
436 \section{Results And Discussions}
437 \subsection{Thermal Conductivity}
438 \subsubsection{Lennard-Jones Fluid}
439 Our thermal conductivity calculations show that scaling method results
440 agree with swapping method. Four different exchange intervals were
441 tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
442 10fs exchange interval, target exchange kinetic energy was set to
443 produce equivalent kinetic energy flux as in swapping method. And
444 similar thermal gradients were observed with similar thermal flux in
445 two simulation methods (Figure \ref{thermalGrad}).
446
447 \begin{table*}
448 \begin{minipage}{\linewidth}
449 \begin{center}
450
451 \caption{Calculation results for thermal conductivity of Lennard-Jones
452 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
453 swap and scale methods at various kinetic energy exchange rates. Results
454 in reduced unit. Errors of calculations in parentheses.}
455
456 \begin{tabular}{ccc}
457 \hline
458 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
459 $\lambda^*_{scale}$\\
460 \hline
461 250 & 7.03(0.34) & 7.30(0.10)\\
462 500 & 7.03(0.14) & 6.95(0.09)\\
463 1000 & 6.91(0.42) & 7.19(0.07)\\
464 2000 & 7.52(0.15) & 7.19(0.28)\\
465 \hline
466 \end{tabular}
467 \label{thermalLJRes}
468 \end{center}
469 \end{minipage}
470 \end{table*}
471
472 \begin{figure}
473 \includegraphics[width=\linewidth]{thermalGrad}
474 \caption{Temperature gradients under various kinetic energy flux of
475 thermal conductivity simulations}
476 \label{thermalGrad}
477 \end{figure}
478
479 During these simulations, molecule velocities were recorded in 1000 of
480 all the snapshots of one single data collection process. These
481 velocity data were used to produce histograms of velocity and speed
482 distribution in different slabs. From these histograms, it is observed
483 that under relatively high unphysical kinetic energy flux, speed and
484 velocity distribution of molecules in slabs where swapping occured
485 could deviate from Maxwell-Boltzmann distribution. Figure
486 \ref{histSwap} illustrates how these distributions deviate from an
487 ideal distribution. In high temperature slab, probability density in
488 low speed is confidently smaller than ideal curve fit; in low
489 temperature slab, probability density in high speed is smaller than
490 ideal, while larger than ideal in low speed. This phenomenon is more
491 obvious in our high swapping rate simulations. And this deviation
492 could also leads to deviation of distribution of velocity in various
493 dimensions. One feature of these deviated distribution is that in high
494 temperature slab, the ideal Gaussian peak was changed into a
495 relatively flat plateau; while in low temperature slab, that peak
496 appears sharper. This problem is rooted in the mechanism of the
497 swapping method. Continually depleting low (high) speed particles in
498 the high (low) temperature slab could not be complemented by
499 diffusions of low (high) speed particles from neighbor slabs, unless
500 in suffciently low swapping rate. Simutaneously, surplus low speed
501 particles in the low temperature slab do not have sufficient time to
502 diffuse to neighbor slabs. However, thermal exchange rate should reach
503 a minimum level to produce an observable thermal gradient under noise
504 interference. Consequently, swapping RNEMD has a relatively narrow
505 choice of swapping rate to satisfy these above restrictions.
506
507 \begin{figure}
508 \includegraphics[width=\linewidth]{histSwap}
509 \caption{Speed distribution for thermal conductivity using swapping
510 RNEMD. Shown is from the simulation with 250 fs exchange interval.}
511 \label{histSwap}
512 \end{figure}
513
514 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
515 curve fit (Figure \ref{histScale}). Essentially, after scaling, a
516 Gaussian distribution function would remain Gaussian. Although a
517 single scaling is non-isotropic in all three dimensions, our scaling
518 coefficient criteria could help maintian the scaling region as
519 isotropic as possible. On the other hand, scaling coefficients are
520 preferred to be as close to 1 as possible, which also helps minimize
521 the difference among different dimensions. This is possible if scaling
522 interval and one-time thermal transfer energy are well
523 chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
524 thermal flux as the previous RNEMD method without large perturbation
525 to the distribution of velocity and speed in the exchange regions.
526
527 \begin{figure}
528 \includegraphics[width=\linewidth]{histScale}
529 \caption{Speed distribution for thermal conductivity using scaling
530 RNEMD. Shown is from the simulation with an equilvalent thermal flux
531 as an 250 fs exchange interval swapping simulation.}
532 \label{histScale}
533 \end{figure}
534
535 \subsubsection{SPC/E Water}
536 Our results of SPC/E water thermal conductivity are comparable to
537 Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
538 previous swapping RNEMD method for their calculation. Bedrov {\it et
539 al.}\cite{ISI:000090151400044} argued that exchange of the molecule
540 center-of-mass velocities instead of single atom velocities in a
541 molecule conserves the total kinetic energy and linear momentum. This
542 principle is adopted in our simulations. Scaling is applied to the
543 velocities of the rigid bodies of SPC/E model water molecules, instead
544 of each hydrogen and oxygen atoms in relevant water molecules. As
545 shown in Figure \ref{spceGrad}, temperature gradients were established
546 similar to their system. However, the average temperature of our
547 system is 300K, while theirs is 318K, which would be attributed for
548 part of the difference between the final calculation results (Table
549 \ref{spceThermal}). Both methods yields values in agreement with
550 experiment. And this shows the applicability of our method to
551 multi-atom molecular system.
552
553 \begin{figure}
554 \includegraphics[width=\linewidth]{spceGrad}
555 \caption{Temperature gradients for SPC/E water thermal conductivity.}
556 \label{spceGrad}
557 \end{figure}
558
559 \begin{table*}
560 \begin{minipage}{\linewidth}
561 \begin{center}
562
563 \caption{Calculation results for thermal conductivity of SPC/E water
564 at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
565 calculations in parentheses. }
566
567 \begin{tabular}{cccc}
568 \hline
569 $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
570 & This work & Previous simulations\cite{ISI:000090151400044} &
571 Experiment$^a$\\
572 \hline
573 0.38 & 0.816(0.044) & & 0.64\\
574 0.81 & 0.770(0.008) & 0.784\\
575 1.54 & 0.813(0.007) & 0.730\\
576 \hline
577 \end{tabular}
578 \label{spceThermal}
579 \end{center}
580 \end{minipage}
581 \end{table*}
582
583 \subsubsection{Crystal Gold}
584 Our results of gold thermal conductivity using two force fields are
585 shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
586 these calculations,the end and middle slabs were excluded in thermal
587 gradient regession and only used as heat source and drain in the
588 systems. Our yielded values using EAM force field are slightly larger
589 than those using QSC force field. However, both series are
590 significantly smaller than experimental value by an order of more than
591 100. It has been verified that this difference is mainly attributed to
592 the lack of electron interaction representation in these force field
593 parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
594 force field parameters in their metal thermal conductivity
595 calculations. The Non-Equilibrium MD method they employed in their
596 simulations produced comparable results to ours. As Zhang {\it et
597 al.}\cite{ISI:000231042800044} stated, thermal conductivity values
598 are influenced mainly by force field. Therefore, it is confident to
599 conclude that NIVS-RNEMD is applicable to metal force field system.
600
601 \begin{figure}
602 \includegraphics[width=\linewidth]{AuGrad}
603 \caption{Temperature gradients for thermal conductivity calculation of
604 crystal gold using QSC force field.}
605 \label{AuGrad}
606 \end{figure}
607
608 \begin{table*}
609 \begin{minipage}{\linewidth}
610 \begin{center}
611
612 \caption{Calculation results for thermal conductivity of crystal gold
613 using QSC force field at ${\langle T\rangle}$ = 300K at various
614 thermal exchange rates. Errors of calculations in parentheses. }
615
616 \begin{tabular}{cc}
617 \hline
618 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
619 \hline
620 1.44 & 1.10(0.01)\\
621 2.86 & 1.08(0.02)\\
622 5.14 & 1.15(0.01)\\
623 \hline
624 \end{tabular}
625 \label{qscThermal}
626 \end{center}
627 \end{minipage}
628 \end{table*}
629
630 \begin{figure}
631 \includegraphics[width=\linewidth]{eamGrad}
632 \caption{Temperature gradients for thermal conductivity calculation of
633 crystal gold using EAM force field.}
634 \label{eamGrad}
635 \end{figure}
636
637 \begin{table*}
638 \begin{minipage}{\linewidth}
639 \begin{center}
640
641 \caption{Calculation results for thermal conductivity of crystal gold
642 using EAM force field at ${\langle T\rangle}$ = 300K at various
643 thermal exchange rates. Errors of calculations in parentheses. }
644
645 \begin{tabular}{cc}
646 \hline
647 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
648 \hline
649 1.24 & 1.24(0.06)\\
650 2.06 & 1.37(0.04)\\
651 2.55 & 1.41(0.03)\\
652 \hline
653 \end{tabular}
654 \label{eamThermal}
655 \end{center}
656 \end{minipage}
657 \end{table*}
658
659
660 \subsection{Interfaciel Thermal Conductivity}
661 After simulations of homogeneous water and gold systems using
662 NIVS-RNEMD method were proved valid, calculation of gold/water
663 interfacial thermal conductivity was followed. It is found out that
664 the low interfacial conductance is probably due to the hydrophobic
665 surface in our system. Figure \ref{interfaceDensity} demonstrates mass
666 density change along $z$-axis, which is perpendicular to the
667 gold/water interface. It is observed that water density significantly
668 decreases when approaching the surface. Under this low thermal
669 conductance, both gold and water phase have sufficient time to
670 eliminate temperature difference inside respectively (Figure
671 \ref{interfaceGrad}). With indistinguishable temperature difference
672 within respective phase, it is valid to assume that the temperature
673 difference between gold and water on surface would be approximately
674 the same as the difference between the gold and water phase. This
675 assumption enables convenient calculation of $G$ using
676 Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
677 layer of water and gold close enough to surface, which would have
678 greater fluctuation and lower accuracy. Reported results (Table
679 \ref{interfaceRes}) are of two orders of magnitude smaller than our
680 calculations on homogeneous systems, and thus have larger relative
681 errors than our calculation results on homogeneous systems.
682
683 \begin{figure}
684 \includegraphics[width=\linewidth]{interfaceDensity}
685 \caption{Density profile for interfacial thermal conductivity
686 simulation box. Significant water density decrease is observed on
687 gold surface.}
688 \label{interfaceDensity}
689 \end{figure}
690
691 \begin{figure}
692 \includegraphics[width=\linewidth]{interfaceGrad}
693 \caption{Temperature profiles for interfacial thermal conductivity
694 simulation box. Temperatures of different slabs in the same phase
695 show no significant difference.}
696 \label{interfaceGrad}
697 \end{figure}
698
699 \begin{table*}
700 \begin{minipage}{\linewidth}
701 \begin{center}
702
703 \caption{Calculation results for interfacial thermal conductivity
704 at ${\langle T\rangle \sim}$ 300K at various thermal exchange
705 rates. Errors of calculations in parentheses. }
706
707 \begin{tabular}{cccc}
708 \hline
709 $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
710 \hline
711 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
712 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
713 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
714 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
715 \hline
716 \end{tabular}
717 \label{interfaceRes}
718 \end{center}
719 \end{minipage}
720 \end{table*}
721
722 \subsection{Shear Viscosity}
723 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
724 produced comparable shear viscosity to swap RNEMD method. In Table
725 \ref{shearRate}, the names of the calculated samples are devided into
726 two parts. The first number refers to total slabs in one simulation
727 box. The second number refers to the swapping interval in swap method, or
728 in scale method the equilvalent swapping interval that the same
729 momentum flux would theoretically result in swap method. All the scale
730 method results were from simulations that had a scaling interval of 10
731 time steps. The average molecular momentum gradients of these samples
732 are shown in Figure \ref{shearGrad}.
733
734 \begin{table*}
735 \begin{minipage}{\linewidth}
736 \begin{center}
737
738 \caption{Calculation results for shear viscosity of Lennard-Jones
739 fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
740 methods at various momentum exchange rates. Results in reduced
741 unit. Errors of calculations in parentheses. }
742
743 \begin{tabular}{ccc}
744 \hline
745 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
746 \hline
747 20-500 & 3.64(0.05) & 3.76(0.09)\\
748 20-1000 & 3.52(0.16) & 3.66(0.06)\\
749 20-2000 & 3.72(0.05) & 3.32(0.18)\\
750 20-2500 & 3.42(0.06) & 3.43(0.08)\\
751 \hline
752 \end{tabular}
753 \label{shearRate}
754 \end{center}
755 \end{minipage}
756 \end{table*}
757
758 \begin{figure}
759 \includegraphics[width=\linewidth]{shearGrad}
760 \caption{Average momentum gradients of shear viscosity simulations}
761 \label{shearGrad}
762 \end{figure}
763
764 \begin{figure}
765 \includegraphics[width=\linewidth]{shearTempScale}
766 \caption{Temperature profile for scaling RNEMD simulation.}
767 \label{shearTempScale}
768 \end{figure}
769 However, observations of temperatures along three dimensions show that
770 inhomogeneity occurs in scaling RNEMD simulations, particularly in the
771 two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
772 relatively large imposed momentum flux, the temperature difference among $x$
773 and the other two dimensions was significant. This would result from the
774 algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
775 momentum gradient is set up, $P_c^x$ would be roughly stable
776 ($<0$). Consequently, scaling factor $x$ would most probably larger
777 than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
778 keep increase after most scaling steps. And if there is not enough time
779 for the kinetic energy to exchange among different dimensions and
780 different slabs, the system would finally build up temperature
781 (kinetic energy) difference among the three dimensions. Also, between
782 $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
783 are closer to neighbor slabs. This is due to momentum transfer along
784 $z$ dimension between slabs.
785
786 Although results between scaling and swapping methods are comparable,
787 the inherent temperature inhomogeneity even in relatively low imposed
788 exchange momentum flux simulations makes scaling RNEMD method less
789 attractive than swapping RNEMD in shear viscosity calculation.
790
791 \section{Conclusions}
792 NIVS-RNEMD simulation method is developed and tested on various
793 systems. Simulation results demonstrate its validity in thermal
794 conductivity calculations, from Lennard-Jones fluid to multi-atom
795 molecule like water and metal crystals. NIVS-RNEMD improves
796 non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
797 methods. Furthermore, it develops a valid means for unphysical thermal
798 transfer between different species of molecules, and thus extends its
799 applicability to interfacial systems. Our calculation of gold/water
800 interfacial thermal conductivity demonstrates this advantage over
801 previous RNEMD methods. NIVS-RNEMD has also limited application on
802 shear viscosity calculations, but could cause temperature difference
803 among different dimensions under high momentum flux. Modification is
804 necessary to extend the applicability of NIVS-RNEMD in shear viscosity
805 calculations.
806
807 \section{Acknowledgments}
808 Support for this project was provided by the National Science
809 Foundation under grant CHE-0848243. Computational time was provided by
810 the Center for Research Computing (CRC) at the University of Notre
811 Dame. \newpage
812
813 \bibliographystyle{jcp2}
814 \bibliography{nivsRnemd}
815 \end{doublespace}
816 \end{document}
817