ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3585
Committed: Wed Apr 14 16:57:52 2010 UTC (14 years, 2 months ago) by skuang
Content type: application/x-tex
File size: 37253 byte(s)
Log Message:
add more citations.

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 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 \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 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 artificial flux is typically created by periodically ``swapping''
75 either the entire momentum vector $\vec{p}$ or single components of
76 this vector ($p_x$) between molecules in each of the two slabs. If
77 the two slabs are separated along the $z$ coordinate, the imposed flux
78 is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
79 response of a simulated system to the imposed momentum flux will
80 typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
81 The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
82 easily obtained by assuming linear response of the system,
83 \begin{eqnarray}
84 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85 J_z & = & \lambda \frac{\partial T}{\partial z}
86 \end{eqnarray}
87 RNEMD has been widely used to provide computational estimates of thermal
88 conductivities and shear viscosities in a wide range of materials,
89 from liquid copper to monatomic liquids to molecular fluids
90 (e.g. ionic liquids).\cite{Bedrov:2000-1,Bedrov:2000,Muller-Plathe:2002,ISI:000184808400018,ISI:000231042800044,Maginn:2007,Muller-Plathe:2008,ISI:000258840700015}
91
92 \begin{figure}
93 \includegraphics[width=\linewidth]{thermalDemo}
94 \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 \label{thermalDemo}
101 \end{figure}
102
103 RNEMD is preferable in many ways to the forward NEMD methods
104 [CITATIONS NEEDED] because it imposes what is typically difficult to measure
105 (a flux or stress) and it is typically much easier to compute momentum
106 gradients or strains (the response). For similar reasons, RNEMD is
107 also preferable to slowly-converging equilibrium methods for measuring
108 thermal conductivity and shear viscosity (using Green-Kubo relations
109 [CITATIONS NEEDED] or the Helfand moment approach of Viscardy {\it et
110 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
111 computing difficult to measure quantities.
112
113 Another attractive feature of RNEMD is that it conserves both total
114 linear momentum and total energy during the swaps (as long as the two
115 molecules have the same identity), so the swapped configurations are
116 typically samples from the same manifold of states in the
117 microcanonical ensemble.
118
119 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
120 some problems with the original RNEMD swap technique. Notably, large
121 momentum fluxes (equivalent to frequent momentum swaps between the
122 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
123 momentum distributions in the two slabs, as well as non-linear thermal
124 and velocity distributions along the direction of the imposed flux
125 ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
126 and self-adjusting metrics for retaining the usability of the method.
127
128 In this paper, we develop and test a method for non-isotropic velocity
129 scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
130 (conservation of linear momentum and total energy, compatibility with
131 periodic boundary conditions) while establishing true thermal
132 distributions in each of the two slabs. In the next section, we
133 present the method for determining the scaling constraints. We then
134 test the method on both single component, multi-component, and
135 non-isotropic mixtures and show that it is capable of providing
136 reasonable estimates of the thermal conductivity and shear viscosity
137 in these cases.
138
139 \section{Methodology}
140 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
141 periodic system is partitioned into a series of thin slabs along one
142 axis ($z$). One of the slabs at the end of the periodic box is
143 designated the ``hot'' slab, while the slab in the center of the box
144 is designated the ``cold'' slab. The artificial momentum flux will be
145 established by transferring momentum from the cold slab and into the
146 hot slab.
147
148 Rather than using momentum swaps, we use a series of velocity scaling
149 moves. For molecules $\{i\}$ located within the cold slab,
150 \begin{equation}
151 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
152 x & 0 & 0 \\
153 0 & y & 0 \\
154 0 & 0 & z \\
155 \end{array} \right) \cdot \vec{v}_i
156 \end{equation}
157 where ${x, y, z}$ are a set of 3 scaling variables for each of the
158 three directions in the system. Likewise, the molecules $\{j\}$
159 located in the hot slab will see a concomitant scaling of velocities,
160 \begin{equation}
161 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
162 x^\prime & 0 & 0 \\
163 0 & y^\prime & 0 \\
164 0 & 0 & z^\prime \\
165 \end{array} \right) \cdot \vec{v}_j
166 \end{equation}
167
168 Conservation of linear momentum in each of the three directions
169 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
170 parameters together:
171 \begin{equation}
172 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
173 \end{equation}
174 where
175 \begin{eqnarray}
176 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
177 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
178 \label{eq:momentumdef}
179 \end{eqnarray}
180 Therefore, for each of the three directions, the hot scaling
181 parameters are a simple function of the cold scaling parameters and
182 the instantaneous linear momentum in each of the two slabs.
183 \begin{equation}
184 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
185 \label{eq:hotcoldscaling}
186 \end{equation}
187 where
188 \begin{equation}
189 p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
190 \end{equation}
191 for convenience.
192
193 Conservation of total energy also places constraints on the scaling:
194 \begin{equation}
195 \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
196 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
197 \end{equation}
198 where the translational kinetic energies, $K_h^\alpha$ and
199 $K_c^\alpha$, are computed for each of the three directions in a
200 similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
201 Substituting in the expressions for the hot scaling parameters
202 ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
203 {\it constraint ellipsoid}:
204 \begin{equation}
205 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
206 \label{eq:constraintEllipsoid}
207 \end{equation}
208 where the constants are obtained from the instantaneous values of the
209 linear momenta and kinetic energies for the hot and cold slabs,
210 \begin{eqnarray}
211 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
212 \left(p_\alpha\right)^2\right) \\
213 b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
214 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
215 \label{eq:constraintEllipsoidConsts}
216 \end{eqnarray}
217 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
218 cold slab scaling parameters which can be applied while preserving
219 both linear momentum in all three directions as well as total kinetic
220 energy.
221
222 The goal of using velocity scaling variables is to transfer linear
223 momentum or kinetic energy from the cold slab to the hot slab. If the
224 hot and cold slabs are separated along the z-axis, the energy flux is
225 given simply by the decrease in kinetic energy of the cold bin:
226 \begin{equation}
227 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
228 \end{equation}
229 The expression for the energy flux can be re-written as another
230 ellipsoid centered on $(x,y,z) = 0$:
231 \begin{equation}
232 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
233 \label{eq:fluxEllipsoid}
234 \end{equation}
235 The spatial extent of the {\it thermal flux ellipsoid} is governed
236 both by a targetted value, $J_z$ as well as the instantaneous values
237 of the kinetic energy components in the cold bin.
238
239 To satisfy an energetic flux as well as the conservation constraints,
240 we must determine the points ${x,y,z}$ which lie on both the
241 constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
242 ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
243 two ellipsoids in 3-dimensional space.
244
245 \begin{figure}
246 \includegraphics[width=\linewidth]{ellipsoids}
247 \caption{Scaling points which maintain both constant energy and
248 constant linear momentum of the system lie on the surface of the
249 {\it constraint ellipsoid} while points which generate the target
250 momentum flux lie on the surface of the {\it flux ellipsoid}. The
251 velocity distributions in the cold bin are scaled by only those
252 points which lie on both ellipsoids.}
253 \label{ellipsoids}
254 \end{figure}
255
256 One may also define {\it momentum} flux (say along the $x$-direction) as:
257 \begin{equation}
258 (1-x) P_c^x = j_z(p_x)\Delta t
259 \label{eq:fluxPlane}
260 \end{equation}
261 The above {\it momentum flux plane} is perpendicular to the $x$-axis,
262 with its position governed both by a target value, $j_z(p_x)$ as well
263 as the instantaneous value of the momentum along the $x$-direction.
264
265 In order to satisfy a momentum flux as well as the conservation
266 constraints, we must determine the points ${x,y,z}$ which lie on both
267 the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
268 flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
269 ellipsoid and a plane in 3-dimensional space.
270
271 In both the momentum and energy flux scenarios, valid scaling
272 parameters are arrived at by solving geometric intersection problems
273 in $x, y, z$ space in order to obtain cold slab scaling parameters.
274 Once the scaling variables for the cold slab are known, the hot slab
275 scaling has also been determined.
276
277
278 The following problem will be choosing an optimal set of scaling
279 variables among the possible sets. Although this method is inherently
280 non-isotropic, the goal is still to maintain the system as isotropic
281 as possible. Under this consideration, one would like the kinetic
282 energies in different directions could become as close as each other
283 after each scaling. Simultaneously, one would also like each scaling
284 as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
285 large perturbation to the system. Therefore, one approach to obtain
286 the scaling variables would be constructing an criteria function, with
287 constraints as above equation sets, and solving the function's minimum
288 by method like Lagrange multipliers.
289
290 In order to save computation time, we have a different approach to a
291 relatively good set of scaling variables with much less calculation
292 than above. Here is the detail of our simplification of the problem.
293
294 In the case of kinetic energy transfer, we impose another constraint
295 ${x = y}$, into the equation sets. Consequently, there are two
296 variables left. And now one only needs to solve a set of two {\it
297 ellipses equations}. This problem would be transformed into solving
298 one quartic equation for one of the two variables. There are known
299 generic methods that solve real roots of quartic equations. Then one
300 can determine the other variable and obtain sets of scaling
301 variables. Among these sets, one can apply the above criteria to
302 choose the best set, while much faster with only a few sets to choose.
303
304 In the case of momentum flux transfer, we impose another constraint to
305 set the kinetic energy transfer as zero. In another word, we apply
306 Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
307 variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
308 of equations on the above kinetic energy transfer problem. Therefore,
309 an approach similar to the above would be sufficient for this as well.
310
311 \section{Computational Details}
312 \subsection{Lennard-Jones Fluid}
313 Our simulation consists of a series of systems. All of these
314 simulations were run with the OpenMD simulation software
315 package\cite{Meineke:2005gd} integrated with RNEMD codes.
316
317 A Lennard-Jones fluid system was built and tested first. In order to
318 compare our method with swapping RNEMD, a series of simulations were
319 performed to calculate the shear viscosity and thermal conductivity of
320 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
321 \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
322 ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
323 comparison between our results and others. These simulations used
324 velocity Verlet algorithm with reduced timestep ${\tau^* =
325 4.6\times10^{-4}}$.
326
327 For shear viscosity calculation, the reduced temperature was ${T^* =
328 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
329 ensemble (NVT), then equilibrated in microcanonical ensemble
330 (NVE). Establishing and stablizing momentum gradient were followed
331 also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
332 adopted.\cite{ISI:000080382700030} The simulation box was under
333 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
334 the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
335 most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
336 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
337 frequency were chosen. According to each result from swapping
338 RNEMD, scaling RNEMD simulations were run with the target momentum
339 flux set to produce a similar momentum flux, and consequently shear
340 rate. Furthermore, various scaling frequencies can be tested for one
341 single swapping rate. To test the temperature homogeneity in our
342 system of swapping and scaling methods, temperatures of different
343 dimensions in all the slabs were observed. Most of the simulations
344 include $10^5$ steps of equilibration without imposing momentum flux,
345 $10^5$ steps of stablization with imposing unphysical momentum
346 transfer, and $10^6$ steps of data collection under RNEMD. For
347 relatively high momentum flux simulations, ${5\times10^5}$ step data
348 collection is sufficient. For some low momentum flux simulations,
349 ${2\times10^6}$ steps were necessary.
350
351 After each simulation, the shear viscosity was calculated in reduced
352 unit. The momentum flux was calculated with total unphysical
353 transferred momentum ${P_x}$ and data collection time $t$:
354 \begin{equation}
355 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
356 \end{equation}
357 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
358 box, and physical momentum transfer occurs in two ways due to our
359 periodic boundary condition settings. And the velocity gradient
360 ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
361 linear regression of the velocity profile. From the shear viscosity
362 $\eta$ calculated with the above parameters, one can further convert
363 it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
364
365 For thermal conductivity calculations, simulations were first run under
366 reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
367 ensemble. Muller-Plathe's algorithm was adopted in the swapping
368 method. Under identical simulation box parameters with our shear
369 viscosity calculations, in each swap, the top slab exchanges all three
370 translational momentum components of the molecule with least kinetic
371 energy with the same components of the molecule in the center slab
372 with most kinetic energy, unless this ``coldest'' molecule in the
373 ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
374 ``cold'' slab. According to swapping RNEMD results, target energy flux
375 for scaling RNEMD simulations can be set. Also, various scaling
376 frequencies can be tested for one target energy flux. To compare the
377 performance between swapping and scaling method, distributions of
378 velocity and speed in different slabs were observed.
379
380 For each swapping rate, thermal conductivity was calculated in reduced
381 unit. The energy flux was calculated similarly to the momentum flux,
382 with total unphysical transferred energy ${E_{total}}$ and data collection
383 time $t$:
384 \begin{equation}
385 J_z = \frac{E_{total}}{2 t L_x L_y}
386 \end{equation}
387 And the temperature gradient ${\langle\partial T/\partial z\rangle}$
388 can be obtained by a linear regression of the temperature
389 profile. From the thermal conductivity $\lambda$ calculated, one can
390 further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
391 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
392
393 \subsection{ Water / Metal Thermal Conductivity}
394 Another series of our simulation is the calculation of interfacial
395 thermal conductivity of a Au/H$_2$O system. Respective calculations of
396 liquid water (Extended Simple Point Charge model) and crystal gold
397 thermal conductivity were performed and compared with current results
398 to ensure the validity of NIVS-RNEMD. After that, a mixture system was
399 simulated.
400
401 For thermal conductivity calculation of bulk water, a simulation box
402 consisting of 1000 molecules were first equilibrated under ambient
403 pressure and temperature conditions using NPT ensemble, followed by
404 equilibration in fixed volume (NVT). The system was then equilibrated in
405 microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
406 stable thermal gradient was followed. The simulation box was under
407 periodic boundary condition and devided into 10 slabs. Data collection
408 process was similar to Lennard-Jones fluid system.
409
410 Thermal conductivity calculation of bulk crystal gold used a similar
411 protocol. Two types of force field parameters, Embedded Atom Method
412 (EAM) and Quantum Sutten-Chen (QSC) force field were used
413 respectively. The face-centered cubic crystal simulation box consists of
414 2880 Au atoms. The lattice was first allowed volume change to relax
415 under ambient temperature and pressure. Equilibrations in canonical and
416 microcanonical ensemble were followed in order. With the simulation
417 lattice devided evenly into 10 slabs, different thermal gradients were
418 established by applying a set of target thermal transfer flux. Data of
419 the series of thermal gradients was collected for calculation.
420
421 After simulations of bulk water and crystal gold, a mixture system was
422 constructed, consisting of 1188 Au atoms and 1862 H$_2$O
423 molecules. Spohr potential was adopted in depicting the interaction
424 between metal atom and water molecule.\cite{ISI:000167766600035} A
425 similar protocol of equilibration was followed. Several thermal
426 gradients was built under different target thermal flux. It was found
427 out that compared to our previous simulation systems, the two phases
428 could have large temperature difference even under a relatively low
429 thermal flux. Therefore, under our low flux conditions, it is assumed
430 that the metal and water phases have respectively homogeneous
431 temperature, excluding the surface regions. In calculating the
432 interfacial thermal conductivity $G$, this assumptioin was applied and
433 thus our formula becomes:
434
435 \begin{equation}
436 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
437 \langle T_{water}\rangle \right)}
438 \label{interfaceCalc}
439 \end{equation}
440 where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
441 and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
442 average observed temperature of gold and water phases respectively.
443
444 \section{Results And Discussions}
445 \subsection{Thermal Conductivity}
446 \subsubsection{Lennard-Jones Fluid}
447 Our thermal conductivity calculations show that scaling method results
448 agree with swapping method. Four different exchange intervals were
449 tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
450 10fs exchange interval, target exchange kinetic energy was set to
451 produce equivalent kinetic energy flux as in swapping method. And
452 similar thermal gradients were observed with similar thermal flux in
453 two simulation methods (Figure \ref{thermalGrad}).
454
455 \begin{table*}
456 \begin{minipage}{\linewidth}
457 \begin{center}
458
459 \caption{Calculation results for thermal conductivity of Lennard-Jones
460 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
461 swap and scale methods at various kinetic energy exchange rates. Results
462 in reduced unit. Errors of calculations in parentheses.}
463
464 \begin{tabular}{ccc}
465 \hline
466 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
467 $\lambda^*_{scale}$\\
468 \hline
469 250 & 7.03(0.34) & 7.30(0.10)\\
470 500 & 7.03(0.14) & 6.95(0.09)\\
471 1000 & 6.91(0.42) & 7.19(0.07)\\
472 2000 & 7.52(0.15) & 7.19(0.28)\\
473 \hline
474 \end{tabular}
475 \label{thermalLJRes}
476 \end{center}
477 \end{minipage}
478 \end{table*}
479
480 \begin{figure}
481 \includegraphics[width=\linewidth]{thermalGrad}
482 \caption{Temperature gradients under various kinetic energy flux of
483 thermal conductivity simulations}
484 \label{thermalGrad}
485 \end{figure}
486
487 During these simulations, molecule velocities were recorded in 1000 of
488 all the snapshots of one single data collection process. These
489 velocity data were used to produce histograms of velocity and speed
490 distribution in different slabs. From these histograms, it is observed
491 that under relatively high unphysical kinetic energy flux, speed and
492 velocity distribution of molecules in slabs where swapping occured
493 could deviate from Maxwell-Boltzmann distribution. Figure
494 \ref{histSwap} illustrates how these distributions deviate from an
495 ideal distribution. In high temperature slab, probability density in
496 low speed is confidently smaller than ideal curve fit; in low
497 temperature slab, probability density in high speed is smaller than
498 ideal, while larger than ideal in low speed. This phenomenon is more
499 obvious in our high swapping rate simulations. And this deviation
500 could also leads to deviation of distribution of velocity in various
501 dimensions. One feature of these deviated distribution is that in high
502 temperature slab, the ideal Gaussian peak was changed into a
503 relatively flat plateau; while in low temperature slab, that peak
504 appears sharper. This problem is rooted in the mechanism of the
505 swapping method. Continually depleting low (high) speed particles in
506 the high (low) temperature slab could not be complemented by
507 diffusions of low (high) speed particles from neighbor slabs, unless
508 in suffciently low swapping rate. Simutaneously, surplus low speed
509 particles in the low temperature slab do not have sufficient time to
510 diffuse to neighbor slabs. However, thermal exchange rate should reach
511 a minimum level to produce an observable thermal gradient under noise
512 interference. Consequently, swapping RNEMD has a relatively narrow
513 choice of swapping rate to satisfy these above restrictions.
514
515 \begin{figure}
516 \includegraphics[width=\linewidth]{histSwap}
517 \caption{Speed distribution for thermal conductivity using swapping
518 RNEMD. Shown is from the simulation with 250 fs exchange interval.}
519 \label{histSwap}
520 \end{figure}
521
522 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
523 curve fit (Figure \ref{histScale}). Essentially, after scaling, a
524 Gaussian distribution function would remain Gaussian. Although a
525 single scaling is non-isotropic in all three dimensions, our scaling
526 coefficient criteria could help maintian the scaling region as
527 isotropic as possible. On the other hand, scaling coefficients are
528 preferred to be as close to 1 as possible, which also helps minimize
529 the difference among different dimensions. This is possible if scaling
530 interval and one-time thermal transfer energy are well
531 chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
532 thermal flux as the previous RNEMD method without large perturbation
533 to the distribution of velocity and speed in the exchange regions.
534
535 \begin{figure}
536 \includegraphics[width=\linewidth]{histScale}
537 \caption{Speed distribution for thermal conductivity using scaling
538 RNEMD. Shown is from the simulation with an equilvalent thermal flux
539 as an 250 fs exchange interval swapping simulation.}
540 \label{histScale}
541 \end{figure}
542
543 \subsubsection{SPC/E Water}
544 Our results of SPC/E water thermal conductivity are comparable to
545 Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
546 previous swapping RNEMD method for their calculation. Bedrov {\it et
547 al.}\cite{ISI:000090151400044} argued that exchange of the molecule
548 center-of-mass velocities instead of single atom velocities in a
549 molecule conserves the total kinetic energy and linear momentum. This
550 principle is adopted in our simulations. Scaling is applied to the
551 velocities of the rigid bodies of SPC/E model water molecules, instead
552 of each hydrogen and oxygen atoms in relevant water molecules. As
553 shown in Figure \ref{spceGrad}, temperature gradients were established
554 similar to their system. However, the average temperature of our
555 system is 300K, while theirs is 318K, which would be attributed for
556 part of the difference between the final calculation results (Table
557 \ref{spceThermal}). Both methods yields values in agreement with
558 experiment. And this shows the applicability of our method to
559 multi-atom molecular system.
560
561 \begin{figure}
562 \includegraphics[width=\linewidth]{spceGrad}
563 \caption{Temperature gradients for SPC/E water thermal conductivity.}
564 \label{spceGrad}
565 \end{figure}
566
567 \begin{table*}
568 \begin{minipage}{\linewidth}
569 \begin{center}
570
571 \caption{Calculation results for thermal conductivity of SPC/E water
572 at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
573 calculations in parentheses. }
574
575 \begin{tabular}{cccc}
576 \hline
577 $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
578 & This work & Previous simulations\cite{ISI:000090151400044} &
579 Experiment$^a$\\
580 \hline
581 0.38 & 0.816(0.044) & & 0.64\\
582 0.81 & 0.770(0.008) & 0.784\\
583 1.54 & 0.813(0.007) & 0.730\\
584 \hline
585 \end{tabular}
586 \label{spceThermal}
587 \end{center}
588 \end{minipage}
589 \end{table*}
590
591 \subsubsection{Crystal Gold}
592 Our results of gold thermal conductivity using two force fields are
593 shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
594 these calculations,the end and middle slabs were excluded in thermal
595 gradient regession and only used as heat source and drain in the
596 systems. Our yielded values using EAM force field are slightly larger
597 than those using QSC force field. However, both series are
598 significantly smaller than experimental value by an order of more than
599 100. It has been verified that this difference is mainly attributed to
600 the lack of electron interaction representation in these force field
601 parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
602 force field parameters in their metal thermal conductivity
603 calculations. The Non-Equilibrium MD method they employed in their
604 simulations produced comparable results to ours. As Zhang {\it et
605 al.}\cite{ISI:000231042800044} stated, thermal conductivity values
606 are influenced mainly by force field. Therefore, it is confident to
607 conclude that NIVS-RNEMD is applicable to metal force field system.
608
609 \begin{figure}
610 \includegraphics[width=\linewidth]{AuGrad}
611 \caption{Temperature gradients for thermal conductivity calculation of
612 crystal gold using QSC force field.}
613 \label{AuGrad}
614 \end{figure}
615
616 \begin{table*}
617 \begin{minipage}{\linewidth}
618 \begin{center}
619
620 \caption{Calculation results for thermal conductivity of crystal gold
621 using QSC force field at ${\langle T\rangle}$ = 300K at various
622 thermal exchange rates. Errors of calculations in parentheses. }
623
624 \begin{tabular}{cc}
625 \hline
626 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
627 \hline
628 1.44 & 1.10(0.01)\\
629 2.86 & 1.08(0.02)\\
630 5.14 & 1.15(0.01)\\
631 \hline
632 \end{tabular}
633 \label{qscThermal}
634 \end{center}
635 \end{minipage}
636 \end{table*}
637
638 \begin{figure}
639 \includegraphics[width=\linewidth]{eamGrad}
640 \caption{Temperature gradients for thermal conductivity calculation of
641 crystal gold using EAM force field.}
642 \label{eamGrad}
643 \end{figure}
644
645 \begin{table*}
646 \begin{minipage}{\linewidth}
647 \begin{center}
648
649 \caption{Calculation results for thermal conductivity of crystal gold
650 using EAM force field at ${\langle T\rangle}$ = 300K at various
651 thermal exchange rates. Errors of calculations in parentheses. }
652
653 \begin{tabular}{cc}
654 \hline
655 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
656 \hline
657 1.24 & 1.24(0.06)\\
658 2.06 & 1.37(0.04)\\
659 2.55 & 1.41(0.03)\\
660 \hline
661 \end{tabular}
662 \label{eamThermal}
663 \end{center}
664 \end{minipage}
665 \end{table*}
666
667
668 \subsection{Interfaciel Thermal Conductivity}
669 After simulations of homogeneous water and gold systems using
670 NIVS-RNEMD method were proved valid, calculation of gold/water
671 interfacial thermal conductivity was followed. It is found out that
672 the low interfacial conductance is probably due to the hydrophobic
673 surface in our system. Figure \ref{interfaceDensity} demonstrates mass
674 density change along $z$-axis, which is perpendicular to the
675 gold/water interface. It is observed that water density significantly
676 decreases when approaching the surface. Under this low thermal
677 conductance, both gold and water phase have sufficient time to
678 eliminate temperature difference inside respectively (Figure
679 \ref{interfaceGrad}). With indistinguishable temperature difference
680 within respective phase, it is valid to assume that the temperature
681 difference between gold and water on surface would be approximately
682 the same as the difference between the gold and water phase. This
683 assumption enables convenient calculation of $G$ using
684 Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
685 layer of water and gold close enough to surface, which would have
686 greater fluctuation and lower accuracy. Reported results (Table
687 \ref{interfaceRes}) are of two orders of magnitude smaller than our
688 calculations on homogeneous systems, and thus have larger relative
689 errors than our calculation results on homogeneous systems.
690
691 \begin{figure}
692 \includegraphics[width=\linewidth]{interfaceDensity}
693 \caption{Density profile for interfacial thermal conductivity
694 simulation box. Significant water density decrease is observed on
695 gold surface.}
696 \label{interfaceDensity}
697 \end{figure}
698
699 \begin{figure}
700 \includegraphics[width=\linewidth]{interfaceGrad}
701 \caption{Temperature profiles for interfacial thermal conductivity
702 simulation box. Temperatures of different slabs in the same phase
703 show no significant difference.}
704 \label{interfaceGrad}
705 \end{figure}
706
707 \begin{table*}
708 \begin{minipage}{\linewidth}
709 \begin{center}
710
711 \caption{Calculation results for interfacial thermal conductivity
712 at ${\langle T\rangle \sim}$ 300K at various thermal exchange
713 rates. Errors of calculations in parentheses. }
714
715 \begin{tabular}{cccc}
716 \hline
717 $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
718 \hline
719 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
720 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
721 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
722 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
723 \hline
724 \end{tabular}
725 \label{interfaceRes}
726 \end{center}
727 \end{minipage}
728 \end{table*}
729
730 \subsection{Shear Viscosity}
731 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
732 produced comparable shear viscosity to swap RNEMD method. In Table
733 \ref{shearRate}, the names of the calculated samples are devided into
734 two parts. The first number refers to total slabs in one simulation
735 box. The second number refers to the swapping interval in swap method, or
736 in scale method the equilvalent swapping interval that the same
737 momentum flux would theoretically result in swap method. All the scale
738 method results were from simulations that had a scaling interval of 10
739 time steps. The average molecular momentum gradients of these samples
740 are shown in Figure \ref{shearGrad}.
741
742 \begin{table*}
743 \begin{minipage}{\linewidth}
744 \begin{center}
745
746 \caption{Calculation results for shear viscosity of Lennard-Jones
747 fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
748 methods at various momentum exchange rates. Results in reduced
749 unit. Errors of calculations in parentheses. }
750
751 \begin{tabular}{ccc}
752 \hline
753 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
754 \hline
755 20-500 & 3.64(0.05) & 3.76(0.09)\\
756 20-1000 & 3.52(0.16) & 3.66(0.06)\\
757 20-2000 & 3.72(0.05) & 3.32(0.18)\\
758 20-2500 & 3.42(0.06) & 3.43(0.08)\\
759 \hline
760 \end{tabular}
761 \label{shearRate}
762 \end{center}
763 \end{minipage}
764 \end{table*}
765
766 \begin{figure}
767 \includegraphics[width=\linewidth]{shearGrad}
768 \caption{Average momentum gradients of shear viscosity simulations}
769 \label{shearGrad}
770 \end{figure}
771
772 \begin{figure}
773 \includegraphics[width=\linewidth]{shearTempScale}
774 \caption{Temperature profile for scaling RNEMD simulation.}
775 \label{shearTempScale}
776 \end{figure}
777 However, observations of temperatures along three dimensions show that
778 inhomogeneity occurs in scaling RNEMD simulations, particularly in the
779 two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
780 relatively large imposed momentum flux, the temperature difference among $x$
781 and the other two dimensions was significant. This would result from the
782 algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
783 momentum gradient is set up, $P_c^x$ would be roughly stable
784 ($<0$). Consequently, scaling factor $x$ would most probably larger
785 than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
786 keep increase after most scaling steps. And if there is not enough time
787 for the kinetic energy to exchange among different dimensions and
788 different slabs, the system would finally build up temperature
789 (kinetic energy) difference among the three dimensions. Also, between
790 $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
791 are closer to neighbor slabs. This is due to momentum transfer along
792 $z$ dimension between slabs.
793
794 Although results between scaling and swapping methods are comparable,
795 the inherent temperature inhomogeneity even in relatively low imposed
796 exchange momentum flux simulations makes scaling RNEMD method less
797 attractive than swapping RNEMD in shear viscosity calculation.
798
799 \section{Conclusions}
800 NIVS-RNEMD simulation method is developed and tested on various
801 systems. Simulation results demonstrate its validity in thermal
802 conductivity calculations, from Lennard-Jones fluid to multi-atom
803 molecule like water and metal crystals. NIVS-RNEMD improves
804 non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
805 methods. Furthermore, it develops a valid means for unphysical thermal
806 transfer between different species of molecules, and thus extends its
807 applicability to interfacial systems. Our calculation of gold/water
808 interfacial thermal conductivity demonstrates this advantage over
809 previous RNEMD methods. NIVS-RNEMD has also limited application on
810 shear viscosity calculations, but could cause temperature difference
811 among different dimensions under high momentum flux. Modification is
812 necessary to extend the applicability of NIVS-RNEMD in shear viscosity
813 calculations.
814
815 \section{Acknowledgments}
816 Support for this project was provided by the National Science
817 Foundation under grant CHE-0848243. Computational time was provided by
818 the Center for Research Computing (CRC) at the University of Notre
819 Dame. \newpage
820
821 \bibliographystyle{aip}
822 \bibliography{nivsRnemd}
823
824 \end{doublespace}
825 \end{document}
826