ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3583
Committed: Tue Apr 13 19:59:50 2010 UTC (14 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 37162 byte(s)
Log Message:
*** empty log message ***

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
69
70 \section{Introduction}
71 The original formulation of Reverse Non-equilibrium Molecular Dynamics
72 (RNEMD) obtains transport coefficients (thermal conductivity and shear
73 viscosity) in a fluid by imposing an artificial momentum flux between
74 two thin parallel slabs of material that are spatially separated in
75 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
76 artificial flux is typically created by periodically ``swapping''
77 either the entire momentum vector $\vec{p}$ or single components of
78 this vector ($p_x$) between molecules in each of the two slabs. If
79 the two slabs are separated along the $z$ coordinate, the imposed flux
80 is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
81 response of a simulated system to the imposed momentum flux will
82 typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
83 The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
84 easily obtained by assuming linear response of the system,
85 \begin{eqnarray}
86 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
87 J_z & = & \lambda \frac{\partial T}{\partial z}
88 \end{eqnarray}
89 RNEMD has been widely used to provide computational estimates of thermal
90 conductivities and shear viscosities in a wide range of materials,
91 from liquid copper to monatomic liquids to molecular fluids
92 (e.g. ionic liquids).\cite{ISI:000246190100032} [MORE CITATIONS HERE]
93
94 \begin{figure}
95 \includegraphics[width=\linewidth]{thermalDemo}
96 \caption{RNEMD methods impose an unphysical transfer of momentum or
97 kinetic energy between a ``hot'' slab and a ``cold'' slab in the
98 simulation box. The molecular system responds to this imposed flux
99 by generating a momentum or temperature gradient. The slope of the
100 gradient can then be used to compute transport properties (e.g.
101 shear viscosity and thermal conductivity).}
102 \label{thermalDemo}
103 \end{figure}
104
105 RNEMD is preferable in many ways to the forward NEMD methods
106 [CITATIONS NEEDED] because it imposes what is typically difficult to measure
107 (a flux or stress) and it is typically much easier to compute momentum
108 gradients or strains (the response). For similar reasons, RNEMD is
109 also preferable to slowly-converging equilibrium methods for measuring
110 thermal conductivity and shear viscosity (using Green-Kubo relations
111 [CITATIONS NEEDED] or the Helfand moment approach of Viscardy {\it et
112 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
113 computing difficult to measure quantities.
114
115 Another attractive feature of RNEMD is that it conserves both total
116 linear momentum and total energy during the swaps (as long as the two
117 molecules have the same identity), so the swapped configurations are
118 typically samples from the same manifold of states in the
119 microcanonical ensemble.
120
121 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
122 some problems with the original RNEMD swap technique. Notably, large
123 momentum fluxes (equivalent to frequent momentum swaps between the
124 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
125 momentum distributions in the two slabs, as well as non-linear thermal
126 and velocity distributions along the direction of the imposed flux
127 ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
128 and self-adjusting metrics for retaining the usability of the method.
129
130 In this paper, we develop and test a method for non-isotropic velocity
131 scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
132 (conservation of linear momentum and total energy, compatibility with
133 periodic boundary conditions) while establishing true thermal
134 distributions in each of the two slabs. In the next section, we
135 present the method for determining the scaling constraints. We then
136 test the method on both single component, multi-component, and
137 non-isotropic mixtures and show that it is capable of providing
138 reasonable estimates of the thermal conductivity and shear viscosity
139 in these cases.
140
141 \section{Methodology}
142 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
143 periodic system is partitioned into a series of thin slabs along one
144 axis ($z$). One of the slabs at the end of the periodic box is
145 designated the ``hot'' slab, while the slab in the center of the box
146 is designated the ``cold'' slab. The artificial momentum flux will be
147 established by transferring momentum from the cold slab and into the
148 hot slab.
149
150 Rather than using momentum swaps, we use a series of velocity scaling
151 moves. For molecules $\{i\}$ located within the cold slab,
152 \begin{equation}
153 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
154 x & 0 & 0 \\
155 0 & y & 0 \\
156 0 & 0 & z \\
157 \end{array} \right) \cdot \vec{v}_i
158 \end{equation}
159 where ${x, y, z}$ are a set of 3 scaling variables for each of the
160 three directions in the system. Likewise, the molecules $\{j\}$
161 located in the hot slab will see a concomitant scaling of velocities,
162 \begin{equation}
163 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
164 x^\prime & 0 & 0 \\
165 0 & y^\prime & 0 \\
166 0 & 0 & z^\prime \\
167 \end{array} \right) \cdot \vec{v}_j
168 \end{equation}
169
170 Conservation of linear momentum in each of the three directions
171 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
172 parameters together:
173 \begin{equation}
174 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
175 \end{equation}
176 where
177 \begin{eqnarray}
178 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
179 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
180 \label{eq:momentumdef}
181 \end{eqnarray}
182 Therefore, for each of the three directions, the hot scaling
183 parameters are a simple function of the cold scaling parameters and
184 the instantaneous linear momentum in each of the two slabs.
185 \begin{equation}
186 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
187 \label{eq:hotcoldscaling}
188 \end{equation}
189 where
190 \begin{equation}
191 p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
192 \end{equation}
193 for convenience.
194
195 Conservation of total energy also places constraints on the scaling:
196 \begin{equation}
197 \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
198 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
199 \end{equation}
200 where the translational kinetic energies, $K_h^\alpha$ and
201 $K_c^\alpha$, are computed for each of the three directions in a
202 similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
203 Substituting in the expressions for the hot scaling parameters
204 ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
205 {\it constraint ellipsoid}:
206 \begin{equation}
207 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
208 \label{eq:constraintEllipsoid}
209 \end{equation}
210 where the constants are obtained from the instantaneous values of the
211 linear momenta and kinetic energies for the hot and cold slabs,
212 \begin{eqnarray}
213 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
214 \left(p_\alpha\right)^2\right) \\
215 b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
216 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
217 \label{eq:constraintEllipsoidConsts}
218 \end{eqnarray}
219 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
220 cold slab scaling parameters which can be applied while preserving
221 both linear momentum in all three directions as well as total kinetic
222 energy.
223
224 The goal of using velocity scaling variables is to transfer linear
225 momentum or kinetic energy from the cold slab to the hot slab. If the
226 hot and cold slabs are separated along the z-axis, the energy flux is
227 given simply by the decrease in kinetic energy of the cold bin:
228 \begin{equation}
229 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
230 \end{equation}
231 The expression for the energy flux can be re-written as another
232 ellipsoid centered on $(x,y,z) = 0$:
233 \begin{equation}
234 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
235 \label{eq:fluxEllipsoid}
236 \end{equation}
237 The spatial extent of the {\it thermal flux ellipsoid} is governed
238 both by a targetted value, $J_z$ as well as the instantaneous values
239 of the kinetic energy components in the cold bin.
240
241 To satisfy an energetic flux as well as the conservation constraints,
242 we must determine the points ${x,y,z}$ which lie on both the
243 constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux
244 ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the
245 two ellipsoids in 3-dimensional space.
246
247 \begin{figure}
248 \includegraphics[width=\linewidth]{ellipsoids}
249 \caption{Scaling points which maintain both constant energy and
250 constant linear momentum of the system lie on the surface of the
251 {\it constraint ellipsoid} while points which generate the target
252 momentum flux lie on the surface of the {\it flux ellipsoid}. The
253 velocity distributions in the cold bin are scaled by only those
254 points which lie on both ellipsoids.}
255 \label{ellipsoids}
256 \end{figure}
257
258 One may also define {\it momentum} flux (say along the $x$-direction) as:
259 \begin{equation}
260 (1-x) P_c^x = j_z(p_x)\Delta t
261 \label{eq:fluxPlane}
262 \end{equation}
263 The above {\it momentum flux plane} is perpendicular to the $x$-axis,
264 with its position governed both by a target value, $j_z(p_x)$ as well
265 as the instantaneous value of the momentum along the $x$-direction.
266
267 In order to satisfy a momentum flux as well as the conservation
268 constraints, we must determine the points ${x,y,z}$ which lie on both
269 the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
270 flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
271 ellipsoid and a plane in 3-dimensional space.
272
273 In both the momentum and energy flux scenarios, valid scaling
274 parameters are arrived at by solving geometric intersection problems
275 in $x, y, z$ space in order to obtain cold slab scaling parameters.
276 Once the scaling variables for the cold slab are known, the hot slab
277 scaling has also been determined.
278
279
280 The following problem will be choosing an optimal set of scaling
281 variables among the possible sets. Although this method is inherently
282 non-isotropic, the goal is still to maintain the system as isotropic
283 as possible. Under this consideration, one would like the kinetic
284 energies in different directions could become as close as each other
285 after each scaling. Simultaneously, one would also like each scaling
286 as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
287 large perturbation to the system. Therefore, one approach to obtain
288 the scaling variables would be constructing an criteria function, with
289 constraints as above equation sets, and solving the function's minimum
290 by method like Lagrange multipliers.
291
292 In order to save computation time, we have a different approach to a
293 relatively good set of scaling variables with much less calculation
294 than above. Here is the detail of our simplification of the problem.
295
296 In the case of kinetic energy transfer, we impose another constraint
297 ${x = y}$, into the equation sets. Consequently, there are two
298 variables left. And now one only needs to solve a set of two {\it
299 ellipses equations}. This problem would be transformed into solving
300 one quartic equation for one of the two variables. There are known
301 generic methods that solve real roots of quartic equations. Then one
302 can determine the other variable and obtain sets of scaling
303 variables. Among these sets, one can apply the above criteria to
304 choose the best set, while much faster with only a few sets to choose.
305
306 In the case of momentum flux transfer, we impose another constraint to
307 set the kinetic energy transfer as zero. In another word, we apply
308 Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
309 variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
310 of equations on the above kinetic energy transfer problem. Therefore,
311 an approach similar to the above would be sufficient for this as well.
312
313 \section{Computational Details}
314 \subsection{Lennard-Jones Fluid}
315 Our simulation consists of a series of systems. All of these
316 simulations were run with the OpenMD simulation software
317 package\cite{Meineke:2005gd} integrated with RNEMD codes.
318
319 A Lennard-Jones fluid system was built and tested first. In order to
320 compare our method with swapping RNEMD, a series of simulations were
321 performed to calculate the shear viscosity and thermal conductivity of
322 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
323 \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
324 ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
325 comparison between our results and others. These simulations used
326 velocity Verlet algorithm with reduced timestep ${\tau^* =
327 4.6\times10^{-4}}$.
328
329 For shear viscosity calculation, the reduced temperature was ${T^* =
330 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
331 ensemble (NVT), then equilibrated in microcanonical ensemble
332 (NVE). Establishing and stablizing momentum gradient were followed
333 also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
334 adopted.\cite{ISI:000080382700030} The simulation box was under
335 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
336 the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
337 most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
338 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
339 frequency were chosen. According to each result from swapping
340 RNEMD, scaling RNEMD simulations were run with the target momentum
341 flux set to produce a similar momentum flux, and consequently shear
342 rate. Furthermore, various scaling frequencies can be tested for one
343 single swapping rate. To test the temperature homogeneity in our
344 system of swapping and scaling methods, temperatures of different
345 dimensions in all the slabs were observed. Most of the simulations
346 include $10^5$ steps of equilibration without imposing momentum flux,
347 $10^5$ steps of stablization with imposing unphysical momentum
348 transfer, and $10^6$ steps of data collection under RNEMD. For
349 relatively high momentum flux simulations, ${5\times10^5}$ step data
350 collection is sufficient. For some low momentum flux simulations,
351 ${2\times10^6}$ steps were necessary.
352
353 After each simulation, the shear viscosity was calculated in reduced
354 unit. The momentum flux was calculated with total unphysical
355 transferred momentum ${P_x}$ and data collection time $t$:
356 \begin{equation}
357 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
358 \end{equation}
359 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
360 box, and physical momentum transfer occurs in two ways due to our
361 periodic boundary condition settings. And the velocity gradient
362 ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
363 linear regression of the velocity profile. From the shear viscosity
364 $\eta$ calculated with the above parameters, one can further convert
365 it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
366
367 For thermal conductivity calculations, simulations were first run under
368 reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
369 ensemble. Muller-Plathe's algorithm was adopted in the swapping
370 method. Under identical simulation box parameters with our shear
371 viscosity calculations, in each swap, the top slab exchanges all three
372 translational momentum components of the molecule with least kinetic
373 energy with the same components of the molecule in the center slab
374 with most kinetic energy, unless this ``coldest'' molecule in the
375 ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
376 ``cold'' slab. According to swapping RNEMD results, target energy flux
377 for scaling RNEMD simulations can be set. Also, various scaling
378 frequencies can be tested for one target energy flux. To compare the
379 performance between swapping and scaling method, distributions of
380 velocity and speed in different slabs were observed.
381
382 For each swapping rate, thermal conductivity was calculated in reduced
383 unit. The energy flux was calculated similarly to the momentum flux,
384 with total unphysical transferred energy ${E_{total}}$ and data collection
385 time $t$:
386 \begin{equation}
387 J_z = \frac{E_{total}}{2 t L_x L_y}
388 \end{equation}
389 And the temperature gradient ${\langle\partial T/\partial z\rangle}$
390 can be obtained by a linear regression of the temperature
391 profile. From the thermal conductivity $\lambda$ calculated, one can
392 further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
393 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
394
395 \subsection{ Water / Metal Thermal Conductivity}
396 Another series of our simulation is the calculation of interfacial
397 thermal conductivity of a Au/H$_2$O system. Respective calculations of
398 liquid water (Extended Simple Point Charge model) and crystal gold
399 thermal conductivity were performed and compared with current results
400 to ensure the validity of NIVS-RNEMD. After that, a mixture system was
401 simulated.
402
403 For thermal conductivity calculation of bulk water, a simulation box
404 consisting of 1000 molecules were first equilibrated under ambient
405 pressure and temperature conditions using NPT ensemble, followed by
406 equilibration in fixed volume (NVT). The system was then equilibrated in
407 microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
408 stable thermal gradient was followed. The simulation box was under
409 periodic boundary condition and devided into 10 slabs. Data collection
410 process was similar to Lennard-Jones fluid system.
411
412 Thermal conductivity calculation of bulk crystal gold used a similar
413 protocol. Two types of force field parameters, Embedded Atom Method
414 (EAM) and Quantum Sutten-Chen (QSC) force field were used
415 respectively. The face-centered cubic crystal simulation box consists of
416 2880 Au atoms. The lattice was first allowed volume change to relax
417 under ambient temperature and pressure. Equilibrations in canonical and
418 microcanonical ensemble were followed in order. With the simulation
419 lattice devided evenly into 10 slabs, different thermal gradients were
420 established by applying a set of target thermal transfer flux. Data of
421 the series of thermal gradients was collected for calculation.
422
423 After simulations of bulk water and crystal gold, a mixture system was
424 constructed, consisting of 1188 Au atoms and 1862 H$_2$O
425 molecules. Spohr potential was adopted in depicting the interaction
426 between metal atom and water molecule.\cite{ISI:000167766600035} A
427 similar protocol of equilibration was followed. Several thermal
428 gradients was built under different target thermal flux. It was found
429 out that compared to our previous simulation systems, the two phases
430 could have large temperature difference even under a relatively low
431 thermal flux. Therefore, under our low flux conditions, it is assumed
432 that the metal and water phases have respectively homogeneous
433 temperature, excluding the surface regions. In calculating the
434 interfacial thermal conductivity $G$, this assumptioin was applied and
435 thus our formula becomes:
436
437 \begin{equation}
438 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
439 \langle T_{water}\rangle \right)}
440 \label{interfaceCalc}
441 \end{equation}
442 where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
443 and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
444 average observed temperature of gold and water phases respectively.
445
446 \section{Results And Discussions}
447 \subsection{Thermal Conductivity}
448 \subsubsection{Lennard-Jones Fluid}
449 Our thermal conductivity calculations show that scaling method results
450 agree with swapping method. Four different exchange intervals were
451 tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
452 10fs exchange interval, target exchange kinetic energy was set to
453 produce equivalent kinetic energy flux as in swapping method. And
454 similar thermal gradients were observed with similar thermal flux in
455 two simulation methods (Figure \ref{thermalGrad}).
456
457 \begin{table*}
458 \begin{minipage}{\linewidth}
459 \begin{center}
460
461 \caption{Calculation results for thermal conductivity of Lennard-Jones
462 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
463 swap and scale methods at various kinetic energy exchange rates. Results
464 in reduced unit. Errors of calculations in parentheses.}
465
466 \begin{tabular}{ccc}
467 \hline
468 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
469 $\lambda^*_{scale}$\\
470 \hline
471 250 & 7.03(0.34) & 7.30(0.10)\\
472 500 & 7.03(0.14) & 6.95(0.09)\\
473 1000 & 6.91(0.42) & 7.19(0.07)\\
474 2000 & 7.52(0.15) & 7.19(0.28)\\
475 \hline
476 \end{tabular}
477 \label{thermalLJRes}
478 \end{center}
479 \end{minipage}
480 \end{table*}
481
482 \begin{figure}
483 \includegraphics[width=\linewidth]{thermalGrad}
484 \caption{Temperature gradients under various kinetic energy flux of
485 thermal conductivity simulations}
486 \label{thermalGrad}
487 \end{figure}
488
489 During these simulations, molecule velocities were recorded in 1000 of
490 all the snapshots of one single data collection process. These
491 velocity data were used to produce histograms of velocity and speed
492 distribution in different slabs. From these histograms, it is observed
493 that under relatively high unphysical kinetic energy flux, speed and
494 velocity distribution of molecules in slabs where swapping occured
495 could deviate from Maxwell-Boltzmann distribution. Figure
496 \ref{histSwap} illustrates how these distributions deviate from an
497 ideal distribution. In high temperature slab, probability density in
498 low speed is confidently smaller than ideal curve fit; in low
499 temperature slab, probability density in high speed is smaller than
500 ideal, while larger than ideal in low speed. This phenomenon is more
501 obvious in our high swapping rate simulations. And this deviation
502 could also leads to deviation of distribution of velocity in various
503 dimensions. One feature of these deviated distribution is that in high
504 temperature slab, the ideal Gaussian peak was changed into a
505 relatively flat plateau; while in low temperature slab, that peak
506 appears sharper. This problem is rooted in the mechanism of the
507 swapping method. Continually depleting low (high) speed particles in
508 the high (low) temperature slab could not be complemented by
509 diffusions of low (high) speed particles from neighbor slabs, unless
510 in suffciently low swapping rate. Simutaneously, surplus low speed
511 particles in the low temperature slab do not have sufficient time to
512 diffuse to neighbor slabs. However, thermal exchange rate should reach
513 a minimum level to produce an observable thermal gradient under noise
514 interference. Consequently, swapping RNEMD has a relatively narrow
515 choice of swapping rate to satisfy these above restrictions.
516
517 \begin{figure}
518 \includegraphics[width=\linewidth]{histSwap}
519 \caption{Speed distribution for thermal conductivity using swapping
520 RNEMD. Shown is from the simulation with 250 fs exchange interval.}
521 \label{histSwap}
522 \end{figure}
523
524 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
525 curve fit (Figure \ref{histScale}). Essentially, after scaling, a
526 Gaussian distribution function would remain Gaussian. Although a
527 single scaling is non-isotropic in all three dimensions, our scaling
528 coefficient criteria could help maintian the scaling region as
529 isotropic as possible. On the other hand, scaling coefficients are
530 preferred to be as close to 1 as possible, which also helps minimize
531 the difference among different dimensions. This is possible if scaling
532 interval and one-time thermal transfer energy are well
533 chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
534 thermal flux as the previous RNEMD method without large perturbation
535 to the distribution of velocity and speed in the exchange regions.
536
537 \begin{figure}
538 \includegraphics[width=\linewidth]{histScale}
539 \caption{Speed distribution for thermal conductivity using scaling
540 RNEMD. Shown is from the simulation with an equilvalent thermal flux
541 as an 250 fs exchange interval swapping simulation.}
542 \label{histScale}
543 \end{figure}
544
545 \subsubsection{SPC/E Water}
546 Our results of SPC/E water thermal conductivity are comparable to
547 Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
548 previous swapping RNEMD method for their calculation. Bedrov {\it et
549 al.}\cite{ISI:000090151400044} argued that exchange of the molecule
550 center-of-mass velocities instead of single atom velocities in a
551 molecule conserves the total kinetic energy and linear momentum. This
552 principle is adopted in our simulations. Scaling is applied to the
553 velocities of the rigid bodies of SPC/E model water molecules, instead
554 of each hydrogen and oxygen atoms in relevant water molecules. As
555 shown in Figure \ref{spceGrad}, temperature gradients were established
556 similar to their system. However, the average temperature of our
557 system is 300K, while theirs is 318K, which would be attributed for
558 part of the difference between the final calculation results (Table
559 \ref{spceThermal}). Both methods yields values in agreement with
560 experiment. And this shows the applicability of our method to
561 multi-atom molecular system.
562
563 \begin{figure}
564 \includegraphics[width=\linewidth]{spceGrad}
565 \caption{Temperature gradients for SPC/E water thermal conductivity.}
566 \label{spceGrad}
567 \end{figure}
568
569 \begin{table*}
570 \begin{minipage}{\linewidth}
571 \begin{center}
572
573 \caption{Calculation results for thermal conductivity of SPC/E water
574 at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
575 calculations in parentheses. }
576
577 \begin{tabular}{cccc}
578 \hline
579 $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
580 & This work & Previous simulations\cite{ISI:000090151400044} &
581 Experiment$^a$\\
582 \hline
583 0.38 & 0.816(0.044) & & 0.64\\
584 0.81 & 0.770(0.008) & 0.784\\
585 1.54 & 0.813(0.007) & 0.730\\
586 \hline
587 \end{tabular}
588 \label{spceThermal}
589 \end{center}
590 \end{minipage}
591 \end{table*}
592
593 \subsubsection{Crystal Gold}
594 Our results of gold thermal conductivity using two force fields are
595 shown separately in Table \ref{qscThermal} and \ref{eamThermal}. In
596 these calculations,the end and middle slabs were excluded in thermal
597 gradient regession and only used as heat source and drain in the
598 systems. Our yielded values using EAM force field are slightly larger
599 than those using QSC force field. However, both series are
600 significantly smaller than experimental value by an order of more than
601 100. It has been verified that this difference is mainly attributed to
602 the lack of electron interaction representation in these force field
603 parameters. Richardson {\it et al.}\cite{Clancy:1992} used EAM
604 force field parameters in their metal thermal conductivity
605 calculations. The Non-Equilibrium MD method they employed in their
606 simulations produced comparable results to ours. As Zhang {\it et
607 al.}\cite{ISI:000231042800044} stated, thermal conductivity values
608 are influenced mainly by force field. Therefore, it is confident to
609 conclude that NIVS-RNEMD is applicable to metal force field system.
610
611 \begin{figure}
612 \includegraphics[width=\linewidth]{AuGrad}
613 \caption{Temperature gradients for thermal conductivity calculation of
614 crystal gold using QSC force field.}
615 \label{AuGrad}
616 \end{figure}
617
618 \begin{table*}
619 \begin{minipage}{\linewidth}
620 \begin{center}
621
622 \caption{Calculation results for thermal conductivity of crystal gold
623 using QSC force field at ${\langle T\rangle}$ = 300K at various
624 thermal exchange rates. Errors of calculations in parentheses. }
625
626 \begin{tabular}{cc}
627 \hline
628 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
629 \hline
630 1.44 & 1.10(0.01)\\
631 2.86 & 1.08(0.02)\\
632 5.14 & 1.15(0.01)\\
633 \hline
634 \end{tabular}
635 \label{qscThermal}
636 \end{center}
637 \end{minipage}
638 \end{table*}
639
640 \begin{figure}
641 \includegraphics[width=\linewidth]{eamGrad}
642 \caption{Temperature gradients for thermal conductivity calculation of
643 crystal gold using EAM force field.}
644 \label{eamGrad}
645 \end{figure}
646
647 \begin{table*}
648 \begin{minipage}{\linewidth}
649 \begin{center}
650
651 \caption{Calculation results for thermal conductivity of crystal gold
652 using EAM force field at ${\langle T\rangle}$ = 300K at various
653 thermal exchange rates. Errors of calculations in parentheses. }
654
655 \begin{tabular}{cc}
656 \hline
657 $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
658 \hline
659 1.24 & 1.24(0.06)\\
660 2.06 & 1.37(0.04)\\
661 2.55 & 1.41(0.03)\\
662 \hline
663 \end{tabular}
664 \label{eamThermal}
665 \end{center}
666 \end{minipage}
667 \end{table*}
668
669
670 \subsection{Interfaciel Thermal Conductivity}
671 After simulations of homogeneous water and gold systems using
672 NIVS-RNEMD method were proved valid, calculation of gold/water
673 interfacial thermal conductivity was followed. It is found out that
674 the low interfacial conductance is probably due to the hydrophobic
675 surface in our system. Figure \ref{interfaceDensity} demonstrates mass
676 density change along $z$-axis, which is perpendicular to the
677 gold/water interface. It is observed that water density significantly
678 decreases when approaching the surface. Under this low thermal
679 conductance, both gold and water phase have sufficient time to
680 eliminate temperature difference inside respectively (Figure
681 \ref{interfaceGrad}). With indistinguishable temperature difference
682 within respective phase, it is valid to assume that the temperature
683 difference between gold and water on surface would be approximately
684 the same as the difference between the gold and water phase. This
685 assumption enables convenient calculation of $G$ using
686 Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
687 layer of water and gold close enough to surface, which would have
688 greater fluctuation and lower accuracy. Reported results (Table
689 \ref{interfaceRes}) are of two orders of magnitude smaller than our
690 calculations on homogeneous systems, and thus have larger relative
691 errors than our calculation results on homogeneous systems.
692
693 \begin{figure}
694 \includegraphics[width=\linewidth]{interfaceDensity}
695 \caption{Density profile for interfacial thermal conductivity
696 simulation box. Significant water density decrease is observed on
697 gold surface.}
698 \label{interfaceDensity}
699 \end{figure}
700
701 \begin{figure}
702 \includegraphics[width=\linewidth]{interfaceGrad}
703 \caption{Temperature profiles for interfacial thermal conductivity
704 simulation box. Temperatures of different slabs in the same phase
705 show no significant difference.}
706 \label{interfaceGrad}
707 \end{figure}
708
709 \begin{table*}
710 \begin{minipage}{\linewidth}
711 \begin{center}
712
713 \caption{Calculation results for interfacial thermal conductivity
714 at ${\langle T\rangle \sim}$ 300K at various thermal exchange
715 rates. Errors of calculations in parentheses. }
716
717 \begin{tabular}{cccc}
718 \hline
719 $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
720 \hline
721 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
722 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
723 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
724 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
725 \hline
726 \end{tabular}
727 \label{interfaceRes}
728 \end{center}
729 \end{minipage}
730 \end{table*}
731
732 \subsection{Shear Viscosity}
733 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
734 produced comparable shear viscosity to swap RNEMD method. In Table
735 \ref{shearRate}, the names of the calculated samples are devided into
736 two parts. The first number refers to total slabs in one simulation
737 box. The second number refers to the swapping interval in swap method, or
738 in scale method the equilvalent swapping interval that the same
739 momentum flux would theoretically result in swap method. All the scale
740 method results were from simulations that had a scaling interval of 10
741 time steps. The average molecular momentum gradients of these samples
742 are shown in Figure \ref{shearGrad}.
743
744 \begin{table*}
745 \begin{minipage}{\linewidth}
746 \begin{center}
747
748 \caption{Calculation results for shear viscosity of Lennard-Jones
749 fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
750 methods at various momentum exchange rates. Results in reduced
751 unit. Errors of calculations in parentheses. }
752
753 \begin{tabular}{ccc}
754 \hline
755 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
756 \hline
757 20-500 & 3.64(0.05) & 3.76(0.09)\\
758 20-1000 & 3.52(0.16) & 3.66(0.06)\\
759 20-2000 & 3.72(0.05) & 3.32(0.18)\\
760 20-2500 & 3.42(0.06) & 3.43(0.08)\\
761 \hline
762 \end{tabular}
763 \label{shearRate}
764 \end{center}
765 \end{minipage}
766 \end{table*}
767
768 \begin{figure}
769 \includegraphics[width=\linewidth]{shearGrad}
770 \caption{Average momentum gradients of shear viscosity simulations}
771 \label{shearGrad}
772 \end{figure}
773
774 \begin{figure}
775 \includegraphics[width=\linewidth]{shearTempScale}
776 \caption{Temperature profile for scaling RNEMD simulation.}
777 \label{shearTempScale}
778 \end{figure}
779 However, observations of temperatures along three dimensions show that
780 inhomogeneity occurs in scaling RNEMD simulations, particularly in the
781 two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
782 relatively large imposed momentum flux, the temperature difference among $x$
783 and the other two dimensions was significant. This would result from the
784 algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
785 momentum gradient is set up, $P_c^x$ would be roughly stable
786 ($<0$). Consequently, scaling factor $x$ would most probably larger
787 than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
788 keep increase after most scaling steps. And if there is not enough time
789 for the kinetic energy to exchange among different dimensions and
790 different slabs, the system would finally build up temperature
791 (kinetic energy) difference among the three dimensions. Also, between
792 $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
793 are closer to neighbor slabs. This is due to momentum transfer along
794 $z$ dimension between slabs.
795
796 Although results between scaling and swapping methods are comparable,
797 the inherent temperature inhomogeneity even in relatively low imposed
798 exchange momentum flux simulations makes scaling RNEMD method less
799 attractive than swapping RNEMD in shear viscosity calculation.
800
801 \section{Conclusions}
802 NIVS-RNEMD simulation method is developed and tested on various
803 systems. Simulation results demonstrate its validity in thermal
804 conductivity calculations, from Lennard-Jones fluid to multi-atom
805 molecule like water and metal crystals. NIVS-RNEMD improves
806 non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
807 methods. Furthermore, it develops a valid means for unphysical thermal
808 transfer between different species of molecules, and thus extends its
809 applicability to interfacial systems. Our calculation of gold/water
810 interfacial thermal conductivity demonstrates this advantage over
811 previous RNEMD methods. NIVS-RNEMD has also limited application on
812 shear viscosity calculations, but could cause temperature difference
813 among different dimensions under high momentum flux. Modification is
814 necessary to extend the applicability of NIVS-RNEMD in shear viscosity
815 calculations.
816
817 \section{Acknowledgments}
818 Support for this project was provided by the National Science
819 Foundation under grant CHE-0848243. Computational time was provided by
820 the Center for Research Computing (CRC) at the University of Notre
821 Dame. \newpage
822
823 \bibliographystyle{aip}
824 \bibliography{nivsRnemd}
825
826 \end{doublespace}
827 \end{document}
828