ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3575
Committed: Sat Mar 13 02:08:11 2010 UTC (14 years, 6 months ago) by skuang
Content type: application/x-tex
File size: 29920 byte(s)
Log Message:
a few changes in intro and method part.

File Contents

# User Rev Content
1 gezelter 3524 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5     \usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{tabularx}
8     \usepackage{graphicx}
9     %\usepackage{booktabs}
10     %\usepackage{bibentry}
11     %\usepackage{mathrsfs}
12     \usepackage[ref]{overcite}
13     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
14     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
15     9.0in \textwidth 6.5in \brokenpenalty=10000
16    
17     % double space list of tables and figures
18     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
19     \setlength{\abovecaptionskip}{20 pt}
20     \setlength{\belowcaptionskip}{30 pt}
21    
22     \renewcommand\citemid{\ } % no comma in optional referenc note
23    
24     \begin{document}
25    
26     \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
27    
28     \author{Shenyu Kuang and J. Daniel
29     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
30     Department of Chemistry and Biochemistry,\\
31     University of Notre Dame\\
32     Notre Dame, Indiana 46556}
33    
34     \date{\today}
35    
36     \maketitle
37    
38     \begin{doublespace}
39    
40     \begin{abstract}
41    
42     \end{abstract}
43    
44     \newpage
45    
46     %\narrowtext
47    
48     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
49     % BODY OF TEXT
50     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51    
52    
53    
54     \section{Introduction}
55     The original formulation of Reverse Non-equilibrium Molecular Dynamics
56     (RNEMD) obtains transport coefficients (thermal conductivity and shear
57     viscosity) in a fluid by imposing an artificial momentum flux between
58     two thin parallel slabs of material that are spatially separated in
59 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
60 skuang 3531 artificial flux is typically created by periodically ``swapping'' either
61 gezelter 3524 the entire momentum vector $\vec{p}$ or single components of this
62     vector ($p_x$) between molecules in each of the two slabs. If the two
63 skuang 3575 slabs are separated along the $z$ coordinate, the imposed flux is either
64 skuang 3532 directional ($j_z(p_x)$) or isotropic ($J_z$), and the response of a
65 gezelter 3524 simulated system to the imposed momentum flux will typically be a
66 skuang 3575 velocity or thermal gradient (Fig. \ref{thermalDemo}). The transport
67     coefficients (shear viscosity and thermal conductivity) are easily
68     obtained by assuming linear response of the system,
69 gezelter 3524 \begin{eqnarray}
70 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
71 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
72 gezelter 3524 \end{eqnarray}
73 skuang 3528 RNEMD has been widely used to provide computational estimates of thermal
74 gezelter 3524 conductivities and shear viscosities in a wide range of materials,
75     from liquid copper to monatomic liquids to molecular fluids
76 skuang 3528 (e.g. ionic liquids).\cite{ISI:000246190100032}
77 gezelter 3524
78 skuang 3574 \begin{figure}
79     \includegraphics[width=\linewidth]{thermalDemo}
80     \caption{Demostration of thermal gradient estalished by RNEMD method.}
81     \label{thermalDemo}
82     \end{figure}
83    
84 gezelter 3524 RNEMD is preferable in many ways to the forward NEMD methods because
85     it imposes what is typically difficult to measure (a flux or stress)
86     and it is typically much easier to compute momentum gradients or
87     strains (the response). For similar reasons, RNEMD is also preferable
88     to slowly-converging equilibrium methods for measuring thermal
89     conductivity and shear viscosity (using Green-Kubo relations or the
90     Helfand moment approach of Viscardy {\it et
91 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
92 gezelter 3524 computing difficult to measure quantities.
93    
94     Another attractive feature of RNEMD is that it conserves both total
95     linear momentum and total energy during the swaps (as long as the two
96     molecules have the same identity), so the swapped configurations are
97     typically samples from the same manifold of states in the
98     microcanonical ensemble.
99    
100 skuang 3565 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
101     some problems with the original RNEMD swap technique. Notably, large
102     momentum fluxes (equivalent to frequent momentum swaps between the
103 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
104     momentum distributions in the two slabs, as well as non-linear thermal
105     and velocity distributions along the direction of the imposed flux
106     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
107     and self-adjusting metrics for retaining the usability of the method.
108 gezelter 3524
109     In this paper, we develop and test a method for non-isotropic velocity
110     scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
111     (conservation of linear momentum and total energy, compatibility with
112     periodic boundary conditions) while establishing true thermal
113     distributions in each of the two slabs. In the next section, we
114     develop the method for determining the scaling constraints. We then
115     test the method on both single component, multi-component, and
116     non-isotropic mixtures and show that it is capable of providing
117     reasonable estimates of the thermal conductivity and shear viscosity
118     in these cases.
119    
120     \section{Methodology}
121     We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
122     system is partitioned into a series of thin slabs along a particular
123     axis ($z$). One of the slabs at the end of the periodic box is
124     designated the ``hot'' slab, while the slab in the center of the box
125     is designated the ``cold'' slab. The artificial momentum flux will be
126     established by transferring momentum from the cold slab and into the
127     hot slab.
128    
129     Rather than using momentum swaps, we use a series of velocity scaling
130 skuang 3528 moves. For molecules $\{i\}$ located within the cold slab,
131 gezelter 3524 \begin{equation}
132 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
133     x & 0 & 0 \\
134     0 & y & 0 \\
135     0 & 0 & z \\
136 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
137     \end{equation}
138     where ${x, y, z}$ are a set of 3 scaling variables for each of the
139     three directions in the system. Likewise, the molecules $\{j\}$
140 skuang 3528 located in the hot slab will see a concomitant scaling of velocities,
141 gezelter 3524 \begin{equation}
142 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
143     x^\prime & 0 & 0 \\
144     0 & y^\prime & 0 \\
145     0 & 0 & z^\prime \\
146 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
147     \end{equation}
148    
149     Conservation of linear momentum in each of the three directions
150     ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
151     parameters together:
152     \begin{equation}
153 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
154 gezelter 3524 \end{equation}
155     where
156 skuang 3565 \begin{eqnarray}
157 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
158 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
159 gezelter 3524 \label{eq:momentumdef}
160 skuang 3565 \end{eqnarray}
161 skuang 3528 Therefore, for each of the three directions, the hot scaling
162     parameters are a simple function of the cold scaling parameters and
163 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
164     \begin{equation}
165 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
166 gezelter 3524 \label{eq:hotcoldscaling}
167     \end{equation}
168 skuang 3528 where
169     \begin{equation}
170     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
171     \end{equation}
172     for convenience.
173 gezelter 3524
174     Conservation of total energy also places constraints on the scaling:
175     \begin{equation}
176     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
177 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
178 gezelter 3524 \end{equation}
179 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
180     $K_c^\alpha$, are computed for each of the three directions in a
181     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
182     Substituting in the expressions for the hot scaling parameters
183     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
184     {\it constraint ellipsoid equation}:
185 gezelter 3524 \begin{equation}
186 skuang 3565 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
187 gezelter 3524 \label{eq:constraintEllipsoid}
188     \end{equation}
189     where the constants are obtained from the instantaneous values of the
190     linear momenta and kinetic energies for the hot and cold slabs,
191 skuang 3565 \begin{eqnarray}
192 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
193     \left(p_\alpha\right)^2\right) \\
194     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
195 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
196 gezelter 3524 \label{eq:constraintEllipsoidConsts}
197 skuang 3565 \end{eqnarray}
198 skuang 3528 This ellipsoid equation defines the set of cold slab scaling
199     parameters which can be applied while preserving both linear momentum
200 skuang 3530 in all three directions as well as kinetic energy.
201 gezelter 3524
202     The goal of using velocity scaling variables is to transfer linear
203     momentum or kinetic energy from the cold slab to the hot slab. If the
204     hot and cold slabs are separated along the z-axis, the energy flux is
205 skuang 3528 given simply by the decrease in kinetic energy of the cold bin:
206 gezelter 3524 \begin{equation}
207 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
208 gezelter 3524 \end{equation}
209     The expression for the energy flux can be re-written as another
210     ellipsoid centered on $(x,y,z) = 0$:
211     \begin{equation}
212 skuang 3534 x^2 K_c^x + y^2 K_c^y + z^2 K_c^z = K_c^x + K_c^y + K_c^z - J_z\Delta t
213 gezelter 3524 \label{eq:fluxEllipsoid}
214     \end{equation}
215 skuang 3575 The spatial extent of the {\it thermal flux ellipsoid equation} is
216     governed both by a targetted value, $J_z$ as well as the instantaneous
217     values of the kinetic energy components in the cold bin.
218 gezelter 3524
219     To satisfy an energetic flux as well as the conservation constraints,
220     it is sufficient to determine the points ${x,y,z}$ which lie on both
221     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
222     flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
223 skuang 3528 the two ellipsoids in 3-dimensional space.
224 gezelter 3524
225 gezelter 3569 \begin{figure}
226     \includegraphics[width=\linewidth]{ellipsoids}
227     \caption{Scaling points which maintain both constant energy and
228     constant linear momentum of the system lie on the surface of the
229     {\it constraint ellipsoid} while points which generate the target
230     momentum flux lie on the surface of the {\it flux ellipsoid}. The
231 skuang 3575 velocity distributions in the cold bin are scaled by only those
232 gezelter 3569 points which lie on both ellipsoids.}
233     \label{ellipsoids}
234     \end{figure}
235    
236 skuang 3575 One may also define momentum flux (say along the $x$-direction) as:
237 gezelter 3524 \begin{equation}
238 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
239 skuang 3531 \label{eq:fluxPlane}
240 gezelter 3524 \end{equation}
241 skuang 3575 The above {\it momentum flux equation} is essentially a plane which is
242     perpendicular to the $x$-axis, with its position governed both by a
243     target value, $j_z(p_x)$ as well as the instantaneous value of the
244     momentum along the $x$-direction.
245 gezelter 3524
246 skuang 3531 Similarly, to satisfy a momentum flux as well as the conservation
247     constraints, it is sufficient to determine the points ${x,y,z}$ which
248     lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
249     and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
250     an ellipsoid and a plane in 3-dimensional space.
251 gezelter 3524
252 skuang 3531 To summarize, by solving respective equation sets, one can determine
253     possible sets of scaling variables for cold slab. And corresponding
254     sets of scaling variables for hot slab can be determine as well.
255 gezelter 3524
256 skuang 3531 The following problem will be choosing an optimal set of scaling
257     variables among the possible sets. Although this method is inherently
258     non-isotropic, the goal is still to maintain the system as isotropic
259     as possible. Under this consideration, one would like the kinetic
260     energies in different directions could become as close as each other
261     after each scaling. Simultaneously, one would also like each scaling
262     as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
263     large perturbation to the system. Therefore, one approach to obtain the
264     scaling variables would be constructing an criteria function, with
265     constraints as above equation sets, and solving the function's minimum
266     by method like Lagrange multipliers.
267 gezelter 3524
268 skuang 3531 In order to save computation time, we have a different approach to a
269     relatively good set of scaling variables with much less calculation
270     than above. Here is the detail of our simplification of the problem.
271 gezelter 3524
272 skuang 3531 In the case of kinetic energy transfer, we impose another constraint
273     ${x = y}$, into the equation sets. Consequently, there are two
274     variables left. And now one only needs to solve a set of two {\it
275     ellipses equations}. This problem would be transformed into solving
276     one quartic equation for one of the two variables. There are known
277     generic methods that solve real roots of quartic equations. Then one
278     can determine the other variable and obtain sets of scaling
279     variables. Among these sets, one can apply the above criteria to
280     choose the best set, while much faster with only a few sets to choose.
281    
282     In the case of momentum flux transfer, we impose another constraint to
283     set the kinetic energy transfer as zero. In another word, we apply
284     Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
285     variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
286     of equations on the above kinetic energy transfer problem. Therefore,
287     an approach similar to the above would be sufficient for this as well.
288    
289     \section{Computational Details}
290 skuang 3534 Our simulation consists of a series of systems. All of these
291 skuang 3565 simulations were run with the OpenMD simulation software
292 skuang 3534 package\cite{Meineke:2005gd} integrated with RNEMD methods.
293 skuang 3531
294 skuang 3532 A Lennard-Jones fluid system was built and tested first. In order to
295     compare our method with swapping RNEMD, a series of simulations were
296     performed to calculate the shear viscosity and thermal conductivity of
297 skuang 3534 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
298     \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
299     ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
300     comparison between our results and others. These simulations used
301 skuang 3565 velocity Verlet algorithm with reduced timestep ${\tau^* =
302 skuang 3534 4.6\times10^{-4}}$.
303 skuang 3532
304     For shear viscosity calculation, the reduced temperature was ${T^* =
305 skuang 3565 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
306     ensemble (NVT), then equilibrated in microcanonical ensemble
307     (NVE). Establishing and stablizing momentum gradient were followed
308     also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
309 skuang 3532 adopted.\cite{ISI:000080382700030} The simulation box was under
310 skuang 3534 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
311     the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
312     most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
313 skuang 3565 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
314 skuang 3534 frequency were chosen. According to each result from swapping
315 skuang 3532 RNEMD, scaling RNEMD simulations were run with the target momentum
316 skuang 3534 flux set to produce a similar momentum flux and shear
317     rate. Furthermore, various scaling frequencies can be tested for one
318     single swapping rate. To compare the performance between swapping and
319     scaling method, temperatures of different dimensions in all the slabs
320 skuang 3538 were observed. Most of the simulations include $10^5$ steps of
321     equilibration without imposing momentum flux, $10^5$ steps of
322     stablization with imposing momentum transfer, and $10^6$ steps of data
323     collection under RNEMD. For relatively high momentum flux simulations,
324     ${5\times10^5}$ step data collection is sufficient. For some low momentum
325     flux simulations, ${2\times10^6}$ steps were necessary.
326 skuang 3532
327 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
328     unit. The momentum flux was calculated with total unphysical
329 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
330 skuang 3534 \begin{equation}
331     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
332     \end{equation}
333     And the velocity gradient ${\langle \partial v_x /\partial z \rangle}$
334     can be obtained by a linear regression of the velocity profile. From
335     the shear viscosity $\eta$ calculated with the above parameters, one
336     can further convert it into reduced unit ${\eta^* = \eta \sigma^2
337     (\varepsilon m)^{-1/2}}$.
338 skuang 3532
339 skuang 3534 For thermal conductivity calculation, simulations were first run under
340     reduced temperature ${T^* = 0.72}$ in NVE ensemble. Muller-Plathe's
341     algorithm was adopted in the swapping method. Under identical
342 skuang 3536 simulation box parameters, in each swap, the top slab exchange the
343     molecule with least kinetic energy with the molecule in the center
344     slab with most kinetic energy, unless this ``coldest'' molecule in the
345     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the ``cold''
346 skuang 3534 slab. According to swapping RNEMD results, target energy flux for
347     scaling RNEMD simulations can be set. Also, various scaling
348     frequencies can be tested for one target energy flux. To compare the
349     performance between swapping and scaling method, distributions of
350     velocity and speed in different slabs were observed.
351    
352     For each swapping rate, thermal conductivity was calculated in reduced
353     unit. The energy flux was calculated similarly to the momentum flux,
354 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
355 skuang 3534 time $t$:
356     \begin{equation}
357     J_z = \frac{E_{total}}{2 t L_x L_y}
358     \end{equation}
359     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
360     can be obtained by a linear regression of the temperature
361     profile. From the thermal conductivity $\lambda$ calculated, one can
362     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
363     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
364    
365 skuang 3565 Another series of our simulation is to calculate the interfacial
366 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
367 skuang 3565 water (SPC/E) and gold (QSC) thermal conductivity were performed and
368 skuang 3563 compared with current results to ensure the validity of
369 skuang 3573 NIVS-RNEMD. After that, a mixture system was simulated.
370 skuang 3563
371 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
372     consisting of 1000 molecules were first equilibrated under ambient
373     pressure and temperature conditions (NPT), followed by equilibration
374     in fixed volume (NVT). The system was then equilibrated in
375     microcanonical ensemble (NVE). Also in NVE ensemble, establishing
376     stable thermal gradient was followed. The simulation box was under
377     periodic boundary condition and devided into 10 slabs. Data collection
378     process was similar to Lennard-Jones fluid system. Thermal
379     conductivity calculation of bulk crystal gold used a similar
380     protocol. And the face centered cubic crystal simulation box consists
381     of 2880 Au atoms.
382    
383     After simulations of bulk water and crystal gold, a mixture system was
384     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
385     molecules. Spohr potential was adopted in depicting the interaction
386     between metal atom and water molecule.\cite{ISI:000167766600035} A
387     similar protocol of equilibration was followed. A thermal gradient was
388     built. It was found out that compared to homogeneous systems, the two
389     phases could have large temperature difference under a relatively low
390     thermal flux. Therefore, under our low flux condition, it is assumed
391     that the metal and water phases have respectively homogeneous
392     temperature. In calculating the interfacial thermal conductivity $G$,
393     this assumptioin was applied and thus our formula becomes:
394    
395     \begin{equation}
396     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
397     \langle T_{water}\rangle \right)}
398     \label{interfaceCalc}
399     \end{equation}
400     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
401     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
402     average observed temperature of gold and water phases respectively.
403    
404 skuang 3534 \section{Results And Discussion}
405     \subsection{Shear Viscosity}
406 skuang 3538 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
407     produced comparable shear viscosity to swap RNEMD method. In Table
408     \ref{shearRate}, the names of the calculated samples are devided into
409     two parts. The first number refers to total slabs in one simulation
410     box. The second number refers to the swapping interval in swap method, or
411     in scale method the equilvalent swapping interval that the same
412     momentum flux would theoretically result in swap method. All the scale
413 skuang 3563 method results were from simulations that had a scaling interval of 10
414     time steps. The average molecular momentum gradients of these samples
415 skuang 3565 are shown in Figure \ref{shearGrad}.
416 skuang 3534
417 skuang 3538 \begin{table*}
418     \begin{minipage}{\linewidth}
419     \begin{center}
420    
421     \caption{Calculation results for shear viscosity of Lennard-Jones
422     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
423     methods at various momentum exchange rates. Results in reduced
424     unit. Errors of calculations in parentheses. }
425    
426 skuang 3565 \begin{tabular}{ccc}
427 skuang 3538 \hline
428 skuang 3565 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
429 skuang 3538 \hline
430     20-500 & 3.64(0.05) & 3.76(0.09)\\
431 skuang 3539 20-1000 & 3.52(0.16) & 3.66(0.06)\\
432     20-2000 & 3.72(0.05) & 3.32(0.18)\\
433 skuang 3566 20-2500 & 3.42(0.06) & 3.43(0.08)\\
434     \hline
435 skuang 3538 \end{tabular}
436     \label{shearRate}
437     \end{center}
438     \end{minipage}
439     \end{table*}
440    
441     \begin{figure}
442 skuang 3565 \includegraphics[width=\linewidth]{shearGrad}
443     \caption{Average momentum gradients of shear viscosity simulations}
444     \label{shearGrad}
445 skuang 3538 \end{figure}
446    
447     \begin{figure}
448 skuang 3565 \includegraphics[width=\linewidth]{shearTempScale}
449 skuang 3538 \caption{Temperature profile for scaling RNEMD simulation.}
450     \label{shearTempScale}
451     \end{figure}
452     However, observations of temperatures along three dimensions show that
453     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
454     two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
455 skuang 3563 relatively large imposed momentum flux, the temperature difference among $x$
456     and the other two dimensions was significant. This would result from the
457     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
458     momentum gradient is set up, $P_c^x$ would be roughly stable
459     ($<0$). Consequently, scaling factor $x$ would most probably larger
460     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
461     keep increase after most scaling steps. And if there is not enough time
462     for the kinetic energy to exchange among different dimensions and
463     different slabs, the system would finally build up temperature
464     (kinetic energy) difference among the three dimensions. Also, between
465     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
466     are closer to neighbor slabs. This is due to momentum transfer along
467     $z$ dimension between slabs.
468 skuang 3538
469     Although results between scaling and swapping methods are comparable,
470 skuang 3563 the inherent temperature inhomogeneity even in relatively low imposed
471     exchange momentum flux simulations makes scaling RNEMD method less
472 skuang 3538 attractive than swapping RNEMD in shear viscosity calculation.
473    
474     \subsection{Thermal Conductivity}
475 skuang 3573 \subsubsection{Lennard-Jones Fluid}
476 skuang 3567 Our thermal conductivity calculations also show that scaling method results
477     agree with swapping method. Table \ref{thermal} lists our simulation
478 skuang 3563 results with similar manner we used in shear viscosity
479     calculation. All the data reported from scaling method were obtained
480     by simulations of 10-step exchange frequency, and the target exchange
481     kinetic energy were set to produce equivalent kinetic energy flux as
482 skuang 3567 in swapping method. Figure \ref{thermalGrad} exhibits similar thermal
483     gradients of respective similar kinetic energy flux.
484 skuang 3538
485 skuang 3563 \begin{table*}
486     \begin{minipage}{\linewidth}
487     \begin{center}
488 skuang 3538
489 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
490 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
491 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
492     in reduced unit. Errors of calculations in parentheses.}
493    
494 skuang 3565 \begin{tabular}{ccc}
495 skuang 3563 \hline
496 skuang 3567 Series & $\lambda^*_{swap}$ & $\lambda^*_{scale}$\\
497 skuang 3565 \hline
498 skuang 3564 20-250 & 7.03(0.34) & 7.30(0.10)\\
499     20-500 & 7.03(0.14) & 6.95(0.09)\\
500 skuang 3563 20-1000 & 6.91(0.42) & 7.19(0.07)\\
501 skuang 3566 20-2000 & 7.52(0.15) & 7.19(0.28)\\
502     \hline
503 skuang 3563 \end{tabular}
504     \label{thermal}
505     \end{center}
506     \end{minipage}
507     \end{table*}
508    
509     \begin{figure}
510 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
511     \caption{Temperature gradients of thermal conductivity simulations}
512     \label{thermalGrad}
513 skuang 3563 \end{figure}
514    
515     During these simulations, molecule velocities were recorded in 1000 of
516     all the snapshots. These velocity data were used to produce histograms
517     of velocity and speed distribution in different slabs. From these
518     histograms, it is observed that with increasing unphysical kinetic
519     energy flux, speed and velocity distribution of molecules in slabs
520     where swapping occured could deviate from Maxwell-Boltzmann
521     distribution. Figure \ref{histSwap} indicates how these distributions
522     deviate from ideal condition. In high temperature slabs, probability
523     density in low speed is confidently smaller than ideal distribution;
524     in low temperature slabs, probability density in high speed is smaller
525     than ideal. This phenomenon is observable even in our relatively low
526 skuang 3568 swapping rate simulations. And this deviation could also leads to
527 skuang 3563 deviation of distribution of velocity in various dimensions. One
528     feature of these deviated distribution is that in high temperature
529     slab, the ideal Gaussian peak was changed into a relatively flat
530     plateau; while in low temperature slab, that peak appears sharper.
531    
532     \begin{figure}
533 skuang 3565 \includegraphics[width=\linewidth]{histSwap}
534 skuang 3563 \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
535     \label{histSwap}
536     \end{figure}
537    
538 skuang 3568 \begin{figure}
539     \includegraphics[width=\linewidth]{histScale}
540     \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
541     \label{histScale}
542     \end{figure}
543    
544 skuang 3573 \subsubsection{SPC/E Water}
545     Our results of SPC/E water thermal conductivity are comparable to
546     Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
547     previous swapping RNEMD method for their calculation. Our simulations
548     were able to produce a similar temperature gradient to their
549     system. However, the average temperature of our system is 300K, while
550     theirs is 318K, which would be attributed for part of the difference
551     between the two series of results. Both methods yields values in
552     agreement with experiment. And this shows the applicability of our
553     method to multi-atom molecular system.
554 skuang 3563
555 skuang 3570 \begin{figure}
556     \includegraphics[width=\linewidth]{spceGrad}
557     \caption{Temperature gradients for SPC/E water thermal conductivity.}
558     \label{spceGrad}
559     \end{figure}
560    
561     \begin{table*}
562     \begin{minipage}{\linewidth}
563     \begin{center}
564    
565     \caption{Calculation results for thermal conductivity of SPC/E water
566     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
567     calculations in parentheses. }
568    
569     \begin{tabular}{cccc}
570     \hline
571     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
572 skuang 3573 & This work & Previous simulations\cite{ISI:000090151400044} &
573     Experiment$^a$\\
574 skuang 3570 \hline
575 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
576     0.81 & 0.770(0.008) & 0.784\\
577     1.54 & 0.813(0.007) & 0.730\\
578 skuang 3570 \hline
579     \end{tabular}
580     \label{spceThermal}
581     \end{center}
582     \end{minipage}
583     \end{table*}
584    
585 skuang 3573 \subsubsection{Crystal Gold}
586 skuang 3574 Our results of gold thermal conductivity used QSC force field are
587     shown in Table \ref{AuThermal}. Although our calculation is smaller
588     than experimental value by an order of more than 100, this difference
589     is mainly attributed to the lack of electron interaction
590     representation in our force field parameters. Richardson {\it et
591     al.}\cite{ISI:A1992HX37800010} used similar force field parameters
592     in their metal thermal conductivity calculations. The EMD method they
593     employed in their simulations produced comparable results to
594     ours. Therefore, it is confident to conclude that NIVS-RNEMD is
595     applicable to metal force field system.
596 skuang 3570
597     \begin{figure}
598     \includegraphics[width=\linewidth]{AuGrad}
599     \caption{Temperature gradients for crystal gold thermal conductivity.}
600     \label{AuGrad}
601     \end{figure}
602    
603     \begin{table*}
604     \begin{minipage}{\linewidth}
605     \begin{center}
606    
607     \caption{Calculation results for thermal conductivity of crystal gold
608     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
609     calculations in parentheses. }
610    
611     \begin{tabular}{ccc}
612     \hline
613     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
614 skuang 3573 & This work & Previous simulations\cite{ISI:A1992HX37800010} \\
615 skuang 3570 \hline
616 skuang 3571 1.44 & 1.10(0.01) & \\
617     2.86 & 1.08(0.02) & \\
618     5.14 & 1.15(0.01) & \\
619 skuang 3570 \hline
620     \end{tabular}
621     \label{AuThermal}
622     \end{center}
623     \end{minipage}
624     \end{table*}
625    
626 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
627 skuang 3574 After valid simulations of homogeneous water and gold systems using
628     NIVS-RNEMD method, calculation of gold/water interfacial thermal
629     conductivity was followed. It is found out that the interfacial
630     conductance is low due to a hydrophobic surface in our system. Figure
631     \ref{interfaceDensity} demonstrates this observance. Consequently, our
632     reported results (Table \ref{interfaceRes}) are of two orders of
633     magnitude smaller than our calculations on homogeneous systems.
634 skuang 3573
635 skuang 3571 \begin{figure}
636     \includegraphics[width=\linewidth]{interfaceDensity}
637     \caption{Density profile for interfacial thermal conductivity
638     simulation box.}
639     \label{interfaceDensity}
640     \end{figure}
641    
642 skuang 3572 \begin{figure}
643     \includegraphics[width=\linewidth]{interfaceGrad}
644     \caption{Temperature profiles for interfacial thermal conductivity
645     simulation box.}
646     \label{interfaceGrad}
647     \end{figure}
648    
649     \begin{table*}
650     \begin{minipage}{\linewidth}
651     \begin{center}
652    
653     \caption{Calculation results for interfacial thermal conductivity
654     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
655     rates. Errors of calculations in parentheses. }
656    
657     \begin{tabular}{cccc}
658     \hline
659     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
660     \hline
661 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
662     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
663     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
664     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
665 skuang 3572 \hline
666     \end{tabular}
667 skuang 3574 \label{interfaceRes}
668 skuang 3572 \end{center}
669     \end{minipage}
670     \end{table*}
671    
672 skuang 3574 \section{Conclusions}
673     NIVS-RNEMD simulation method is developed and tested on various
674     systems. Simulation results demonstrate its validity of thermal
675     conductivity calculations. NIVS-RNEMD improves non-Boltzmann-Maxwell
676     distributions existing in previous RNEMD methods, and extends its
677     applicability to interfacial systems. NIVS-RNEMD has also limited
678     application on shear viscosity calculations, but under high momentum
679     flux, it could cause temperature difference among different
680     dimensions. Modification is necessary to extend the applicability of
681     NIVS-RNEMD in shear viscosity calculations.
682 skuang 3572
683 gezelter 3524 \section{Acknowledgments}
684     Support for this project was provided by the National Science
685     Foundation under grant CHE-0848243. Computational time was provided by
686     the Center for Research Computing (CRC) at the University of Notre
687     Dame. \newpage
688    
689     \bibliographystyle{jcp2}
690     \bibliography{nivsRnemd}
691     \end{doublespace}
692     \end{document}
693