ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3574
Committed: Fri Mar 12 21:36:16 2010 UTC (14 years, 6 months ago) by skuang
Content type: application/x-tex
File size: 29857 byte(s)
Log Message:
draw a demo for RNEMD. finish a rough version.

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     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     velocity or thermal gradient. The transport coefficients (shear
67     viscosity and thermal conductivity) are easily obtained by assuming
68     linear response of the system,
69     \begin{eqnarray}
70 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
71 gezelter 3524 J & = & \lambda \frac{\partial T}{\partial z}
72     \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     slabs) can result in ``notched'', ``peaked'' and generally non-thermal momentum
104 gezelter 3524 distributions in the two slabs, as well as non-linear thermal and
105     velocity distributions along the direction of the imposed flux ($z$).
106     Tenney and Maginn obtained reasonable limits on imposed flux and
107     self-adjusting metrics for retaining the usability of the method.
108    
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     where the kinetic energies, $K_h^\alpha$ and $K_c^\alpha$, are computed
180     for each of the three directions in a similar manner to the linear momenta
181     (Eq. \ref{eq:momentumdef}). Substituting in the expressions for the
182 skuang 3528 hot scaling parameters ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}),
183     we obtain the {\it constraint ellipsoid equation}:
184 gezelter 3524 \begin{equation}
185 skuang 3565 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
186 gezelter 3524 \label{eq:constraintEllipsoid}
187     \end{equation}
188     where the constants are obtained from the instantaneous values of the
189     linear momenta and kinetic energies for the hot and cold slabs,
190 skuang 3565 \begin{eqnarray}
191 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
192     \left(p_\alpha\right)^2\right) \\
193     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
194 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
195 gezelter 3524 \label{eq:constraintEllipsoidConsts}
196 skuang 3565 \end{eqnarray}
197 skuang 3528 This ellipsoid equation defines the set of cold slab scaling
198     parameters which can be applied while preserving both linear momentum
199 skuang 3530 in all three directions as well as kinetic energy.
200 gezelter 3524
201     The goal of using velocity scaling variables is to transfer linear
202     momentum or kinetic energy from the cold slab to the hot slab. If the
203     hot and cold slabs are separated along the z-axis, the energy flux is
204 skuang 3528 given simply by the decrease in kinetic energy of the cold bin:
205 gezelter 3524 \begin{equation}
206 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
207 gezelter 3524 \end{equation}
208     The expression for the energy flux can be re-written as another
209     ellipsoid centered on $(x,y,z) = 0$:
210     \begin{equation}
211 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
212 gezelter 3524 \label{eq:fluxEllipsoid}
213     \end{equation}
214 skuang 3529 The spatial extent of the {\it flux ellipsoid equation} is governed
215     both by a targetted value, $J_z$ as well as the instantaneous values of the
216 skuang 3530 kinetic energy components in the cold bin.
217 gezelter 3524
218     To satisfy an energetic flux as well as the conservation constraints,
219     it is sufficient to determine the points ${x,y,z}$ which lie on both
220     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
221     flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
222 skuang 3528 the two ellipsoids in 3-dimensional space.
223 gezelter 3524
224 gezelter 3569 \begin{figure}
225     \includegraphics[width=\linewidth]{ellipsoids}
226     \caption{Scaling points which maintain both constant energy and
227     constant linear momentum of the system lie on the surface of the
228     {\it constraint ellipsoid} while points which generate the target
229     momentum flux lie on the surface of the {\it flux ellipsoid}. The
230     velocity distributions in the hot bin are scaled by only those
231     points which lie on both ellipsoids.}
232     \label{ellipsoids}
233     \end{figure}
234    
235 gezelter 3524 One may also define momentum flux (say along the x-direction) as:
236     \begin{equation}
237 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
238 skuang 3531 \label{eq:fluxPlane}
239 gezelter 3524 \end{equation}
240 skuang 3531 The above {\it flux equation} is essentially a plane which is
241     perpendicular to the x-axis, with its position governed both by a
242     targetted value, $j_z(p_x)$ as well as the instantaneous value of the
243     momentum along the x-direction.
244 gezelter 3524
245 skuang 3531 Similarly, to satisfy a momentum flux as well as the conservation
246     constraints, it is sufficient to determine the points ${x,y,z}$ which
247     lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
248     and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
249     an ellipsoid and a plane in 3-dimensional space.
250 gezelter 3524
251 skuang 3531 To summarize, by solving respective equation sets, one can determine
252     possible sets of scaling variables for cold slab. And corresponding
253     sets of scaling variables for hot slab can be determine as well.
254 gezelter 3524
255 skuang 3531 The following problem will be choosing an optimal set of scaling
256     variables among the possible sets. Although this method is inherently
257     non-isotropic, the goal is still to maintain the system as isotropic
258     as possible. Under this consideration, one would like the kinetic
259     energies in different directions could become as close as each other
260     after each scaling. Simultaneously, one would also like each scaling
261     as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
262     large perturbation to the system. Therefore, one approach to obtain the
263     scaling variables would be constructing an criteria function, with
264     constraints as above equation sets, and solving the function's minimum
265     by method like Lagrange multipliers.
266 gezelter 3524
267 skuang 3531 In order to save computation time, we have a different approach to a
268     relatively good set of scaling variables with much less calculation
269     than above. Here is the detail of our simplification of the problem.
270 gezelter 3524
271 skuang 3531 In the case of kinetic energy transfer, we impose another constraint
272     ${x = y}$, into the equation sets. Consequently, there are two
273     variables left. And now one only needs to solve a set of two {\it
274     ellipses equations}. This problem would be transformed into solving
275     one quartic equation for one of the two variables. There are known
276     generic methods that solve real roots of quartic equations. Then one
277     can determine the other variable and obtain sets of scaling
278     variables. Among these sets, one can apply the above criteria to
279     choose the best set, while much faster with only a few sets to choose.
280    
281     In the case of momentum flux transfer, we impose another constraint to
282     set the kinetic energy transfer as zero. In another word, we apply
283     Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
284     variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
285     of equations on the above kinetic energy transfer problem. Therefore,
286     an approach similar to the above would be sufficient for this as well.
287    
288     \section{Computational Details}
289 skuang 3534 Our simulation consists of a series of systems. All of these
290 skuang 3565 simulations were run with the OpenMD simulation software
291 skuang 3534 package\cite{Meineke:2005gd} integrated with RNEMD methods.
292 skuang 3531
293 skuang 3532 A Lennard-Jones fluid system was built and tested first. In order to
294     compare our method with swapping RNEMD, a series of simulations were
295     performed to calculate the shear viscosity and thermal conductivity of
296 skuang 3534 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
297     \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
298     ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
299     comparison between our results and others. These simulations used
300 skuang 3565 velocity Verlet algorithm with reduced timestep ${\tau^* =
301 skuang 3534 4.6\times10^{-4}}$.
302 skuang 3532
303     For shear viscosity calculation, the reduced temperature was ${T^* =
304 skuang 3565 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
305     ensemble (NVT), then equilibrated in microcanonical ensemble
306     (NVE). Establishing and stablizing momentum gradient were followed
307     also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
308 skuang 3532 adopted.\cite{ISI:000080382700030} The simulation box was under
309 skuang 3534 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
310     the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
311     most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
312 skuang 3565 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
313 skuang 3534 frequency were chosen. According to each result from swapping
314 skuang 3532 RNEMD, scaling RNEMD simulations were run with the target momentum
315 skuang 3534 flux set to produce a similar momentum flux and shear
316     rate. Furthermore, various scaling frequencies can be tested for one
317     single swapping rate. To compare the performance between swapping and
318     scaling method, temperatures of different dimensions in all the slabs
319 skuang 3538 were observed. Most of the simulations include $10^5$ steps of
320     equilibration without imposing momentum flux, $10^5$ steps of
321     stablization with imposing momentum transfer, and $10^6$ steps of data
322     collection under RNEMD. For relatively high momentum flux simulations,
323     ${5\times10^5}$ step data collection is sufficient. For some low momentum
324     flux simulations, ${2\times10^6}$ steps were necessary.
325 skuang 3532
326 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
327     unit. The momentum flux was calculated with total unphysical
328 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
329 skuang 3534 \begin{equation}
330     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
331     \end{equation}
332     And the velocity gradient ${\langle \partial v_x /\partial z \rangle}$
333     can be obtained by a linear regression of the velocity profile. From
334     the shear viscosity $\eta$ calculated with the above parameters, one
335     can further convert it into reduced unit ${\eta^* = \eta \sigma^2
336     (\varepsilon m)^{-1/2}}$.
337 skuang 3532
338 skuang 3534 For thermal conductivity calculation, simulations were first run under
339     reduced temperature ${T^* = 0.72}$ in NVE ensemble. Muller-Plathe's
340     algorithm was adopted in the swapping method. Under identical
341 skuang 3536 simulation box parameters, in each swap, the top slab exchange the
342     molecule with least kinetic energy with the molecule in the center
343     slab with most kinetic energy, unless this ``coldest'' molecule in the
344     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the ``cold''
345 skuang 3534 slab. According to swapping RNEMD results, target energy flux for
346     scaling RNEMD simulations can be set. Also, various scaling
347     frequencies can be tested for one target energy flux. To compare the
348     performance between swapping and scaling method, distributions of
349     velocity and speed in different slabs were observed.
350    
351     For each swapping rate, thermal conductivity was calculated in reduced
352     unit. The energy flux was calculated similarly to the momentum flux,
353 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
354 skuang 3534 time $t$:
355     \begin{equation}
356     J_z = \frac{E_{total}}{2 t L_x L_y}
357     \end{equation}
358     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
359     can be obtained by a linear regression of the temperature
360     profile. From the thermal conductivity $\lambda$ calculated, one can
361     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
362     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
363    
364 skuang 3565 Another series of our simulation is to calculate the interfacial
365 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
366 skuang 3565 water (SPC/E) and gold (QSC) thermal conductivity were performed and
367 skuang 3563 compared with current results to ensure the validity of
368 skuang 3573 NIVS-RNEMD. After that, a mixture system was simulated.
369 skuang 3563
370 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
371     consisting of 1000 molecules were first equilibrated under ambient
372     pressure and temperature conditions (NPT), followed by equilibration
373     in fixed volume (NVT). The system was then equilibrated in
374     microcanonical ensemble (NVE). Also in NVE ensemble, establishing
375     stable thermal gradient was followed. The simulation box was under
376     periodic boundary condition and devided into 10 slabs. Data collection
377     process was similar to Lennard-Jones fluid system. Thermal
378     conductivity calculation of bulk crystal gold used a similar
379     protocol. And the face centered cubic crystal simulation box consists
380     of 2880 Au atoms.
381    
382     After simulations of bulk water and crystal gold, a mixture system was
383     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
384     molecules. Spohr potential was adopted in depicting the interaction
385     between metal atom and water molecule.\cite{ISI:000167766600035} A
386     similar protocol of equilibration was followed. A thermal gradient was
387     built. It was found out that compared to homogeneous systems, the two
388     phases could have large temperature difference under a relatively low
389     thermal flux. Therefore, under our low flux condition, it is assumed
390     that the metal and water phases have respectively homogeneous
391     temperature. In calculating the interfacial thermal conductivity $G$,
392     this assumptioin was applied and thus our formula becomes:
393    
394     \begin{equation}
395     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
396     \langle T_{water}\rangle \right)}
397     \label{interfaceCalc}
398     \end{equation}
399     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
400     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
401     average observed temperature of gold and water phases respectively.
402    
403 skuang 3534 \section{Results And Discussion}
404     \subsection{Shear Viscosity}
405 skuang 3538 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
406     produced comparable shear viscosity to swap RNEMD method. In Table
407     \ref{shearRate}, the names of the calculated samples are devided into
408     two parts. The first number refers to total slabs in one simulation
409     box. The second number refers to the swapping interval in swap method, or
410     in scale method the equilvalent swapping interval that the same
411     momentum flux would theoretically result in swap method. All the scale
412 skuang 3563 method results were from simulations that had a scaling interval of 10
413     time steps. The average molecular momentum gradients of these samples
414 skuang 3565 are shown in Figure \ref{shearGrad}.
415 skuang 3534
416 skuang 3538 \begin{table*}
417     \begin{minipage}{\linewidth}
418     \begin{center}
419    
420     \caption{Calculation results for shear viscosity of Lennard-Jones
421     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
422     methods at various momentum exchange rates. Results in reduced
423     unit. Errors of calculations in parentheses. }
424    
425 skuang 3565 \begin{tabular}{ccc}
426 skuang 3538 \hline
427 skuang 3565 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
428 skuang 3538 \hline
429     20-500 & 3.64(0.05) & 3.76(0.09)\\
430 skuang 3539 20-1000 & 3.52(0.16) & 3.66(0.06)\\
431     20-2000 & 3.72(0.05) & 3.32(0.18)\\
432 skuang 3566 20-2500 & 3.42(0.06) & 3.43(0.08)\\
433     \hline
434 skuang 3538 \end{tabular}
435     \label{shearRate}
436     \end{center}
437     \end{minipage}
438     \end{table*}
439    
440     \begin{figure}
441 skuang 3565 \includegraphics[width=\linewidth]{shearGrad}
442     \caption{Average momentum gradients of shear viscosity simulations}
443     \label{shearGrad}
444 skuang 3538 \end{figure}
445    
446     \begin{figure}
447 skuang 3565 \includegraphics[width=\linewidth]{shearTempScale}
448 skuang 3538 \caption{Temperature profile for scaling RNEMD simulation.}
449     \label{shearTempScale}
450     \end{figure}
451     However, observations of temperatures along three dimensions show that
452     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
453     two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
454 skuang 3563 relatively large imposed momentum flux, the temperature difference among $x$
455     and the other two dimensions was significant. This would result from the
456     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
457     momentum gradient is set up, $P_c^x$ would be roughly stable
458     ($<0$). Consequently, scaling factor $x$ would most probably larger
459     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
460     keep increase after most scaling steps. And if there is not enough time
461     for the kinetic energy to exchange among different dimensions and
462     different slabs, the system would finally build up temperature
463     (kinetic energy) difference among the three dimensions. Also, between
464     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
465     are closer to neighbor slabs. This is due to momentum transfer along
466     $z$ dimension between slabs.
467 skuang 3538
468     Although results between scaling and swapping methods are comparable,
469 skuang 3563 the inherent temperature inhomogeneity even in relatively low imposed
470     exchange momentum flux simulations makes scaling RNEMD method less
471 skuang 3538 attractive than swapping RNEMD in shear viscosity calculation.
472    
473     \subsection{Thermal Conductivity}
474 skuang 3573 \subsubsection{Lennard-Jones Fluid}
475 skuang 3567 Our thermal conductivity calculations also show that scaling method results
476     agree with swapping method. Table \ref{thermal} lists our simulation
477 skuang 3563 results with similar manner we used in shear viscosity
478     calculation. All the data reported from scaling method were obtained
479     by simulations of 10-step exchange frequency, and the target exchange
480     kinetic energy were set to produce equivalent kinetic energy flux as
481 skuang 3567 in swapping method. Figure \ref{thermalGrad} exhibits similar thermal
482     gradients of respective similar kinetic energy flux.
483 skuang 3538
484 skuang 3563 \begin{table*}
485     \begin{minipage}{\linewidth}
486     \begin{center}
487 skuang 3538
488 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
489 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
490 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
491     in reduced unit. Errors of calculations in parentheses.}
492    
493 skuang 3565 \begin{tabular}{ccc}
494 skuang 3563 \hline
495 skuang 3567 Series & $\lambda^*_{swap}$ & $\lambda^*_{scale}$\\
496 skuang 3565 \hline
497 skuang 3564 20-250 & 7.03(0.34) & 7.30(0.10)\\
498     20-500 & 7.03(0.14) & 6.95(0.09)\\
499 skuang 3563 20-1000 & 6.91(0.42) & 7.19(0.07)\\
500 skuang 3566 20-2000 & 7.52(0.15) & 7.19(0.28)\\
501     \hline
502 skuang 3563 \end{tabular}
503     \label{thermal}
504     \end{center}
505     \end{minipage}
506     \end{table*}
507    
508     \begin{figure}
509 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
510     \caption{Temperature gradients of thermal conductivity simulations}
511     \label{thermalGrad}
512 skuang 3563 \end{figure}
513    
514     During these simulations, molecule velocities were recorded in 1000 of
515     all the snapshots. These velocity data were used to produce histograms
516     of velocity and speed distribution in different slabs. From these
517     histograms, it is observed that with increasing unphysical kinetic
518     energy flux, speed and velocity distribution of molecules in slabs
519     where swapping occured could deviate from Maxwell-Boltzmann
520     distribution. Figure \ref{histSwap} indicates how these distributions
521     deviate from ideal condition. In high temperature slabs, probability
522     density in low speed is confidently smaller than ideal distribution;
523     in low temperature slabs, probability density in high speed is smaller
524     than ideal. This phenomenon is observable even in our relatively low
525 skuang 3568 swapping rate simulations. And this deviation could also leads to
526 skuang 3563 deviation of distribution of velocity in various dimensions. One
527     feature of these deviated distribution is that in high temperature
528     slab, the ideal Gaussian peak was changed into a relatively flat
529     plateau; while in low temperature slab, that peak appears sharper.
530    
531     \begin{figure}
532 skuang 3565 \includegraphics[width=\linewidth]{histSwap}
533 skuang 3563 \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
534     \label{histSwap}
535     \end{figure}
536    
537 skuang 3568 \begin{figure}
538     \includegraphics[width=\linewidth]{histScale}
539     \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
540     \label{histScale}
541     \end{figure}
542    
543 skuang 3573 \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. Our simulations
547     were able to produce a similar temperature gradient to their
548     system. However, the average temperature of our system is 300K, while
549     theirs is 318K, which would be attributed for part of the difference
550     between the two series of results. Both methods yields values in
551     agreement with experiment. And this shows the applicability of our
552     method to multi-atom molecular system.
553 skuang 3563
554 skuang 3570 \begin{figure}
555     \includegraphics[width=\linewidth]{spceGrad}
556     \caption{Temperature gradients for SPC/E water thermal conductivity.}
557     \label{spceGrad}
558     \end{figure}
559    
560     \begin{table*}
561     \begin{minipage}{\linewidth}
562     \begin{center}
563    
564     \caption{Calculation results for thermal conductivity of SPC/E water
565     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
566     calculations in parentheses. }
567    
568     \begin{tabular}{cccc}
569     \hline
570     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
571 skuang 3573 & This work & Previous simulations\cite{ISI:000090151400044} &
572     Experiment$^a$\\
573 skuang 3570 \hline
574 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
575     0.81 & 0.770(0.008) & 0.784\\
576     1.54 & 0.813(0.007) & 0.730\\
577 skuang 3570 \hline
578     \end{tabular}
579     \label{spceThermal}
580     \end{center}
581     \end{minipage}
582     \end{table*}
583    
584 skuang 3573 \subsubsection{Crystal Gold}
585 skuang 3574 Our results of gold thermal conductivity used QSC force field are
586     shown in Table \ref{AuThermal}. Although our calculation is smaller
587     than experimental value by an order of more than 100, this difference
588     is mainly attributed to the lack of electron interaction
589     representation in our force field parameters. Richardson {\it et
590     al.}\cite{ISI:A1992HX37800010} used similar force field parameters
591     in their metal thermal conductivity calculations. The EMD method they
592     employed in their simulations produced comparable results to
593     ours. Therefore, it is confident to conclude that NIVS-RNEMD is
594     applicable to metal force field system.
595 skuang 3570
596     \begin{figure}
597     \includegraphics[width=\linewidth]{AuGrad}
598     \caption{Temperature gradients for crystal gold thermal conductivity.}
599     \label{AuGrad}
600     \end{figure}
601    
602     \begin{table*}
603     \begin{minipage}{\linewidth}
604     \begin{center}
605    
606     \caption{Calculation results for thermal conductivity of crystal gold
607     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
608     calculations in parentheses. }
609    
610     \begin{tabular}{ccc}
611     \hline
612     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
613 skuang 3573 & This work & Previous simulations\cite{ISI:A1992HX37800010} \\
614 skuang 3570 \hline
615 skuang 3571 1.44 & 1.10(0.01) & \\
616     2.86 & 1.08(0.02) & \\
617     5.14 & 1.15(0.01) & \\
618 skuang 3570 \hline
619     \end{tabular}
620     \label{AuThermal}
621     \end{center}
622     \end{minipage}
623     \end{table*}
624    
625 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
626 skuang 3574 After valid simulations of homogeneous water and gold systems using
627     NIVS-RNEMD method, calculation of gold/water interfacial thermal
628     conductivity was followed. It is found out that the interfacial
629     conductance is low due to a hydrophobic surface in our system. Figure
630     \ref{interfaceDensity} demonstrates this observance. Consequently, our
631     reported results (Table \ref{interfaceRes}) are of two orders of
632     magnitude smaller than our calculations on homogeneous systems.
633 skuang 3573
634 skuang 3571 \begin{figure}
635     \includegraphics[width=\linewidth]{interfaceDensity}
636     \caption{Density profile for interfacial thermal conductivity
637     simulation box.}
638     \label{interfaceDensity}
639     \end{figure}
640    
641 skuang 3572 \begin{figure}
642     \includegraphics[width=\linewidth]{interfaceGrad}
643     \caption{Temperature profiles for interfacial thermal conductivity
644     simulation box.}
645     \label{interfaceGrad}
646     \end{figure}
647    
648     \begin{table*}
649     \begin{minipage}{\linewidth}
650     \begin{center}
651    
652     \caption{Calculation results for interfacial thermal conductivity
653     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
654     rates. Errors of calculations in parentheses. }
655    
656     \begin{tabular}{cccc}
657     \hline
658     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
659     \hline
660 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
661     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
662     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
663     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
664 skuang 3572 \hline
665     \end{tabular}
666 skuang 3574 \label{interfaceRes}
667 skuang 3572 \end{center}
668     \end{minipage}
669     \end{table*}
670    
671 skuang 3574 \section{Conclusions}
672     NIVS-RNEMD simulation method is developed and tested on various
673     systems. Simulation results demonstrate its validity of thermal
674     conductivity calculations. NIVS-RNEMD improves non-Boltzmann-Maxwell
675     distributions existing in previous RNEMD methods, and extends its
676     applicability to interfacial systems. NIVS-RNEMD has also limited
677     application on shear viscosity calculations, but under high momentum
678     flux, it could cause temperature difference among different
679     dimensions. Modification is necessary to extend the applicability of
680     NIVS-RNEMD in shear viscosity calculations.
681 skuang 3572
682 gezelter 3524 \section{Acknowledgments}
683     Support for this project was provided by the National Science
684     Foundation under grant CHE-0848243. Computational time was provided by
685     the Center for Research Computing (CRC) at the University of Notre
686     Dame. \newpage
687    
688     \bibliographystyle{jcp2}
689     \bibliography{nivsRnemd}
690     \end{doublespace}
691     \end{document}
692