ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3573
Committed: Thu Mar 11 23:41:33 2010 UTC (14 years, 6 months ago) by skuang
Content type: application/x-tex
File size: 28000 byte(s)
Log Message:
rearrange some parts

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     RNEMD is preferable in many ways to the forward NEMD methods because
79     it imposes what is typically difficult to measure (a flux or stress)
80     and it is typically much easier to compute momentum gradients or
81     strains (the response). For similar reasons, RNEMD is also preferable
82     to slowly-converging equilibrium methods for measuring thermal
83     conductivity and shear viscosity (using Green-Kubo relations or the
84     Helfand moment approach of Viscardy {\it et
85 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
86 gezelter 3524 computing difficult to measure quantities.
87    
88     Another attractive feature of RNEMD is that it conserves both total
89     linear momentum and total energy during the swaps (as long as the two
90     molecules have the same identity), so the swapped configurations are
91     typically samples from the same manifold of states in the
92     microcanonical ensemble.
93    
94 skuang 3565 Recently, Tenney and Maginn\cite{ISI:000273472300004} have discovered
95     some problems with the original RNEMD swap technique. Notably, large
96     momentum fluxes (equivalent to frequent momentum swaps between the
97     slabs) can result in ``notched'', ``peaked'' and generally non-thermal momentum
98 gezelter 3524 distributions in the two slabs, as well as non-linear thermal and
99     velocity distributions along the direction of the imposed flux ($z$).
100     Tenney and Maginn obtained reasonable limits on imposed flux and
101     self-adjusting metrics for retaining the usability of the method.
102    
103     In this paper, we develop and test a method for non-isotropic velocity
104     scaling (NIVS-RNEMD) which retains the desirable features of RNEMD
105     (conservation of linear momentum and total energy, compatibility with
106     periodic boundary conditions) while establishing true thermal
107     distributions in each of the two slabs. In the next section, we
108     develop the method for determining the scaling constraints. We then
109     test the method on both single component, multi-component, and
110     non-isotropic mixtures and show that it is capable of providing
111     reasonable estimates of the thermal conductivity and shear viscosity
112     in these cases.
113    
114     \section{Methodology}
115     We retain the basic idea of Muller-Plathe's RNEMD method; the periodic
116     system is partitioned into a series of thin slabs along a particular
117     axis ($z$). One of the slabs at the end of the periodic box is
118     designated the ``hot'' slab, while the slab in the center of the box
119     is designated the ``cold'' slab. The artificial momentum flux will be
120     established by transferring momentum from the cold slab and into the
121     hot slab.
122    
123     Rather than using momentum swaps, we use a series of velocity scaling
124 skuang 3528 moves. For molecules $\{i\}$ located within the cold slab,
125 gezelter 3524 \begin{equation}
126 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
127     x & 0 & 0 \\
128     0 & y & 0 \\
129     0 & 0 & z \\
130 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
131     \end{equation}
132     where ${x, y, z}$ are a set of 3 scaling variables for each of the
133     three directions in the system. Likewise, the molecules $\{j\}$
134 skuang 3528 located in the hot slab will see a concomitant scaling of velocities,
135 gezelter 3524 \begin{equation}
136 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
137     x^\prime & 0 & 0 \\
138     0 & y^\prime & 0 \\
139     0 & 0 & z^\prime \\
140 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
141     \end{equation}
142    
143     Conservation of linear momentum in each of the three directions
144     ($\alpha = x,y,z$) ties the values of the hot and cold bin scaling
145     parameters together:
146     \begin{equation}
147 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
148 gezelter 3524 \end{equation}
149     where
150 skuang 3565 \begin{eqnarray}
151 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
152 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
153 gezelter 3524 \label{eq:momentumdef}
154 skuang 3565 \end{eqnarray}
155 skuang 3528 Therefore, for each of the three directions, the hot scaling
156     parameters are a simple function of the cold scaling parameters and
157 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
158     \begin{equation}
159 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
160 gezelter 3524 \label{eq:hotcoldscaling}
161     \end{equation}
162 skuang 3528 where
163     \begin{equation}
164     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
165     \end{equation}
166     for convenience.
167 gezelter 3524
168     Conservation of total energy also places constraints on the scaling:
169     \begin{equation}
170     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
171 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
172 gezelter 3524 \end{equation}
173     where the kinetic energies, $K_h^\alpha$ and $K_c^\alpha$, are computed
174     for each of the three directions in a similar manner to the linear momenta
175     (Eq. \ref{eq:momentumdef}). Substituting in the expressions for the
176 skuang 3528 hot scaling parameters ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}),
177     we obtain the {\it constraint ellipsoid equation}:
178 gezelter 3524 \begin{equation}
179 skuang 3565 \sum_{\alpha = x,y,z} a_\alpha \alpha^2 + b_\alpha \alpha + c_\alpha = 0
180 gezelter 3524 \label{eq:constraintEllipsoid}
181     \end{equation}
182     where the constants are obtained from the instantaneous values of the
183     linear momenta and kinetic energies for the hot and cold slabs,
184 skuang 3565 \begin{eqnarray}
185 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
186     \left(p_\alpha\right)^2\right) \\
187     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
188 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
189 gezelter 3524 \label{eq:constraintEllipsoidConsts}
190 skuang 3565 \end{eqnarray}
191 skuang 3528 This ellipsoid equation defines the set of cold slab scaling
192     parameters which can be applied while preserving both linear momentum
193 skuang 3530 in all three directions as well as kinetic energy.
194 gezelter 3524
195     The goal of using velocity scaling variables is to transfer linear
196     momentum or kinetic energy from the cold slab to the hot slab. If the
197     hot and cold slabs are separated along the z-axis, the energy flux is
198 skuang 3528 given simply by the decrease in kinetic energy of the cold bin:
199 gezelter 3524 \begin{equation}
200 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
201 gezelter 3524 \end{equation}
202     The expression for the energy flux can be re-written as another
203     ellipsoid centered on $(x,y,z) = 0$:
204     \begin{equation}
205 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
206 gezelter 3524 \label{eq:fluxEllipsoid}
207     \end{equation}
208 skuang 3529 The spatial extent of the {\it flux ellipsoid equation} is governed
209     both by a targetted value, $J_z$ as well as the instantaneous values of the
210 skuang 3530 kinetic energy components in the cold bin.
211 gezelter 3524
212     To satisfy an energetic flux as well as the conservation constraints,
213     it is sufficient to determine the points ${x,y,z}$ which lie on both
214     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
215     flux ellipsoid (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of
216 skuang 3528 the two ellipsoids in 3-dimensional space.
217 gezelter 3524
218 gezelter 3569 \begin{figure}
219     \includegraphics[width=\linewidth]{ellipsoids}
220     \caption{Scaling points which maintain both constant energy and
221     constant linear momentum of the system lie on the surface of the
222     {\it constraint ellipsoid} while points which generate the target
223     momentum flux lie on the surface of the {\it flux ellipsoid}. The
224     velocity distributions in the hot bin are scaled by only those
225     points which lie on both ellipsoids.}
226     \label{ellipsoids}
227     \end{figure}
228    
229 gezelter 3524 One may also define momentum flux (say along the x-direction) as:
230     \begin{equation}
231 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
232 skuang 3531 \label{eq:fluxPlane}
233 gezelter 3524 \end{equation}
234 skuang 3531 The above {\it flux equation} is essentially a plane which is
235     perpendicular to the x-axis, with its position governed both by a
236     targetted value, $j_z(p_x)$ as well as the instantaneous value of the
237     momentum along the x-direction.
238 gezelter 3524
239 skuang 3531 Similarly, to satisfy a momentum flux as well as the conservation
240     constraints, it is sufficient to determine the points ${x,y,z}$ which
241     lie on both the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid})
242     and the flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of
243     an ellipsoid and a plane in 3-dimensional space.
244 gezelter 3524
245 skuang 3531 To summarize, by solving respective equation sets, one can determine
246     possible sets of scaling variables for cold slab. And corresponding
247     sets of scaling variables for hot slab can be determine as well.
248 gezelter 3524
249 skuang 3531 The following problem will be choosing an optimal set of scaling
250     variables among the possible sets. Although this method is inherently
251     non-isotropic, the goal is still to maintain the system as isotropic
252     as possible. Under this consideration, one would like the kinetic
253     energies in different directions could become as close as each other
254     after each scaling. Simultaneously, one would also like each scaling
255     as gentle as possible, i.e. ${x,y,z \rightarrow 1}$, in order to avoid
256     large perturbation to the system. Therefore, one approach to obtain the
257     scaling variables would be constructing an criteria function, with
258     constraints as above equation sets, and solving the function's minimum
259     by method like Lagrange multipliers.
260 gezelter 3524
261 skuang 3531 In order to save computation time, we have a different approach to a
262     relatively good set of scaling variables with much less calculation
263     than above. Here is the detail of our simplification of the problem.
264 gezelter 3524
265 skuang 3531 In the case of kinetic energy transfer, we impose another constraint
266     ${x = y}$, into the equation sets. Consequently, there are two
267     variables left. And now one only needs to solve a set of two {\it
268     ellipses equations}. This problem would be transformed into solving
269     one quartic equation for one of the two variables. There are known
270     generic methods that solve real roots of quartic equations. Then one
271     can determine the other variable and obtain sets of scaling
272     variables. Among these sets, one can apply the above criteria to
273     choose the best set, while much faster with only a few sets to choose.
274    
275     In the case of momentum flux transfer, we impose another constraint to
276     set the kinetic energy transfer as zero. In another word, we apply
277     Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. After that, with one
278     variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar set
279     of equations on the above kinetic energy transfer problem. Therefore,
280     an approach similar to the above would be sufficient for this as well.
281    
282     \section{Computational Details}
283 skuang 3534 Our simulation consists of a series of systems. All of these
284 skuang 3565 simulations were run with the OpenMD simulation software
285 skuang 3534 package\cite{Meineke:2005gd} integrated with RNEMD methods.
286 skuang 3531
287 skuang 3532 A Lennard-Jones fluid system was built and tested first. In order to
288     compare our method with swapping RNEMD, a series of simulations were
289     performed to calculate the shear viscosity and thermal conductivity of
290 skuang 3534 argon. 2592 atoms were in a orthorhombic cell, which was ${10.06\sigma
291     \times 10.06\sigma \times 30.18\sigma}$ by size. The reduced density
292     ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled direct
293     comparison between our results and others. These simulations used
294 skuang 3565 velocity Verlet algorithm with reduced timestep ${\tau^* =
295 skuang 3534 4.6\times10^{-4}}$.
296 skuang 3532
297     For shear viscosity calculation, the reduced temperature was ${T^* =
298 skuang 3565 k_B T/\varepsilon = 0.72}$. Simulations were first equilibrated in canonical
299     ensemble (NVT), then equilibrated in microcanonical ensemble
300     (NVE). Establishing and stablizing momentum gradient were followed
301     also in NVE ensemble. For the swapping part, Muller-Plathe's algorithm was
302 skuang 3532 adopted.\cite{ISI:000080382700030} The simulation box was under
303 skuang 3534 periodic boundary condition, and devided into ${N = 20}$ slabs. In each swap,
304     the top slab ${(n = 1)}$ exchange the most negative $x$ momentum with the
305     most positive $x$ momentum in the center slab ${(n = N/2 + 1)}$. Referred
306 skuang 3565 to Tenney {\it et al.}\cite{ISI:000273472300004}, a series of swapping
307 skuang 3534 frequency were chosen. According to each result from swapping
308 skuang 3532 RNEMD, scaling RNEMD simulations were run with the target momentum
309 skuang 3534 flux set to produce a similar momentum flux and shear
310     rate. Furthermore, various scaling frequencies can be tested for one
311     single swapping rate. To compare the performance between swapping and
312     scaling method, temperatures of different dimensions in all the slabs
313 skuang 3538 were observed. Most of the simulations include $10^5$ steps of
314     equilibration without imposing momentum flux, $10^5$ steps of
315     stablization with imposing momentum transfer, and $10^6$ steps of data
316     collection under RNEMD. For relatively high momentum flux simulations,
317     ${5\times10^5}$ step data collection is sufficient. For some low momentum
318     flux simulations, ${2\times10^6}$ steps were necessary.
319 skuang 3532
320 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
321     unit. The momentum flux was calculated with total unphysical
322 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
323 skuang 3534 \begin{equation}
324     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
325     \end{equation}
326     And the velocity gradient ${\langle \partial v_x /\partial z \rangle}$
327     can be obtained by a linear regression of the velocity profile. From
328     the shear viscosity $\eta$ calculated with the above parameters, one
329     can further convert it into reduced unit ${\eta^* = \eta \sigma^2
330     (\varepsilon m)^{-1/2}}$.
331 skuang 3532
332 skuang 3534 For thermal conductivity calculation, simulations were first run under
333     reduced temperature ${T^* = 0.72}$ in NVE ensemble. Muller-Plathe's
334     algorithm was adopted in the swapping method. Under identical
335 skuang 3536 simulation box parameters, in each swap, the top slab exchange the
336     molecule with least kinetic energy with the molecule in the center
337     slab with most kinetic energy, unless this ``coldest'' molecule in the
338     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the ``cold''
339 skuang 3534 slab. According to swapping RNEMD results, target energy flux for
340     scaling RNEMD simulations can be set. Also, various scaling
341     frequencies can be tested for one target energy flux. To compare the
342     performance between swapping and scaling method, distributions of
343     velocity and speed in different slabs were observed.
344    
345     For each swapping rate, thermal conductivity was calculated in reduced
346     unit. The energy flux was calculated similarly to the momentum flux,
347 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
348 skuang 3534 time $t$:
349     \begin{equation}
350     J_z = \frac{E_{total}}{2 t L_x L_y}
351     \end{equation}
352     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
353     can be obtained by a linear regression of the temperature
354     profile. From the thermal conductivity $\lambda$ calculated, one can
355     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
356     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
357    
358 skuang 3565 Another series of our simulation is to calculate the interfacial
359 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
360 skuang 3565 water (SPC/E) and gold (QSC) thermal conductivity were performed and
361 skuang 3563 compared with current results to ensure the validity of
362 skuang 3573 NIVS-RNEMD. After that, a mixture system was simulated.
363 skuang 3563
364 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
365     consisting of 1000 molecules were first equilibrated under ambient
366     pressure and temperature conditions (NPT), followed by equilibration
367     in fixed volume (NVT). The system was then equilibrated in
368     microcanonical ensemble (NVE). Also in NVE ensemble, establishing
369     stable thermal gradient was followed. The simulation box was under
370     periodic boundary condition and devided into 10 slabs. Data collection
371     process was similar to Lennard-Jones fluid system. Thermal
372     conductivity calculation of bulk crystal gold used a similar
373     protocol. And the face centered cubic crystal simulation box consists
374     of 2880 Au atoms.
375    
376     After simulations of bulk water and crystal gold, a mixture system was
377     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
378     molecules. Spohr potential was adopted in depicting the interaction
379     between metal atom and water molecule.\cite{ISI:000167766600035} A
380     similar protocol of equilibration was followed. A thermal gradient was
381     built. It was found out that compared to homogeneous systems, the two
382     phases could have large temperature difference under a relatively low
383     thermal flux. Therefore, under our low flux condition, it is assumed
384     that the metal and water phases have respectively homogeneous
385     temperature. In calculating the interfacial thermal conductivity $G$,
386     this assumptioin was applied and thus our formula becomes:
387    
388     \begin{equation}
389     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
390     \langle T_{water}\rangle \right)}
391     \label{interfaceCalc}
392     \end{equation}
393     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
394     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
395     average observed temperature of gold and water phases respectively.
396    
397 skuang 3534 \section{Results And Discussion}
398     \subsection{Shear Viscosity}
399 skuang 3538 Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
400     produced comparable shear viscosity to swap RNEMD method. In Table
401     \ref{shearRate}, the names of the calculated samples are devided into
402     two parts. The first number refers to total slabs in one simulation
403     box. The second number refers to the swapping interval in swap method, or
404     in scale method the equilvalent swapping interval that the same
405     momentum flux would theoretically result in swap method. All the scale
406 skuang 3563 method results were from simulations that had a scaling interval of 10
407     time steps. The average molecular momentum gradients of these samples
408 skuang 3565 are shown in Figure \ref{shearGrad}.
409 skuang 3534
410 skuang 3538 \begin{table*}
411     \begin{minipage}{\linewidth}
412     \begin{center}
413    
414     \caption{Calculation results for shear viscosity of Lennard-Jones
415     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
416     methods at various momentum exchange rates. Results in reduced
417     unit. Errors of calculations in parentheses. }
418    
419 skuang 3565 \begin{tabular}{ccc}
420 skuang 3538 \hline
421 skuang 3565 Series & $\eta^*_{swap}$ & $\eta^*_{scale}$\\
422 skuang 3538 \hline
423     20-500 & 3.64(0.05) & 3.76(0.09)\\
424 skuang 3539 20-1000 & 3.52(0.16) & 3.66(0.06)\\
425     20-2000 & 3.72(0.05) & 3.32(0.18)\\
426 skuang 3566 20-2500 & 3.42(0.06) & 3.43(0.08)\\
427     \hline
428 skuang 3538 \end{tabular}
429     \label{shearRate}
430     \end{center}
431     \end{minipage}
432     \end{table*}
433    
434     \begin{figure}
435 skuang 3565 \includegraphics[width=\linewidth]{shearGrad}
436     \caption{Average momentum gradients of shear viscosity simulations}
437     \label{shearGrad}
438 skuang 3538 \end{figure}
439    
440     \begin{figure}
441 skuang 3565 \includegraphics[width=\linewidth]{shearTempScale}
442 skuang 3538 \caption{Temperature profile for scaling RNEMD simulation.}
443     \label{shearTempScale}
444     \end{figure}
445     However, observations of temperatures along three dimensions show that
446     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
447     two slabs which were scaled. Figure \ref{shearTempScale} indicate that with
448 skuang 3563 relatively large imposed momentum flux, the temperature difference among $x$
449     and the other two dimensions was significant. This would result from the
450     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
451     momentum gradient is set up, $P_c^x$ would be roughly stable
452     ($<0$). Consequently, scaling factor $x$ would most probably larger
453     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
454     keep increase after most scaling steps. And if there is not enough time
455     for the kinetic energy to exchange among different dimensions and
456     different slabs, the system would finally build up temperature
457     (kinetic energy) difference among the three dimensions. Also, between
458     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
459     are closer to neighbor slabs. This is due to momentum transfer along
460     $z$ dimension between slabs.
461 skuang 3538
462     Although results between scaling and swapping methods are comparable,
463 skuang 3563 the inherent temperature inhomogeneity even in relatively low imposed
464     exchange momentum flux simulations makes scaling RNEMD method less
465 skuang 3538 attractive than swapping RNEMD in shear viscosity calculation.
466    
467     \subsection{Thermal Conductivity}
468 skuang 3573 \subsubsection{Lennard-Jones Fluid}
469 skuang 3567 Our thermal conductivity calculations also show that scaling method results
470     agree with swapping method. Table \ref{thermal} lists our simulation
471 skuang 3563 results with similar manner we used in shear viscosity
472     calculation. All the data reported from scaling method were obtained
473     by simulations of 10-step exchange frequency, and the target exchange
474     kinetic energy were set to produce equivalent kinetic energy flux as
475 skuang 3567 in swapping method. Figure \ref{thermalGrad} exhibits similar thermal
476     gradients of respective similar kinetic energy flux.
477 skuang 3538
478 skuang 3563 \begin{table*}
479     \begin{minipage}{\linewidth}
480     \begin{center}
481 skuang 3538
482 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
483 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
484 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
485     in reduced unit. Errors of calculations in parentheses.}
486    
487 skuang 3565 \begin{tabular}{ccc}
488 skuang 3563 \hline
489 skuang 3567 Series & $\lambda^*_{swap}$ & $\lambda^*_{scale}$\\
490 skuang 3565 \hline
491 skuang 3564 20-250 & 7.03(0.34) & 7.30(0.10)\\
492     20-500 & 7.03(0.14) & 6.95(0.09)\\
493 skuang 3563 20-1000 & 6.91(0.42) & 7.19(0.07)\\
494 skuang 3566 20-2000 & 7.52(0.15) & 7.19(0.28)\\
495     \hline
496 skuang 3563 \end{tabular}
497     \label{thermal}
498     \end{center}
499     \end{minipage}
500     \end{table*}
501    
502     \begin{figure}
503 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
504     \caption{Temperature gradients of thermal conductivity simulations}
505     \label{thermalGrad}
506 skuang 3563 \end{figure}
507    
508     During these simulations, molecule velocities were recorded in 1000 of
509     all the snapshots. These velocity data were used to produce histograms
510     of velocity and speed distribution in different slabs. From these
511     histograms, it is observed that with increasing unphysical kinetic
512     energy flux, speed and velocity distribution of molecules in slabs
513     where swapping occured could deviate from Maxwell-Boltzmann
514     distribution. Figure \ref{histSwap} indicates how these distributions
515     deviate from ideal condition. In high temperature slabs, probability
516     density in low speed is confidently smaller than ideal distribution;
517     in low temperature slabs, probability density in high speed is smaller
518     than ideal. This phenomenon is observable even in our relatively low
519 skuang 3568 swapping rate simulations. And this deviation could also leads to
520 skuang 3563 deviation of distribution of velocity in various dimensions. One
521     feature of these deviated distribution is that in high temperature
522     slab, the ideal Gaussian peak was changed into a relatively flat
523     plateau; while in low temperature slab, that peak appears sharper.
524    
525     \begin{figure}
526 skuang 3565 \includegraphics[width=\linewidth]{histSwap}
527 skuang 3563 \caption{Speed distribution for thermal conductivity using swapping RNEMD.}
528     \label{histSwap}
529     \end{figure}
530    
531 skuang 3568 \begin{figure}
532     \includegraphics[width=\linewidth]{histScale}
533     \caption{Speed distribution for thermal conductivity using scaling RNEMD.}
534     \label{histScale}
535     \end{figure}
536    
537 skuang 3573 \subsubsection{SPC/E Water}
538     Our results of SPC/E water thermal conductivity are comparable to
539     Bedrov {\it et al.}\cite{ISI:000090151400044}, which employed the
540     previous swapping RNEMD method for their calculation. Our simulations
541     were able to produce a similar temperature gradient to their
542     system. However, the average temperature of our system is 300K, while
543     theirs is 318K, which would be attributed for part of the difference
544     between the two series of results. Both methods yields values in
545     agreement with experiment. And this shows the applicability of our
546     method to multi-atom molecular system.
547 skuang 3563
548 skuang 3570 \begin{figure}
549     \includegraphics[width=\linewidth]{spceGrad}
550     \caption{Temperature gradients for SPC/E water thermal conductivity.}
551     \label{spceGrad}
552     \end{figure}
553    
554     \begin{table*}
555     \begin{minipage}{\linewidth}
556     \begin{center}
557    
558     \caption{Calculation results for thermal conductivity of SPC/E water
559     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
560     calculations in parentheses. }
561    
562     \begin{tabular}{cccc}
563     \hline
564     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
565 skuang 3573 & This work & Previous simulations\cite{ISI:000090151400044} &
566     Experiment$^a$\\
567 skuang 3570 \hline
568 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
569     0.81 & 0.770(0.008) & 0.784\\
570     1.54 & 0.813(0.007) & 0.730\\
571 skuang 3570 \hline
572     \end{tabular}
573     \label{spceThermal}
574     \end{center}
575     \end{minipage}
576     \end{table*}
577    
578 skuang 3573 \subsubsection{Crystal Gold}
579 skuang 3570
580     \begin{figure}
581     \includegraphics[width=\linewidth]{AuGrad}
582     \caption{Temperature gradients for crystal gold thermal conductivity.}
583     \label{AuGrad}
584     \end{figure}
585    
586     \begin{table*}
587     \begin{minipage}{\linewidth}
588     \begin{center}
589    
590     \caption{Calculation results for thermal conductivity of crystal gold
591     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
592     calculations in parentheses. }
593    
594     \begin{tabular}{ccc}
595     \hline
596     $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
597 skuang 3573 & This work & Previous simulations\cite{ISI:A1992HX37800010} \\
598 skuang 3570 \hline
599 skuang 3571 1.44 & 1.10(0.01) & \\
600     2.86 & 1.08(0.02) & \\
601     5.14 & 1.15(0.01) & \\
602 skuang 3570 \hline
603     \end{tabular}
604     \label{AuThermal}
605     \end{center}
606     \end{minipage}
607     \end{table*}
608    
609    
610 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
611    
612    
613 skuang 3571 \begin{figure}
614     \includegraphics[width=\linewidth]{interfaceDensity}
615     \caption{Density profile for interfacial thermal conductivity
616     simulation box.}
617     \label{interfaceDensity}
618     \end{figure}
619    
620    
621 skuang 3572 \begin{figure}
622     \includegraphics[width=\linewidth]{interfaceGrad}
623     \caption{Temperature profiles for interfacial thermal conductivity
624     simulation box.}
625     \label{interfaceGrad}
626     \end{figure}
627    
628    
629    
630     \begin{table*}
631     \begin{minipage}{\linewidth}
632     \begin{center}
633    
634     \caption{Calculation results for interfacial thermal conductivity
635     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
636     rates. Errors of calculations in parentheses. }
637    
638     \begin{tabular}{cccc}
639     \hline
640     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
641     \hline
642 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
643     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
644     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
645     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
646 skuang 3572 \hline
647     \end{tabular}
648     \label{AuThermal}
649     \end{center}
650     \end{minipage}
651     \end{table*}
652    
653    
654 gezelter 3524 \section{Acknowledgments}
655     Support for this project was provided by the National Science
656     Foundation under grant CHE-0848243. Computational time was provided by
657     the Center for Research Computing (CRC) at the University of Notre
658     Dame. \newpage
659    
660     \bibliographystyle{jcp2}
661     \bibliography{nivsRnemd}
662     \end{doublespace}
663     \end{document}
664