ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3600
Committed: Thu Apr 22 20:58:46 2010 UTC (14 years, 4 months ago) by gezelter
Content type: application/x-tex
File size: 38009 byte(s)
Log Message:
more changes

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 gezelter 3583 We present a new method for introducing stable non-equilibrium
42     velocity and temperature distributions in molecular dynamics
43     simulations of heterogeneous systems. This method extends some
44     earlier Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods
45     which use momentum exchange swapping moves that can create
46     non-thermal velocity distributions (and which are difficult to use
47     for interfacial calculations). By using non-isotropic velocity
48     scaling (NIVS) on the molecules in specific regions of a system, it
49     is possible to impose momentum or thermal flux between regions of a
50     simulation and stable thermal and momentum gradients can then be
51     established. The scaling method we have developed conserves the
52     total linear momentum and total energy of the system. To test the
53     methods, we have computed the thermal conductivity of model liquid
54     and solid systems as well as the interfacial thermal conductivity of
55     a metal-water interface. We find that the NIVS-RNEMD improves the
56     problematic velocity distributions that develop in other RNEMD
57     methods.
58 gezelter 3524 \end{abstract}
59    
60     \newpage
61    
62     %\narrowtext
63    
64     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65     % BODY OF TEXT
66     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67    
68     \section{Introduction}
69     The original formulation of Reverse Non-equilibrium Molecular Dynamics
70     (RNEMD) obtains transport coefficients (thermal conductivity and shear
71     viscosity) in a fluid by imposing an artificial momentum flux between
72     two thin parallel slabs of material that are spatially separated in
73 skuang 3534 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
74 gezelter 3583 artificial flux is typically created by periodically ``swapping''
75     either the entire momentum vector $\vec{p}$ or single components of
76     this vector ($p_x$) between molecules in each of the two slabs. If
77     the two slabs are separated along the $z$ coordinate, the imposed flux
78     is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
79     response of a simulated system to the imposed momentum flux will
80     typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
81     The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
82     easily obtained by assuming linear response of the system,
83 gezelter 3524 \begin{eqnarray}
84 skuang 3532 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
85 skuang 3575 J_z & = & \lambda \frac{\partial T}{\partial z}
86 gezelter 3524 \end{eqnarray}
87 gezelter 3600 RNEMD has been widely used to provide computational estimates of
88     thermal conductivities and shear viscosities in a wide range of
89     materials, from liquid copper to both monatomic and molecular fluids
90     (e.g. ionic
91     liquids).\cite{Bedrov:2000-1,Bedrov:2000,Muller-Plathe:2002,ISI:000184808400018,ISI:000231042800044,Maginn:2007,Muller-Plathe:2008,ISI:000258460400020,ISI:000258840700015,ISI:000261835100054}
92 gezelter 3524
93 skuang 3574 \begin{figure}
94     \includegraphics[width=\linewidth]{thermalDemo}
95 gezelter 3583 \caption{RNEMD methods impose an unphysical transfer of momentum or
96     kinetic energy between a ``hot'' slab and a ``cold'' slab in the
97     simulation box. The molecular system responds to this imposed flux
98     by generating a momentum or temperature gradient. The slope of the
99     gradient can then be used to compute transport properties (e.g.
100     shear viscosity and thermal conductivity).}
101 skuang 3574 \label{thermalDemo}
102     \end{figure}
103    
104 skuang 3591 RNEMD is preferable in many ways to the forward NEMD
105 skuang 3592 methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
106     because it imposes what is typically difficult to measure (a flux or
107 gezelter 3600 stress) and it is typically much easier to compute the response
108     (momentum gradients or strains. For similar reasons, RNEMD is also
109 skuang 3592 preferable to slowly-converging equilibrium methods for measuring
110     thermal conductivity and shear viscosity (using Green-Kubo
111 skuang 3591 relations\cite{daivis:541,mondello:9327} or the Helfand moment
112     approach of Viscardy {\it et
113 skuang 3527 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
114 gezelter 3524 computing difficult to measure quantities.
115    
116     Another attractive feature of RNEMD is that it conserves both total
117     linear momentum and total energy during the swaps (as long as the two
118     molecules have the same identity), so the swapped configurations are
119     typically samples from the same manifold of states in the
120     microcanonical ensemble.
121    
122 skuang 3588 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered
123 skuang 3565 some problems with the original RNEMD swap technique. Notably, large
124     momentum fluxes (equivalent to frequent momentum swaps between the
125 skuang 3575 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
126     momentum distributions in the two slabs, as well as non-linear thermal
127     and velocity distributions along the direction of the imposed flux
128     ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
129     and self-adjusting metrics for retaining the usability of the method.
130 gezelter 3524
131     In this paper, we develop and test a method for non-isotropic velocity
132 gezelter 3600 scaling (NIVS) which retains the desirable features of RNEMD
133 gezelter 3524 (conservation of linear momentum and total energy, compatibility with
134     periodic boundary conditions) while establishing true thermal
135 gezelter 3600 distributions in each of the two slabs. In the next section, we
136 gezelter 3583 present the method for determining the scaling constraints. We then
137 gezelter 3600 test the method on both liquids and solids as well as a non-isotropic
138     liquid-solid interface and show that it is capable of providing
139 gezelter 3524 reasonable estimates of the thermal conductivity and shear viscosity
140 gezelter 3600 in all of these cases.
141 gezelter 3524
142     \section{Methodology}
143 gezelter 3583 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
144     periodic system is partitioned into a series of thin slabs along one
145 gezelter 3524 axis ($z$). One of the slabs at the end of the periodic box is
146     designated the ``hot'' slab, while the slab in the center of the box
147     is designated the ``cold'' slab. The artificial momentum flux will be
148     established by transferring momentum from the cold slab and into the
149     hot slab.
150    
151     Rather than using momentum swaps, we use a series of velocity scaling
152 gezelter 3583 moves. For molecules $\{i\}$ located within the cold slab,
153 gezelter 3524 \begin{equation}
154 skuang 3565 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
155     x & 0 & 0 \\
156     0 & y & 0 \\
157     0 & 0 & z \\
158 gezelter 3524 \end{array} \right) \cdot \vec{v}_i
159     \end{equation}
160 gezelter 3600 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
161     of the three directions in the system. Likewise, the molecules
162     $\{j\}$ located in the hot slab will see a concomitant scaling of
163     velocities,
164 gezelter 3524 \begin{equation}
165 skuang 3565 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
166     x^\prime & 0 & 0 \\
167     0 & y^\prime & 0 \\
168     0 & 0 & z^\prime \\
169 gezelter 3524 \end{array} \right) \cdot \vec{v}_j
170     \end{equation}
171    
172     Conservation of linear momentum in each of the three directions
173 gezelter 3583 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
174 gezelter 3524 parameters together:
175     \begin{equation}
176 skuang 3528 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
177 gezelter 3524 \end{equation}
178     where
179 skuang 3565 \begin{eqnarray}
180 skuang 3528 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i \left[\vec{v}_i\right]_\alpha \\
181 skuang 3565 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j \left[\vec{v}_j\right]_\alpha
182 gezelter 3524 \label{eq:momentumdef}
183 skuang 3565 \end{eqnarray}
184 skuang 3528 Therefore, for each of the three directions, the hot scaling
185     parameters are a simple function of the cold scaling parameters and
186 gezelter 3524 the instantaneous linear momentum in each of the two slabs.
187     \begin{equation}
188 skuang 3528 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
189 gezelter 3524 \label{eq:hotcoldscaling}
190     \end{equation}
191 skuang 3528 where
192     \begin{equation}
193     p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
194     \end{equation}
195     for convenience.
196 gezelter 3524
197     Conservation of total energy also places constraints on the scaling:
198     \begin{equation}
199     \sum_{\alpha = x,y,z} K_h^\alpha + K_c^\alpha = \sum_{\alpha = x,y,z}
200 skuang 3565 \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha
201 gezelter 3524 \end{equation}
202 skuang 3575 where the translational kinetic energies, $K_h^\alpha$ and
203     $K_c^\alpha$, are computed for each of the three directions in a
204     similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
205     Substituting in the expressions for the hot scaling parameters
206     ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
207 gezelter 3583 {\it constraint ellipsoid}:
208 gezelter 3524 \begin{equation}
209 gezelter 3600 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
210     c_\alpha \right) = 0
211 gezelter 3524 \label{eq:constraintEllipsoid}
212     \end{equation}
213     where the constants are obtained from the instantaneous values of the
214     linear momenta and kinetic energies for the hot and cold slabs,
215 skuang 3565 \begin{eqnarray}
216 skuang 3528 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
217     \left(p_\alpha\right)^2\right) \\
218     b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
219 skuang 3565 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
220 gezelter 3524 \label{eq:constraintEllipsoidConsts}
221 skuang 3565 \end{eqnarray}
222 gezelter 3583 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
223 gezelter 3600 cold slab scaling parameters which, when applied, preserve the linear
224     momentum of the system in all three directions as well as total
225     kinetic energy.
226 gezelter 3524
227 gezelter 3600 The goal of using these velocity scaling variables is to transfer
228     linear momentum or kinetic energy from the cold slab to the hot slab.
229     If the hot and cold slabs are separated along the z-axis, the energy
230     flux is given simply by the decrease in kinetic energy of the cold
231     bin:
232 gezelter 3524 \begin{equation}
233 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
234 gezelter 3524 \end{equation}
235     The expression for the energy flux can be re-written as another
236     ellipsoid centered on $(x,y,z) = 0$:
237     \begin{equation}
238 gezelter 3600 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = -J_z \Delta t +
239     \sum_{\alpha = x,y,z} K_c^\alpha
240 gezelter 3524 \label{eq:fluxEllipsoid}
241     \end{equation}
242 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
243 gezelter 3600 both by the target flux, $J_z$ as well as the instantaneous values of
244     the kinetic energy components in the cold bin.
245 gezelter 3524
246     To satisfy an energetic flux as well as the conservation constraints,
247 gezelter 3600 we must determine the points ${x,y,z}$ that lie on both the constraint
248     ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
249     (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
250     ellipsoids in 3-dimensional space.
251 gezelter 3524
252 gezelter 3569 \begin{figure}
253     \includegraphics[width=\linewidth]{ellipsoids}
254 gezelter 3600 \caption{Velocity scaling coefficients which maintain both constant
255     energy and constant linear momentum of the system lie on the surface
256     of the {\it constraint ellipsoid} while points which generate the
257     target momentum flux lie on the surface of the {\it flux ellipsoid}.
258     The velocity distributions in the cold bin are scaled by only those
259     points which lie on both ellipsoids.}
260 gezelter 3569 \label{ellipsoids}
261     \end{figure}
262    
263 gezelter 3600 Since ellipsoids can be expressed as polynomials up to second order in
264     each of the three coordinates, finding the the intersection points of
265     two ellipsoids is isomorphic to finding the roots a polynomial of
266     degree 16. There are a number of polynomial root-finding methods in
267     the literature, [CITATIONS NEEDED] but numerically finding the roots
268     of high-degree polynomials is generally an ill-conditioned
269     problem.[CITATION NEEDED] One way around this is to try to maintain
270     velocity scalings that are {\it as isotropic as possible}. To do
271     this, we impose $x=y$, and to treat both the constraint and flux
272     ellipsoids as 2-dimensional ellipses. In reduced dimensionality, the
273     intersecting-ellipse problem reduces to finding the roots of
274     polynomials of degree 4.
275    
276     Depending on the target flux and current velocity distributions, the
277     ellipsoids can have between 0 and 4 intersection points. If there are
278     no intersection points, it is not possible to satisfy the constraints
279     while performing a non-equilibrium scaling move, and no change is made
280     to the dynamics.
281    
282     With multiple intersection points, any of the scaling points will
283     conserve the linear momentum and kinetic energy of the system and will
284     generate the correct target flux. Although this method is inherently
285     non-isotropic, the goal is still to maintain the system as close to an
286     isotropic fluid as possible. With this in mind, we would like the
287     kinetic energies in the three different directions could become as
288     close as each other as possible after each scaling. Simultaneously,
289     one would also like each scaling as gentle as possible, i.e. ${x,y,z
290     \rightarrow 1}$, in order to avoid large perturbation to the system.
291     To do this, we pick the intersection point which maintains the scaling
292     variables ${x=y, z}$ as well as the ratio of kinetic energies
293     ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
294    
295     After the valid scaling parameters are arrived at by solving geometric
296     intersection problems in $x, y, z$ space in order to obtain cold slab
297     scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
298     determine the conjugate hot slab scaling variables.
299    
300     \subsection{Introducing shear stress via velocity scaling}
301     Rather than using this method to induce a thermal flux, it is possible
302     to use the random fluctuations of the average momentum in each of the
303     bins to induce a momentum flux. Doing this repeatedly will create a
304     shear stress on the system which will respond with an easily-measured
305     strain. The momentum flux (say along the $x$-direction) may be
306     defined as:
307 gezelter 3524 \begin{equation}
308 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
309 skuang 3531 \label{eq:fluxPlane}
310 gezelter 3524 \end{equation}
311 gezelter 3600 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
312     its position governed both by a target value, $j_z(p_x)$ as well as
313     the instantaneous value of the momentum along the $x$-direction.
314 gezelter 3524
315 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
316     constraints, we must determine the points ${x,y,z}$ which lie on both
317     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
318     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
319 gezelter 3600 ellipsoid and a plane in 3-dimensional space.
320 gezelter 3524
321 gezelter 3600 In the case of momentum flux transfer, we also impose another
322     constraint to set the kinetic energy transfer as zero. In another
323     word, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
324     one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
325     set of quartic equations to the above kinetic energy transfer problem.
326 gezelter 3524
327 gezelter 3600 \section{Computational Details}
328 gezelter 3583
329 gezelter 3600 We have implemented this methodology in our molecular dynamics
330     code,\cite{Meineke:2005gd} by performing the NIVS scaling moves after
331     each MD step. We have tested it for a variety of different
332     situations, including homogeneous fluids (Lennard-Jones and SPC/E
333     water), crystalline solids (EAM and Sutton-Chen models for Gold), and
334     heterogeneous interfaces (EAM gold - SPC/E water). The last of these
335     systems would have been very difficult to study using previous RNEMD
336     methods, but using velocity scaling moves, we can even obtain
337     estimates of the interfacial thermal conductivity.
338 gezelter 3524
339 gezelter 3600 \subsection{Lennard-Jones Fluid}
340 gezelter 3524
341 gezelter 3600 2592 Lennard-Jones atoms were placed in an orthorhombic cell
342     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
343     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
344     direct comparison between our results and previous methods. These
345     simulations were carried out with a reduced timestep ${\tau^* =
346     4.6\times10^{-4}}$. For the shear viscosity calculation, the mean
347     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. Simulations were
348     first thermalized in canonical ensemble (NVT), then equilibrated in
349     microcanonical ensemble (NVE) before introducing any non-equilibrium
350     method.
351 skuang 3531
352 gezelter 3600 We have compared the momentum gradients established using our method
353     to those obtained using the original M\"{u}ller-Plathe swapping
354     algorithm.\cite{ISI:000080382700030} In both cases, the simulation box
355     was divided into ${N = 20}$ slabs. In the swapping algorithm, the top
356     slab $(n = 1)$ exchanges the most negative $x$ momentum with the most
357     positive $x$ momentum in the center slab $(n = N/2 + 1)$. The rate at
358     which the swapping moves are carried out defines the momentum or
359     thermal flux between the two slabs. In their work, Tenney {\it et
360     al.}\cite{Maginn:2010} found problematic behavior with large swap
361     frequencies.
362 skuang 3531
363 gezelter 3600 According to each result from swapping RNEMD, scaling RNEMD
364     simulations were run with the target momentum flux set to produce a
365     similar momentum flux, and consequently shear rate. Furthermore,
366     various scaling frequencies can be tested for one single swapping
367     rate. To test the temperature homogeneity in our system of swapping
368     and scaling methods, temperatures of different dimensions in all the
369     slabs were observed. Most of the simulations include $10^5$ steps of
370     equilibration without imposing momentum flux, $10^5$ steps of
371     stablization with imposing unphysical momentum transfer, and $10^6$
372     steps of data collection under RNEMD. For relatively high momentum
373     flux simulations, ${5\times10^5}$ step data collection is sufficient.
374     For some low momentum flux simulations, ${2\times10^6}$ steps were
375     necessary.
376 skuang 3531
377 skuang 3534 After each simulation, the shear viscosity was calculated in reduced
378     unit. The momentum flux was calculated with total unphysical
379 skuang 3565 transferred momentum ${P_x}$ and data collection time $t$:
380 skuang 3534 \begin{equation}
381     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
382     \end{equation}
383 skuang 3576 where $L_x$ and $L_y$ denotes $x$ and $y$ lengths of the simulation
384     box, and physical momentum transfer occurs in two ways due to our
385     periodic boundary condition settings. And the velocity gradient
386     ${\langle \partial v_x /\partial z \rangle}$ can be obtained by a
387     linear regression of the velocity profile. From the shear viscosity
388     $\eta$ calculated with the above parameters, one can further convert
389     it into reduced unit ${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$.
390 skuang 3532
391 skuang 3576 For thermal conductivity calculations, simulations were first run under
392     reduced temperature ${\langle T^*\rangle = 0.72}$ in NVE
393     ensemble. Muller-Plathe's algorithm was adopted in the swapping
394     method. Under identical simulation box parameters with our shear
395     viscosity calculations, in each swap, the top slab exchanges all three
396     translational momentum components of the molecule with least kinetic
397     energy with the same components of the molecule in the center slab
398     with most kinetic energy, unless this ``coldest'' molecule in the
399     ``hot'' slab is still ``hotter'' than the ``hottest'' molecule in the
400     ``cold'' slab. According to swapping RNEMD results, target energy flux
401     for scaling RNEMD simulations can be set. Also, various scaling
402 skuang 3534 frequencies can be tested for one target energy flux. To compare the
403     performance between swapping and scaling method, distributions of
404     velocity and speed in different slabs were observed.
405    
406     For each swapping rate, thermal conductivity was calculated in reduced
407     unit. The energy flux was calculated similarly to the momentum flux,
408 skuang 3565 with total unphysical transferred energy ${E_{total}}$ and data collection
409 skuang 3534 time $t$:
410     \begin{equation}
411     J_z = \frac{E_{total}}{2 t L_x L_y}
412     \end{equation}
413     And the temperature gradient ${\langle\partial T/\partial z\rangle}$
414     can be obtained by a linear regression of the temperature
415     profile. From the thermal conductivity $\lambda$ calculated, one can
416     further convert it into reduced unit ${\lambda^*=\lambda \sigma^2
417     m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$.
418    
419 skuang 3576 \subsection{ Water / Metal Thermal Conductivity}
420     Another series of our simulation is the calculation of interfacial
421 skuang 3573 thermal conductivity of a Au/H$_2$O system. Respective calculations of
422 skuang 3579 liquid water (Extended Simple Point Charge model) and crystal gold
423 skuang 3580 thermal conductivity were performed and compared with current results
424     to ensure the validity of NIVS-RNEMD. After that, a mixture system was
425     simulated.
426 skuang 3563
427 skuang 3573 For thermal conductivity calculation of bulk water, a simulation box
428     consisting of 1000 molecules were first equilibrated under ambient
429 skuang 3576 pressure and temperature conditions using NPT ensemble, followed by
430     equilibration in fixed volume (NVT). The system was then equilibrated in
431     microcanonical ensemble (NVE). Also in NVE ensemble, establishing a
432 skuang 3573 stable thermal gradient was followed. The simulation box was under
433     periodic boundary condition and devided into 10 slabs. Data collection
434 skuang 3576 process was similar to Lennard-Jones fluid system.
435 skuang 3573
436 skuang 3576 Thermal conductivity calculation of bulk crystal gold used a similar
437 skuang 3580 protocol. Two types of force field parameters, Embedded Atom Method
438     (EAM) and Quantum Sutten-Chen (QSC) force field were used
439     respectively. The face-centered cubic crystal simulation box consists of
440 skuang 3576 2880 Au atoms. The lattice was first allowed volume change to relax
441     under ambient temperature and pressure. Equilibrations in canonical and
442     microcanonical ensemble were followed in order. With the simulation
443     lattice devided evenly into 10 slabs, different thermal gradients were
444     established by applying a set of target thermal transfer flux. Data of
445     the series of thermal gradients was collected for calculation.
446    
447 skuang 3573 After simulations of bulk water and crystal gold, a mixture system was
448     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
449     molecules. Spohr potential was adopted in depicting the interaction
450     between metal atom and water molecule.\cite{ISI:000167766600035} A
451 skuang 3576 similar protocol of equilibration was followed. Several thermal
452     gradients was built under different target thermal flux. It was found
453     out that compared to our previous simulation systems, the two phases
454     could have large temperature difference even under a relatively low
455     thermal flux. Therefore, under our low flux conditions, it is assumed
456 skuang 3573 that the metal and water phases have respectively homogeneous
457 skuang 3576 temperature, excluding the surface regions. In calculating the
458     interfacial thermal conductivity $G$, this assumptioin was applied and
459     thus our formula becomes:
460 skuang 3573
461     \begin{equation}
462     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
463     \langle T_{water}\rangle \right)}
464     \label{interfaceCalc}
465     \end{equation}
466     where ${E_{total}}$ is the imposed unphysical kinetic energy transfer
467     and ${\langle T_{gold}\rangle}$ and ${\langle T_{water}\rangle}$ are the
468     average observed temperature of gold and water phases respectively.
469    
470 skuang 3577 \section{Results And Discussions}
471 skuang 3538 \subsection{Thermal Conductivity}
472 skuang 3573 \subsubsection{Lennard-Jones Fluid}
473 skuang 3577 Our thermal conductivity calculations show that scaling method results
474     agree with swapping method. Four different exchange intervals were
475     tested (Table \ref{thermalLJRes}) using swapping method. With a fixed
476     10fs exchange interval, target exchange kinetic energy was set to
477     produce equivalent kinetic energy flux as in swapping method. And
478     similar thermal gradients were observed with similar thermal flux in
479     two simulation methods (Figure \ref{thermalGrad}).
480 skuang 3538
481 skuang 3563 \begin{table*}
482     \begin{minipage}{\linewidth}
483     \begin{center}
484 skuang 3538
485 skuang 3563 \caption{Calculation results for thermal conductivity of Lennard-Jones
486 skuang 3565 fluid at ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$, with
487 skuang 3563 swap and scale methods at various kinetic energy exchange rates. Results
488     in reduced unit. Errors of calculations in parentheses.}
489    
490 skuang 3565 \begin{tabular}{ccc}
491 skuang 3563 \hline
492 skuang 3577 (Equilvalent) Exchange Interval (fs) & $\lambda^*_{swap}$ &
493     $\lambda^*_{scale}$\\
494 skuang 3565 \hline
495 skuang 3577 250 & 7.03(0.34) & 7.30(0.10)\\
496     500 & 7.03(0.14) & 6.95(0.09)\\
497     1000 & 6.91(0.42) & 7.19(0.07)\\
498     2000 & 7.52(0.15) & 7.19(0.28)\\
499 skuang 3566 \hline
500 skuang 3563 \end{tabular}
501 skuang 3577 \label{thermalLJRes}
502 skuang 3563 \end{center}
503     \end{minipage}
504     \end{table*}
505    
506     \begin{figure}
507 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
508 skuang 3589 \caption{NIVS-RNEMD method introduced similar temperature gradients
509     compared to ``swapping'' method under various kinetic energy flux in
510     thermal conductivity simulations.}
511 skuang 3567 \label{thermalGrad}
512 skuang 3563 \end{figure}
513    
514     During these simulations, molecule velocities were recorded in 1000 of
515 skuang 3578 all the snapshots of one single data collection process. These
516     velocity data were used to produce histograms of velocity and speed
517     distribution in different slabs. From these histograms, it is observed
518     that under relatively high unphysical kinetic energy flux, speed and
519     velocity distribution of molecules in slabs where swapping occured
520     could deviate from Maxwell-Boltzmann distribution. Figure
521 skuang 3589 \ref{thermalHist} a) illustrates how these distributions deviate from an
522 skuang 3578 ideal distribution. In high temperature slab, probability density in
523     low speed is confidently smaller than ideal curve fit; in low
524     temperature slab, probability density in high speed is smaller than
525     ideal, while larger than ideal in low speed. This phenomenon is more
526     obvious in our high swapping rate simulations. And this deviation
527     could also leads to deviation of distribution of velocity in various
528     dimensions. One feature of these deviated distribution is that in high
529     temperature slab, the ideal Gaussian peak was changed into a
530     relatively flat plateau; while in low temperature slab, that peak
531     appears sharper. This problem is rooted in the mechanism of the
532     swapping method. Continually depleting low (high) speed particles in
533     the high (low) temperature slab could not be complemented by
534     diffusions of low (high) speed particles from neighbor slabs, unless
535     in suffciently low swapping rate. Simutaneously, surplus low speed
536     particles in the low temperature slab do not have sufficient time to
537     diffuse to neighbor slabs. However, thermal exchange rate should reach
538     a minimum level to produce an observable thermal gradient under noise
539     interference. Consequently, swapping RNEMD has a relatively narrow
540     choice of swapping rate to satisfy these above restrictions.
541 skuang 3563
542 skuang 3578 Comparatively, NIVS-RNEMD has a speed distribution closer to the ideal
543 skuang 3589 curve fit (Figure \ref{thermalHist} b). Essentially, after scaling, a
544 skuang 3578 Gaussian distribution function would remain Gaussian. Although a
545     single scaling is non-isotropic in all three dimensions, our scaling
546     coefficient criteria could help maintian the scaling region as
547     isotropic as possible. On the other hand, scaling coefficients are
548     preferred to be as close to 1 as possible, which also helps minimize
549     the difference among different dimensions. This is possible if scaling
550     interval and one-time thermal transfer energy are well
551     chosen. Consequently, NIVS-RNEMD is able to impose an unphysical
552     thermal flux as the previous RNEMD method without large perturbation
553     to the distribution of velocity and speed in the exchange regions.
554    
555 skuang 3568 \begin{figure}
556 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
557     \caption{Speed distribution for thermal conductivity using a)
558     ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
559     simulations with an exchange or equilvalent exchange interval of 250
560 skuang 3593 fs. In circled areas, distributions from ``swapping'' RNEMD
561     simulation have deviation from ideal Maxwell-Boltzmann distribution
562     (curves fit for each distribution).}
563 skuang 3589 \label{thermalHist}
564 skuang 3568 \end{figure}
565    
566 skuang 3573 \subsubsection{SPC/E Water}
567     Our results of SPC/E water thermal conductivity are comparable to
568 gezelter 3594 Bedrov {\it et al.}\cite{Bedrov:2000}, which employed the
569 skuang 3579 previous swapping RNEMD method for their calculation. Bedrov {\it et
570 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
571 skuang 3579 center-of-mass velocities instead of single atom velocities in a
572     molecule conserves the total kinetic energy and linear momentum. This
573     principle is adopted in our simulations. Scaling is applied to the
574     velocities of the rigid bodies of SPC/E model water molecules, instead
575     of each hydrogen and oxygen atoms in relevant water molecules. As
576     shown in Figure \ref{spceGrad}, temperature gradients were established
577     similar to their system. However, the average temperature of our
578     system is 300K, while theirs is 318K, which would be attributed for
579     part of the difference between the final calculation results (Table
580     \ref{spceThermal}). Both methods yields values in agreement with
581     experiment. And this shows the applicability of our method to
582     multi-atom molecular system.
583 skuang 3563
584 skuang 3570 \begin{figure}
585     \includegraphics[width=\linewidth]{spceGrad}
586 skuang 3590 \caption{Temperature gradients in SPC/E water thermal conductivity
587     simulations.}
588 skuang 3570 \label{spceGrad}
589     \end{figure}
590    
591     \begin{table*}
592     \begin{minipage}{\linewidth}
593     \begin{center}
594    
595     \caption{Calculation results for thermal conductivity of SPC/E water
596     at ${\langle T\rangle}$ = 300K at various thermal exchange rates. Errors of
597     calculations in parentheses. }
598    
599     \begin{tabular}{cccc}
600     \hline
601     $\langle dT/dz\rangle$(K/\AA) & & $\lambda$(W/m/K) & \\
602 gezelter 3594 & This work & Previous simulations\cite{Bedrov:2000} &
603 skuang 3573 Experiment$^a$\\
604 skuang 3570 \hline
605 skuang 3573 0.38 & 0.816(0.044) & & 0.64\\
606     0.81 & 0.770(0.008) & 0.784\\
607     1.54 & 0.813(0.007) & 0.730\\
608 skuang 3570 \hline
609     \end{tabular}
610     \label{spceThermal}
611     \end{center}
612     \end{minipage}
613     \end{table*}
614    
615 skuang 3573 \subsubsection{Crystal Gold}
616 skuang 3580 Our results of gold thermal conductivity using two force fields are
617 skuang 3598 shown in Table \ref{AuThermal}. In these calculations,the end and
618     middle slabs were excluded in thermal gradient regession and only used
619     as heat source and drain in the systems. Our yielded values using EAM
620     force field are slightly larger than those using QSC force
621     field. However, both series are significantly smaller than
622     experimental value by a factor of more than 200. It has been verified
623     that this difference is mainly attributed to the lack of electron
624     interaction representation in these force field parameters. Richardson
625     {\it et al.}\cite{Clancy:1992} used EAM force field parameters in
626     their metal thermal conductivity calculations. The Non-Equilibrium MD
627     method they employed in their simulations produced comparable results
628     to ours. As Zhang {\it et al.}\cite{ISI:000231042800044} stated,
629     thermal conductivity values are influenced mainly by force
630     field. Another factor that affects the calculation results could be
631     the density of the metal. After equilibration under
632     isobaric-isothermal conditions, our crystall simulation cell expanded
633     by the order of 1\%. Under longer lattice constant than default,
634 skuang 3599 lower thermal conductance would be expected. Furthermore, the result
635     of Richardson {\it et al.} were obtained between 300K and 850K, which
636     are significantly higher than in our simulations. Therefore, it is
637     still confident to conclude that NIVS-RNEMD is applicable to metal
638     force field system.
639 skuang 3570
640     \begin{table*}
641     \begin{minipage}{\linewidth}
642     \begin{center}
643    
644     \caption{Calculation results for thermal conductivity of crystal gold
645 skuang 3598 using different force fields at ${\langle T\rangle}$ = 300K at
646     various thermal exchange rates. Errors of calculations in parentheses.}
647 skuang 3570
648 skuang 3598 \begin{tabular}{ccc}
649 skuang 3570 \hline
650 skuang 3598 Force Field Used & $\langle dT/dz\rangle$(K/\AA) & $\lambda$(W/m/K)\\
651 skuang 3570 \hline
652 skuang 3598 & 1.44 & 1.10(0.01)\\
653     QSC & 2.86 & 1.08(0.02)\\
654     & 5.14 & 1.15(0.01)\\
655 skuang 3570 \hline
656 skuang 3598 & 1.24 & 1.24(0.06)\\
657     EAM & 2.06 & 1.37(0.04)\\
658     & 2.55 & 1.41(0.03)\\
659 skuang 3580 \hline
660     \end{tabular}
661 skuang 3598 \label{AuThermal}
662 skuang 3580 \end{center}
663     \end{minipage}
664     \end{table*}
665    
666    
667 skuang 3573 \subsection{Interfaciel Thermal Conductivity}
668 skuang 3581 After simulations of homogeneous water and gold systems using
669     NIVS-RNEMD method were proved valid, calculation of gold/water
670     interfacial thermal conductivity was followed. It is found out that
671     the low interfacial conductance is probably due to the hydrophobic
672 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
673 skuang 3581 density change along $z$-axis, which is perpendicular to the
674     gold/water interface. It is observed that water density significantly
675     decreases when approaching the surface. Under this low thermal
676     conductance, both gold and water phase have sufficient time to
677     eliminate temperature difference inside respectively (Figure
678 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
679 skuang 3581 within respective phase, it is valid to assume that the temperature
680     difference between gold and water on surface would be approximately
681     the same as the difference between the gold and water phase. This
682     assumption enables convenient calculation of $G$ using
683     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
684     layer of water and gold close enough to surface, which would have
685     greater fluctuation and lower accuracy. Reported results (Table
686     \ref{interfaceRes}) are of two orders of magnitude smaller than our
687     calculations on homogeneous systems, and thus have larger relative
688     errors than our calculation results on homogeneous systems.
689 skuang 3573
690 skuang 3571 \begin{figure}
691 skuang 3595 \includegraphics[width=\linewidth]{interface}
692     \caption{Simulation results for Gold/Water interfacial thermal
693     conductivity: (a) Significant water density decrease is observed on
694 skuang 3597 crystalline gold surface, which indicates low surface contact and
695     leads to low thermal conductance. (b) Temperature profiles for a
696     series of simulations. Temperatures of different slabs in the same
697     phase show no significant differences.}
698 skuang 3595 \label{interface}
699 skuang 3571 \end{figure}
700    
701 skuang 3572 \begin{table*}
702     \begin{minipage}{\linewidth}
703     \begin{center}
704    
705     \caption{Calculation results for interfacial thermal conductivity
706     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
707     rates. Errors of calculations in parentheses. }
708    
709     \begin{tabular}{cccc}
710     \hline
711     $J_z$(MW/m$^2$) & $T_{gold}$ & $T_{water}$ & $G$(MW/m$^2$/K)\\
712     \hline
713 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
714     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
715     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
716     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
717 skuang 3572 \hline
718     \end{tabular}
719 skuang 3574 \label{interfaceRes}
720 skuang 3572 \end{center}
721     \end{minipage}
722     \end{table*}
723    
724 skuang 3576 \subsection{Shear Viscosity}
725     Our calculations (Table \ref{shearRate}) shows that scale RNEMD method
726     produced comparable shear viscosity to swap RNEMD method. In Table
727     \ref{shearRate}, the names of the calculated samples are devided into
728     two parts. The first number refers to total slabs in one simulation
729     box. The second number refers to the swapping interval in swap method, or
730     in scale method the equilvalent swapping interval that the same
731     momentum flux would theoretically result in swap method. All the scale
732     method results were from simulations that had a scaling interval of 10
733     time steps. The average molecular momentum gradients of these samples
734 skuang 3590 are shown in Figure \ref{shear} (a) and (b).
735 skuang 3576
736     \begin{table*}
737     \begin{minipage}{\linewidth}
738     \begin{center}
739    
740     \caption{Calculation results for shear viscosity of Lennard-Jones
741     fluid at ${T^* = 0.72}$ and ${\rho^* = 0.85}$, with swap and scale
742     methods at various momentum exchange rates. Results in reduced
743     unit. Errors of calculations in parentheses. }
744    
745     \begin{tabular}{ccc}
746     \hline
747 skuang 3598 (Equilvalent) Exchange Interval (fs) & $\eta^*_{swap}$ &
748     $\eta^*_{scale}$\\
749 skuang 3576 \hline
750 skuang 3598 500 & 3.64(0.05) & 3.76(0.09)\\
751     1000 & 3.52(0.16) & 3.66(0.06)\\
752     2000 & 3.72(0.05) & 3.32(0.18)\\
753     2500 & 3.42(0.06) & 3.43(0.08)\\
754 skuang 3576 \hline
755     \end{tabular}
756     \label{shearRate}
757     \end{center}
758     \end{minipage}
759     \end{table*}
760    
761     \begin{figure}
762 skuang 3590 \includegraphics[width=\linewidth]{shear}
763     \caption{Average momentum gradients in shear viscosity simulations,
764     using (a) ``swapping'' method and (b) NIVS-RNEMD method
765     respectively. (c) Temperature difference among x and y, z dimensions
766     observed when using NIVS-RNEMD with equivalent exchange interval of
767     500 fs.}
768     \label{shear}
769 skuang 3576 \end{figure}
770    
771     However, observations of temperatures along three dimensions show that
772     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
773 skuang 3590 two slabs which were scaled. Figure \ref{shear} (c) indicate that with
774 skuang 3576 relatively large imposed momentum flux, the temperature difference among $x$
775     and the other two dimensions was significant. This would result from the
776     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
777     momentum gradient is set up, $P_c^x$ would be roughly stable
778     ($<0$). Consequently, scaling factor $x$ would most probably larger
779     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
780     keep increase after most scaling steps. And if there is not enough time
781     for the kinetic energy to exchange among different dimensions and
782     different slabs, the system would finally build up temperature
783     (kinetic energy) difference among the three dimensions. Also, between
784     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
785     are closer to neighbor slabs. This is due to momentum transfer along
786     $z$ dimension between slabs.
787    
788     Although results between scaling and swapping methods are comparable,
789     the inherent temperature inhomogeneity even in relatively low imposed
790     exchange momentum flux simulations makes scaling RNEMD method less
791     attractive than swapping RNEMD in shear viscosity calculation.
792    
793 skuang 3574 \section{Conclusions}
794     NIVS-RNEMD simulation method is developed and tested on various
795 skuang 3581 systems. Simulation results demonstrate its validity in thermal
796     conductivity calculations, from Lennard-Jones fluid to multi-atom
797     molecule like water and metal crystals. NIVS-RNEMD improves
798     non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
799     methods. Furthermore, it develops a valid means for unphysical thermal
800     transfer between different species of molecules, and thus extends its
801     applicability to interfacial systems. Our calculation of gold/water
802     interfacial thermal conductivity demonstrates this advantage over
803     previous RNEMD methods. NIVS-RNEMD has also limited application on
804     shear viscosity calculations, but could cause temperature difference
805     among different dimensions under high momentum flux. Modification is
806     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
807     calculations.
808 skuang 3572
809 gezelter 3524 \section{Acknowledgments}
810     Support for this project was provided by the National Science
811     Foundation under grant CHE-0848243. Computational time was provided by
812     the Center for Research Computing (CRC) at the University of Notre
813     Dame. \newpage
814    
815 gezelter 3583 \bibliographystyle{aip}
816 gezelter 3524 \bibliography{nivsRnemd}
817 gezelter 3583
818 gezelter 3524 \end{doublespace}
819     \end{document}
820