ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3611
Committed: Wed Jul 14 15:20:31 2010 UTC (14 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 40620 byte(s)
Log Message:
edits

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 gezelter 3609 \usepackage{multirow}
10 gezelter 3524 %\usepackage{booktabs}
11     %\usepackage{bibentry}
12     %\usepackage{mathrsfs}
13     \usepackage[ref]{overcite}
14     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
15     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
16     9.0in \textwidth 6.5in \brokenpenalty=10000
17    
18     % double space list of tables and figures
19     \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
20     \setlength{\abovecaptionskip}{20 pt}
21     \setlength{\belowcaptionskip}{30 pt}
22    
23     \renewcommand\citemid{\ } % no comma in optional referenc note
24    
25     \begin{document}
26    
27     \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
28    
29     \author{Shenyu Kuang and J. Daniel
30     Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
31     Department of Chemistry and Biochemistry,\\
32     University of Notre Dame\\
33     Notre Dame, Indiana 46556}
34    
35     \date{\today}
36    
37     \maketitle
38    
39     \begin{doublespace}
40    
41     \begin{abstract}
42 gezelter 3583 We present a new method for introducing stable non-equilibrium
43     velocity and temperature distributions in molecular dynamics
44 gezelter 3609 simulations of heterogeneous systems. This method extends earlier
45     Reverse Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
46     momentum exchange swapping moves that can create non-thermal
47     velocity distributions and are difficult to use for interfacial
48     calculations. By using non-isotropic velocity scaling (NIVS) on the
49     molecules in specific regions of a system, it is possible to impose
50     momentum or thermal flux between regions of a simulation and stable
51     thermal and momentum gradients can then be established. The scaling
52     method we have developed conserves the total linear momentum and
53     total energy of the system. To test the methods, we have computed
54     the thermal conductivity of model liquid and solid systems as well
55     as the interfacial thermal conductivity of a metal-water interface.
56     We find that the NIVS-RNEMD improves the problematic velocity
57     distributions that develop in other RNEMD 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 gezelter 3609 (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 gezelter 3609 kinetic energy from the cold slab to the hot slab. If the hot and
229     cold slabs are separated along the z-axis, the energy flux is given
230     simply by the decrease in kinetic energy of the cold bin:
231 gezelter 3524 \begin{equation}
232 skuang 3534 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
233 gezelter 3524 \end{equation}
234     The expression for the energy flux can be re-written as another
235     ellipsoid centered on $(x,y,z) = 0$:
236     \begin{equation}
237 skuang 3604 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
238     K_c^\alpha -J_z \Delta t
239 gezelter 3524 \label{eq:fluxEllipsoid}
240     \end{equation}
241 gezelter 3583 The spatial extent of the {\it thermal flux ellipsoid} is governed
242 gezelter 3600 both by the target flux, $J_z$ as well as the instantaneous values of
243     the kinetic energy components in the cold bin.
244 gezelter 3524
245     To satisfy an energetic flux as well as the conservation constraints,
246 gezelter 3600 we must determine the points ${x,y,z}$ that lie on both the constraint
247     ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
248     (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
249     ellipsoids in 3-dimensional space.
250 gezelter 3524
251 gezelter 3569 \begin{figure}
252     \includegraphics[width=\linewidth]{ellipsoids}
253 gezelter 3600 \caption{Velocity scaling coefficients which maintain both constant
254     energy and constant linear momentum of the system lie on the surface
255     of the {\it constraint ellipsoid} while points which generate the
256     target momentum flux lie on the surface of the {\it flux ellipsoid}.
257     The velocity distributions in the cold bin are scaled by only those
258     points which lie on both ellipsoids.}
259 gezelter 3569 \label{ellipsoids}
260     \end{figure}
261    
262 gezelter 3600 Since ellipsoids can be expressed as polynomials up to second order in
263     each of the three coordinates, finding the the intersection points of
264     two ellipsoids is isomorphic to finding the roots a polynomial of
265     degree 16. There are a number of polynomial root-finding methods in
266     the literature, [CITATIONS NEEDED] but numerically finding the roots
267     of high-degree polynomials is generally an ill-conditioned
268 gezelter 3609 problem.[CITATION NEEDED] One simplification is to maintain velocity
269     scalings that are {\it as isotropic as possible}. To do this, we
270     impose $x=y$, and to treat both the constraint and flux ellipsoids as
271     2-dimensional ellipses. In reduced dimensionality, the
272 gezelter 3600 intersecting-ellipse problem reduces to finding the roots of
273 gezelter 3609 polynomials of degree 4.
274 gezelter 3600
275     Depending on the target flux and current velocity distributions, the
276     ellipsoids can have between 0 and 4 intersection points. If there are
277     no intersection points, it is not possible to satisfy the constraints
278     while performing a non-equilibrium scaling move, and no change is made
279     to the dynamics.
280    
281     With multiple intersection points, any of the scaling points will
282     conserve the linear momentum and kinetic energy of the system and will
283     generate the correct target flux. Although this method is inherently
284     non-isotropic, the goal is still to maintain the system as close to an
285     isotropic fluid as possible. With this in mind, we would like the
286     kinetic energies in the three different directions could become as
287     close as each other as possible after each scaling. Simultaneously,
288     one would also like each scaling as gentle as possible, i.e. ${x,y,z
289     \rightarrow 1}$, in order to avoid large perturbation to the system.
290 gezelter 3609 To do this, we pick the intersection point which maintains the three
291     scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
292 gezelter 3600 ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
293    
294     After the valid scaling parameters are arrived at by solving geometric
295     intersection problems in $x, y, z$ space in order to obtain cold slab
296     scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
297     determine the conjugate hot slab scaling variables.
298    
299     \subsection{Introducing shear stress via velocity scaling}
300 gezelter 3609 It is also possible to use this method to magnify the random
301     fluctuations of the average momentum in each of the bins to induce a
302     momentum flux. Doing this repeatedly will create a shear stress on
303     the system which will respond with an easily-measured strain. The
304     momentum flux (say along the $x$-direction) may be defined as:
305 gezelter 3524 \begin{equation}
306 skuang 3565 (1-x) P_c^x = j_z(p_x)\Delta t
307 skuang 3531 \label{eq:fluxPlane}
308 gezelter 3524 \end{equation}
309 gezelter 3600 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
310     its position governed both by a target value, $j_z(p_x)$ as well as
311     the instantaneous value of the momentum along the $x$-direction.
312 gezelter 3524
313 gezelter 3583 In order to satisfy a momentum flux as well as the conservation
314     constraints, we must determine the points ${x,y,z}$ which lie on both
315     the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
316     flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
317 gezelter 3600 ellipsoid and a plane in 3-dimensional space.
318 gezelter 3524
319 gezelter 3600 In the case of momentum flux transfer, we also impose another
320 gezelter 3609 constraint to set the kinetic energy transfer as zero. In other
321     words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
322 gezelter 3600 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
323     set of quartic equations to the above kinetic energy transfer problem.
324 gezelter 3524
325 gezelter 3600 \section{Computational Details}
326 gezelter 3583
327 gezelter 3609 We have implemented this methodology in our molecular dynamics code,
328     OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
329     after each MD step. We have tested the method in a variety of
330     different systems, including homogeneous fluids (Lennard-Jones and
331     SPC/E water), crystalline solids ({\sc eam}~\cite{PhysRevB.33.7983} and
332     quantum Sutton-Chen ({\sc q-sc})~\cite{PhysRevB.59.3527}
333     models for Gold), and heterogeneous interfaces (QSC gold - SPC/E
334     water). The last of these systems would have been difficult to study
335     using previous RNEMD methods, but using velocity scaling moves, we can
336     even obtain estimates of the interfacial thermal conductivities ($G$).
337 gezelter 3524
338 gezelter 3609 \subsection{Simulation Cells}
339 gezelter 3524
340 gezelter 3609 In each of the systems studied, the dynamics was carried out in a
341     rectangular simulation cell using periodic boundary conditions in all
342     three dimensions. The cells were longer along the $z$ axis and the
343     space was divided into $N$ slabs along this axis (typically $N=20$).
344     The top slab ($n=1$) was designated the ``cold'' slab, while the
345     central slab ($n= N/2 + 1$) was designated the ``hot'' slab. In all
346     cases, simulations were first thermalized in canonical ensemble (NVT)
347     using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
348 gezelter 3600 microcanonical ensemble (NVE) before introducing any non-equilibrium
349     method.
350 skuang 3531
351 gezelter 3609 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
352 skuang 3531
353 gezelter 3609 In order to compare our new methodology with the original
354     M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
355     first performed simulations using the original technique.
356 skuang 3531
357 gezelter 3609 \subsection{RNEMD with NIVS scaling}
358    
359     For each simulation utilizing the swapping method, a corresponding
360     NIVS-RNEMD simulation was carried out using a target momentum flux set
361     to produce a the same momentum or energy flux exhibited in the
362     swapping simulation.
363    
364     To test the temperature homogeneity (and to compute transport
365     coefficients), directional momentum and temperature distributions were
366     accumulated for molecules in each of the slabs.
367    
368     \subsection{Shear viscosities}
369    
370     The momentum flux was calculated using the total non-physical momentum
371     transferred (${P_x}$) and the data collection time ($t$):
372 skuang 3534 \begin{equation}
373     j_z(p_x) = \frac{P_x}{2 t L_x L_y}
374     \end{equation}
375 gezelter 3609 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
376     box. The factor of two in the denominator is present because physical
377     momentum transfer occurs in two directions due to our periodic
378     boundary conditions. The velocity gradient ${\langle \partial v_x
379     /\partial z \rangle}$ was obtained using linear regression of the
380     velocity profiles in the bins. For Lennard-Jones simulations, shear
381     viscosities are reporte in reduced units (${\eta^* = \eta \sigma^2
382     (\varepsilon m)^{-1/2}}$).
383 skuang 3532
384 gezelter 3609 \subsection{Thermal Conductivities}
385 skuang 3534
386 gezelter 3609 The energy flux was calculated similarly to the momentum flux, using
387     the total non-physical energy transferred (${E_{total}}$) and the data
388     collection time $t$:
389 skuang 3534 \begin{equation}
390     J_z = \frac{E_{total}}{2 t L_x L_y}
391     \end{equation}
392 gezelter 3609 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
393     obtained by a linear regression of the temperature profile. For
394     Lennard-Jones simulations, thermal conductivities are reported in
395     reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
396     k_B^{-1}\varepsilon^{-1/2}}$).
397 skuang 3534
398 gezelter 3609 \subsection{Interfacial Thermal Conductivities}
399 skuang 3563
400 gezelter 3609 For materials with a relatively low interfacial conductance, and in
401     cases where the flux between the materials is small, the bulk regions
402     on either side of an interface rapidly come to a state in which the
403     two phases have relatively homogeneous (but distinct) temperatures.
404     In calculating the interfacial thermal conductivity $G$, this
405     assumption was made, and the conductance can be approximated as:
406 skuang 3573
407     \begin{equation}
408     G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
409     \langle T_{water}\rangle \right)}
410     \label{interfaceCalc}
411     \end{equation}
412 gezelter 3609 where ${E_{total}}$ is the imposed non-physical kinetic energy
413     transfer and ${\langle T_{gold}\rangle}$ and ${\langle
414     T_{water}\rangle}$ are the average observed temperature of gold and
415     water phases respectively.
416 skuang 3573
417 gezelter 3609 \section{Results}
418 skuang 3538
419 gezelter 3609 \subsection{Lennard-Jones Fluid}
420     2592 Lennard-Jones atoms were placed in an orthorhombic cell
421     ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
422     reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
423     direct comparison between our results and previous methods. These
424     simulations were carried out with a reduced timestep ${\tau^* =
425     4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
426     temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
427     conductivity calculations, simulations were first run under reduced
428     temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
429     ensemble, but other temperatures ([XXX, YYY, and ZZZ]) were also
430     sampled. The simulations included $10^5$ steps of equilibration
431     without any momentum flux, $10^5$ steps of stablization with an
432     imposed momentum transfer to create a gradient, and $10^6$ steps of
433     data collection under RNEMD.
434    
435 gezelter 3611 \subsubsection*{Thermal Conductivity}
436    
437 gezelter 3609 Our thermal conductivity calculations show that the NIVS method agrees
438     well with the swapping method. Four different swap intervals were
439     tested (Table \ref{thermalLJRes}). With a fixed 10 fs [WHY NOT REDUCED
440     UNITS???] scaling interval, the target exchange kinetic energy
441     produced equivalent kinetic energy flux as in the swapping method.
442     Similar thermal gradients were observed with similar thermal flux
443     under the two different methods (Figure \ref{thermalGrad}).
444    
445 skuang 3563 \begin{table*}
446 gezelter 3609 \begin{minipage}{\linewidth}
447     \begin{center}
448 skuang 3538
449 gezelter 3609 \caption{Thermal conductivity (in reduced units) of a
450     Lennard-Jones fluid at ${\langle T^* \rangle = 0.72}$ and
451     ${\rho^* = 0.85}$ for the swapping and scaling methods at
452     various kinetic energy exchange rates. Uncertainties are
453     indicated in parentheses.}
454    
455     \begin{tabular}{|cc|cc|}
456     \hline
457     \multicolumn{2}{|c|}{Swapping RNEMD} &
458     \multicolumn{2}{|c|}{NIVS-RNEMD} \\
459     \hline
460     Swap Interval (fs) & $\lambda^*_{swap}$ & Equilvalent $J_z^*$ & $\lambda^*_{scale}$\\
461     \hline
462     250 & 7.03(0.34) & 0.16 & 7.30(0.10)\\
463     500 & 7.03(0.14) & 0.09 & 6.95(0.09)\\
464     1000 & 6.91(0.42) & 0.047 & 7.19(0.07)\\
465     2000 & 7.52(0.15) & 0.024 & 7.19(0.28)\\
466     \hline
467     \end{tabular}
468     \label{thermalLJRes}
469     \end{center}
470     \end{minipage}
471 skuang 3563 \end{table*}
472    
473     \begin{figure}
474 skuang 3567 \includegraphics[width=\linewidth]{thermalGrad}
475 gezelter 3609 \caption{NIVS-RNEMD method creates similar temperature gradients
476     compared with the swapping method under a variety of imposed kinetic
477     energy flux values.}
478 skuang 3567 \label{thermalGrad}
479 skuang 3563 \end{figure}
480    
481 gezelter 3609 During these simulations, velocities were recorded every 1000 steps
482     and was used to produce distributions of both velocity and speed in
483     each of the slabs. From these distributions, we observed that under
484     relatively high non-physical kinetic energy flux, the spee of
485     molecules in slabs where swapping occured could deviate from the
486     Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
487     and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
488     distributions deviate from an ideal distribution. In the ``hot'' slab,
489     the probability density is notched at low speeds and has a substantial
490     shoulder at higher speeds relative to the ideal MB distribution. In
491     the cold slab, the opposite notching and shouldering occurs. This
492     phenomenon is more obvious at higher swapping rates.
493 skuang 3563
494 gezelter 3609 In the velocity distributions, the ideal Gaussian peak is
495     substantially flattened in the hot slab, and is overly sharp (with
496     truncated wings) in the cold slab. This problem is rooted in the
497     mechanism of the swapping method. Continually depleting low (high)
498     speed particles in the high (low) temperature slab is not complemented
499     by diffusions of low (high) speed particles from neighboring slabs,
500     unless the swapping rate is sufficiently small. Simutaneously, surplus
501     low speed particles in the low temperature slab do not have sufficient
502     time to diffuse to neighboring slabs. Since the thermal exchange rate
503     must reach a minimum level to produce an observable thermal gradient,
504     the swapping-method RNEMD has a relatively narrow choice of exchange
505     times that can be utilized.
506 skuang 3578
507 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
508     Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
509     this is simple; upon velocity scaling, a Gaussian distribution remains
510     Gaussian. Although a single scaling move is non-isotropic in three
511     dimensions, our criteria for choosing a set of scaling coefficients
512     helps maintain the distributions as close to isotropic as possible.
513     Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
514     as the previous RNEMD methods but without large perturbations to the
515     velocity distributions in the two slabs.
516    
517 skuang 3568 \begin{figure}
518 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
519     \caption{Speed distribution for thermal conductivity using a)
520     ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
521     simulations with an exchange or equilvalent exchange interval of 250
522 skuang 3593 fs. In circled areas, distributions from ``swapping'' RNEMD
523     simulation have deviation from ideal Maxwell-Boltzmann distribution
524     (curves fit for each distribution).}
525 skuang 3589 \label{thermalHist}
526 skuang 3568 \end{figure}
527    
528 gezelter 3611
529     \subsubsection*{Shear Viscosity}
530     Our calculations (Table \ref{shearRate}) show that velocity-scaling
531     RNEMD predicted comparable shear viscosities to swap RNEMD method. All
532     the scale method results were from simulations that had a scaling
533     interval of 10 time steps. The average molecular momentum gradients of
534     these samples are shown in Figure \ref{shear} (a) and (b).
535    
536     \begin{table*}
537     \begin{minipage}{\linewidth}
538     \begin{center}
539    
540     \caption{Shear viscosities of Lennard-Jones fluid at ${T^* =
541     0.72}$ and ${\rho^* = 0.85}$ using swapping and NIVS methods
542     at various momentum exchange rates. Uncertainties are
543     indicated in parentheses. }
544    
545     \begin{tabular}{ccccc}
546     Swapping method & & & NIVS-RNEMD & \\
547     \hline
548     Swap Interval (fs) & $\eta^*_{swap}$ & & Equilvalent $j_p^*(v_x)$ &
549     $\eta^*_{scale}$\\
550     \hline
551     500 & 3.64(0.05) & & 0.09 & 3.76(0.09)\\
552     1000 & 3.52(0.16) & & 0.046 & 3.66(0.06)\\
553     2000 & 3.72(0.05) & & 0.024 & 3.32(0.18)\\
554     2500 & 3.42(0.06) & & 0.019 & 3.43(0.08)\\
555     \hline
556     \end{tabular}
557     \label{shearRate}
558     \end{center}
559     \end{minipage}
560     \end{table*}
561    
562     \begin{figure}
563     \includegraphics[width=\linewidth]{shear}
564     \caption{Average momentum gradients in shear viscosity simulations,
565     using (a) ``swapping'' method and (b) NIVS-RNEMD method
566     respectively. (c) Temperature difference among x and y, z dimensions
567     observed when using NIVS-RNEMD with equivalent exchange interval of
568     500 fs.}
569     \label{shear}
570     \end{figure}
571    
572     However, observations of temperatures along three dimensions show that
573     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
574     two slabs which were scaled. Figure \ref{shear} (c) indicate that with
575     relatively large imposed momentum flux, the temperature difference among $x$
576     and the other two dimensions was significant. This would result from the
577     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
578     momentum gradient is set up, $P_c^x$ would be roughly stable
579     ($<0$). Consequently, scaling factor $x$ would most probably larger
580     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
581     keep increase after most scaling steps. And if there is not enough time
582     for the kinetic energy to exchange among different dimensions and
583     different slabs, the system would finally build up temperature
584     (kinetic energy) difference among the three dimensions. Also, between
585     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
586     are closer to neighbor slabs. This is due to momentum transfer along
587     $z$ dimension between slabs.
588    
589     Although results between scaling and swapping methods are comparable,
590     the inherent temperature inhomogeneity even in relatively low imposed
591     exchange momentum flux simulations makes scaling RNEMD method less
592     attractive than swapping RNEMD in shear viscosity calculation.
593    
594    
595 gezelter 3609 \subsection{Bulk SPC/E water}
596    
597     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
598     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
599     the original swapping RNEMD method. Bedrov {\it et
600 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
601 skuang 3579 center-of-mass velocities instead of single atom velocities in a
602 gezelter 3609 molecule conserves the total kinetic energy and linear momentum. This
603     principle is also adopted in our simulations. Scaling was applied to
604     the center-of-mass velocities of the rigid bodies of SPC/E model water
605     molecules.
606 skuang 3563
607 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
608     molecules were first equilibrated under ambient pressure and
609     temperature conditions using the isobaric-isothermal (NPT)
610     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
611     average volume observed in the NPT simulations, and this was followed
612     by equilibration, first in the canonical (NVT) ensemble, followed by a
613     [XXX ps] period under constant-NVE conditions without any momentum
614     flux. [YYY ps] was allowed to stabilize the system with an imposed
615     momentum transfer to create a gradient, and [ZZZ ps] was alotted for
616     data collection under RNEMD.
617    
618     As shown in Figure \ref{spceGrad}, temperature gradients were
619     established similar to the previous work. However, the average
620     temperature of our system is 300K, while that in Bedrov {\it et al.}
621     is 318K, which would be attributed for part of the difference between
622     the final calculation results (Table \ref{spceThermal}). [WHY DIDN'T
623     WE DO 318 K?] Both methods yield values in reasonable agreement with
624     experiment [DONE AT WHAT TEMPERATURE?]
625    
626 skuang 3570 \begin{figure}
627 gezelter 3609 \includegraphics[width=\linewidth]{spceGrad}
628     \caption{Temperature gradients in SPC/E water thermal conductivity
629     simulations.}
630     \label{spceGrad}
631 skuang 3570 \end{figure}
632    
633     \begin{table*}
634 gezelter 3609 \begin{minipage}{\linewidth}
635     \begin{center}
636    
637     \caption{Thermal conductivity of SPC/E water under various
638     imposed thermal gradients. Uncertainties are indicated in
639     parentheses.}
640    
641     \begin{tabular}{|c|ccc|}
642     \hline
643     $\langle dT/dz\rangle$(K/\AA) & \multicolumn{3}{|c|}{$\lambda
644     (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
645     & This work (300K) & Previous simulations (318K)\cite{Bedrov:2000} &
646     Experiment\cite{WagnerKruse}\\
647     \hline
648     0.38 & 0.816(0.044) & & 0.64\\
649     0.81 & 0.770(0.008) & 0.784 & \\
650     1.54 & 0.813(0.007) & 0.730 & \\
651     \hline
652     \end{tabular}
653     \label{spceThermal}
654     \end{center}
655     \end{minipage}
656     \end{table*}
657 skuang 3570
658 gezelter 3609 \subsection{Crystalline Gold}
659 skuang 3570
660 gezelter 3609 To see how the method performed in a solid, we calculated thermal
661     conductivities using two atomistic models for gold. Several different
662     potential models have been developed that reasonably describe
663     interactions in transition metals. In particular, the Embedded Atom
664     Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc
665     sc})~\cite{Chen90} potential have been used to study a wide range of
666     phenomena in both bulk materials and
667     nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
668     Both potentials are based on a model of a metal which treats the
669     nuclei and core electrons as pseudo-atoms embedded in the electron
670     density due to the valence electrons on all of the other atoms in the
671     system. The {\sc sc} potential has a simple form that closely
672     resembles the Lennard Jones potential,
673     \begin{equation}
674     \label{eq:SCP1}
675     U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
676     \end{equation}
677     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
678     \begin{equation}
679     \label{eq:SCP2}
680     V^{pair}_{ij}(r)=\left( \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}.
681     \end{equation}
682     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
683     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
684     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
685     the interactions between the valence electrons and the cores of the
686     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
687     scale, $c_i$ scales the attractive portion of the potential relative
688     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
689     that assures a dimensionless form for $\rho$. These parameters are
690     tuned to various experimental properties such as the density, cohesive
691     energy, and elastic moduli for FCC transition metals. The quantum
692     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
693     including zero-point quantum corrections for different transition
694     metals.\cite{PhysRevB.59.3527} The {\sc eam} functional forms differ
695     slightly from {\sc sc} but the overall method is very similar.
696 skuang 3570
697 gezelter 3609 In this work, we have utilized both the {\sc eam} and the {\sc q-sc}
698     potentials to test the behavior of scaling RNEMD.
699 skuang 3570
700 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
701     atoms. [LxMxN UNIT CELLS]. Simulations were run both with and
702     without isobaric-isothermal (NPT)~\cite{melchionna93}
703     pre-equilibration at a target pressure of 1 atm. When equilibrated
704     under NPT conditions, our simulation box expanded by approximately 1\%
705 gezelter 3610 [IN VOLUME OR LINEAR DIMENSIONS ?]. Following adjustment of the box
706     volume, equilibrations in both the canonical and microcanonical
707     ensembles were carried out. With the simulation cell divided evenly
708     into 10 slabs, different thermal gradients were established by
709     applying a set of target thermal transfer fluxes.
710 skuang 3570
711 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
712     \ref{AuThermal}. In these calculations, the end and middle slabs were
713 gezelter 3610 excluded in thermal gradient linear regession. {\sc eam} predicts
714 gezelter 3609 slightly larger thermal conductivities than {\sc q-sc}. However, both
715     values are smaller than experimental value by a factor of more than
716     200. This behavior has been observed previously by Richardson and
717     Clancy, and has been attributed to the lack of electronic effects in
718     these force fields.\cite{Clancy:1992} The non-equilibrium MD method
719 gezelter 3610 employed in their simulations produced comparable results to ours. It
720     should be noted that the density of the metal being simulated also
721     greatly affects the thermal conductivity. With an expanded lattice,
722     lower thermal conductance is expected (and observed). We also observed
723     a decrease in thermal conductance at higher temperatures, a trend that
724     agrees with experimental measurements [PAGE
725 gezelter 3609 NUMBERS?].\cite{AshcroftMermin}
726 skuang 3570
727 gezelter 3609 \begin{table*}
728     \begin{minipage}{\linewidth}
729     \begin{center}
730    
731     \caption{Calculated thermal conductivity of crystalline gold
732     using two related force fields. Calculations were done at both
733     experimental and equilibrated densities and at a range of
734     temperatures and thermal flux rates. Uncertainties are
735     indicated in parentheses. [CLANCY COMPARISON? SWAPPING
736     COMPARISON?]}
737    
738     \begin{tabular}{|c|c|c|cc|}
739     \hline
740     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
741     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
742     \hline
743     \multirow{7}{*}{\sc q-sc} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
744     & & & 2.86 & 1.08(0.05)\\
745     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
746     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
747     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
748     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
749     & & & 4.84 & 0.92(0.05)\\
750     \hline
751     \multirow{8}{*}{\sc eam} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
752     & & & 2.06 & 1.37(0.04)\\
753     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
754     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
755     & & & 2.04 & 1.41(0.07)\\
756     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
757     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
758     & & & 4.14 & 1.08(0.05)\\
759     \hline
760     \end{tabular}
761     \label{AuThermal}
762     \end{center}
763     \end{minipage}
764 skuang 3580 \end{table*}
765    
766 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
767     The most attractive aspect of the scaling approach for RNEMD is the
768     ability to use the method in non-homogeneous systems, where molecules
769     of different identities are segregated in different slabs. To test
770     this application, we simulated a Gold (111) / water interface. To
771     construct the interface, a box containing a lattice of 1188 Au atoms
772     (with the 111 surface in the +z and -z directions) was allowed to
773     relax under ambient temperature and pressure. A separate (but
774     identically sized) box of SPC/E water was also equilibrated at ambient
775     conditions. The two boxes were combined by removing all water
776     molecules withing 3 \AA radius of any gold atom. The final
777     configuration contained 1862 SPC/E water molecules.
778 skuang 3580
779 gezelter 3609 After simulations of bulk water and crystal gold, a mixture system was
780     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
781     molecules. Spohr potential was adopted in depicting the interaction
782     between metal atom and water molecule.\cite{ISI:000167766600035} A
783     similar protocol of equilibration was followed. Several thermal
784     gradients was built under different target thermal flux. It was found
785     out that compared to our previous simulation systems, the two phases
786     could have large temperature difference even under a relatively low
787     thermal flux.
788    
789    
790 skuang 3581 After simulations of homogeneous water and gold systems using
791     NIVS-RNEMD method were proved valid, calculation of gold/water
792     interfacial thermal conductivity was followed. It is found out that
793     the low interfacial conductance is probably due to the hydrophobic
794 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
795 skuang 3581 density change along $z$-axis, which is perpendicular to the
796     gold/water interface. It is observed that water density significantly
797     decreases when approaching the surface. Under this low thermal
798     conductance, both gold and water phase have sufficient time to
799     eliminate temperature difference inside respectively (Figure
800 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
801 skuang 3581 within respective phase, it is valid to assume that the temperature
802     difference between gold and water on surface would be approximately
803     the same as the difference between the gold and water phase. This
804     assumption enables convenient calculation of $G$ using
805     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
806     layer of water and gold close enough to surface, which would have
807     greater fluctuation and lower accuracy. Reported results (Table
808     \ref{interfaceRes}) are of two orders of magnitude smaller than our
809     calculations on homogeneous systems, and thus have larger relative
810     errors than our calculation results on homogeneous systems.
811 skuang 3573
812 skuang 3571 \begin{figure}
813 skuang 3595 \includegraphics[width=\linewidth]{interface}
814     \caption{Simulation results for Gold/Water interfacial thermal
815     conductivity: (a) Significant water density decrease is observed on
816 skuang 3597 crystalline gold surface, which indicates low surface contact and
817     leads to low thermal conductance. (b) Temperature profiles for a
818     series of simulations. Temperatures of different slabs in the same
819     phase show no significant differences.}
820 skuang 3595 \label{interface}
821 skuang 3571 \end{figure}
822    
823 skuang 3572 \begin{table*}
824     \begin{minipage}{\linewidth}
825     \begin{center}
826    
827     \caption{Calculation results for interfacial thermal conductivity
828     at ${\langle T\rangle \sim}$ 300K at various thermal exchange
829     rates. Errors of calculations in parentheses. }
830    
831     \begin{tabular}{cccc}
832     \hline
833 skuang 3606 $J_z$ (MW/m$^2$) & $T_{gold}$ (K) & $T_{water}$ (K) & $G$
834     (MW/m$^2$/K)\\
835 skuang 3572 \hline
836 skuang 3573 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
837     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
838     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
839     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
840 skuang 3572 \hline
841     \end{tabular}
842 skuang 3574 \label{interfaceRes}
843 skuang 3572 \end{center}
844     \end{minipage}
845     \end{table*}
846    
847 skuang 3576
848 skuang 3574 \section{Conclusions}
849     NIVS-RNEMD simulation method is developed and tested on various
850 skuang 3581 systems. Simulation results demonstrate its validity in thermal
851     conductivity calculations, from Lennard-Jones fluid to multi-atom
852     molecule like water and metal crystals. NIVS-RNEMD improves
853     non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
854     methods. Furthermore, it develops a valid means for unphysical thermal
855     transfer between different species of molecules, and thus extends its
856     applicability to interfacial systems. Our calculation of gold/water
857     interfacial thermal conductivity demonstrates this advantage over
858     previous RNEMD methods. NIVS-RNEMD has also limited application on
859     shear viscosity calculations, but could cause temperature difference
860     among different dimensions under high momentum flux. Modification is
861     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
862     calculations.
863 skuang 3572
864 gezelter 3524 \section{Acknowledgments}
865     Support for this project was provided by the National Science
866     Foundation under grant CHE-0848243. Computational time was provided by
867     the Center for Research Computing (CRC) at the University of Notre
868     Dame. \newpage
869    
870 gezelter 3583 \bibliographystyle{aip}
871 gezelter 3524 \bibliography{nivsRnemd}
872 gezelter 3583
873 gezelter 3524 \end{doublespace}
874     \end{document}
875