ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3612
Committed: Wed Jul 14 15:52:18 2010 UTC (14 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 40568 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 gezelter 3612 tested (Table \ref{LJ}). With a fixed 10 fs [WHY NOT REDUCED
440 gezelter 3609 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 3612 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
450     ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
451     ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
452     at various momentum fluxes. The original swapping method and
453     the velocity scaling method give similar results.
454     Uncertainties are indicated in parentheses.}
455 gezelter 3609
456 gezelter 3612 \begin{tabular}{|cc|cc|cc|}
457 gezelter 3609 \hline
458 gezelter 3612 \multicolumn{2}{|c}{Momentum Exchange} &
459     \multicolumn{2}{|c}{Swapping RNEMD} &
460 gezelter 3609 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
461     \hline
462 gezelter 3612 \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
463    
464     \multirow{2}{*}{$\lambda^*_{swap}$} &
465     \multirow{2}{*}{$\eta^*_{swap}$} &
466     \multirow{2}{*}{$\lambda^*_{scale}$} &
467     \multirow{2}{*}{$\eta^*_{scale}$} \\
468     & $j_p^*(v_x)$ (reduced units) & & & & \\
469 gezelter 3609 \hline
470 gezelter 3612 250 & 0.16 & 7.03(0.34) & & 7.30(0.10) & \\
471     500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
472     1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
473     2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
474     2500 & 0.019 & & 3.42(0.06) & & 3.43(0.08)\\
475 gezelter 3609 \hline
476     \end{tabular}
477 gezelter 3612 \label{LJ}
478 gezelter 3609 \end{center}
479     \end{minipage}
480 skuang 3563 \end{table*}
481    
482     \begin{figure}
483 gezelter 3612 \includegraphics[width=\linewidth]{thermalGrad}
484     \caption{NIVS-RNEMD method creates similar temperature gradients
485     compared with the swapping method under a variety of imposed kinetic
486     energy flux values.}
487     \label{thermalGrad}
488 skuang 3563 \end{figure}
489    
490 gezelter 3612 \subsubsection*{Velocity Distributions}
491    
492 gezelter 3609 During these simulations, velocities were recorded every 1000 steps
493     and was used to produce distributions of both velocity and speed in
494     each of the slabs. From these distributions, we observed that under
495     relatively high non-physical kinetic energy flux, the spee of
496     molecules in slabs where swapping occured could deviate from the
497     Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
498     and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
499     distributions deviate from an ideal distribution. In the ``hot'' slab,
500     the probability density is notched at low speeds and has a substantial
501     shoulder at higher speeds relative to the ideal MB distribution. In
502     the cold slab, the opposite notching and shouldering occurs. This
503     phenomenon is more obvious at higher swapping rates.
504 skuang 3563
505 gezelter 3609 In the velocity distributions, the ideal Gaussian peak is
506     substantially flattened in the hot slab, and is overly sharp (with
507     truncated wings) in the cold slab. This problem is rooted in the
508     mechanism of the swapping method. Continually depleting low (high)
509     speed particles in the high (low) temperature slab is not complemented
510     by diffusions of low (high) speed particles from neighboring slabs,
511     unless the swapping rate is sufficiently small. Simutaneously, surplus
512     low speed particles in the low temperature slab do not have sufficient
513     time to diffuse to neighboring slabs. Since the thermal exchange rate
514     must reach a minimum level to produce an observable thermal gradient,
515     the swapping-method RNEMD has a relatively narrow choice of exchange
516     times that can be utilized.
517 skuang 3578
518 gezelter 3609 For comparison, NIVS-RNEMD produces a speed distribution closer to the
519     Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
520     this is simple; upon velocity scaling, a Gaussian distribution remains
521     Gaussian. Although a single scaling move is non-isotropic in three
522     dimensions, our criteria for choosing a set of scaling coefficients
523     helps maintain the distributions as close to isotropic as possible.
524     Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
525     as the previous RNEMD methods but without large perturbations to the
526     velocity distributions in the two slabs.
527    
528 skuang 3568 \begin{figure}
529 skuang 3589 \includegraphics[width=\linewidth]{thermalHist}
530     \caption{Speed distribution for thermal conductivity using a)
531     ``swapping'' and b) NIVS- RNEMD methods. Shown is from the
532     simulations with an exchange or equilvalent exchange interval of 250
533 skuang 3593 fs. In circled areas, distributions from ``swapping'' RNEMD
534     simulation have deviation from ideal Maxwell-Boltzmann distribution
535     (curves fit for each distribution).}
536 skuang 3589 \label{thermalHist}
537 skuang 3568 \end{figure}
538    
539 gezelter 3611
540     \subsubsection*{Shear Viscosity}
541 gezelter 3612 Our calculations (Table \ref{LJ}) show that velocity-scaling
542 gezelter 3611 RNEMD predicted comparable shear viscosities to swap RNEMD method. All
543     the scale method results were from simulations that had a scaling
544     interval of 10 time steps. The average molecular momentum gradients of
545     these samples are shown in Figure \ref{shear} (a) and (b).
546    
547     \begin{figure}
548     \includegraphics[width=\linewidth]{shear}
549     \caption{Average momentum gradients in shear viscosity simulations,
550     using (a) ``swapping'' method and (b) NIVS-RNEMD method
551     respectively. (c) Temperature difference among x and y, z dimensions
552     observed when using NIVS-RNEMD with equivalent exchange interval of
553     500 fs.}
554     \label{shear}
555     \end{figure}
556    
557     However, observations of temperatures along three dimensions show that
558     inhomogeneity occurs in scaling RNEMD simulations, particularly in the
559     two slabs which were scaled. Figure \ref{shear} (c) indicate that with
560     relatively large imposed momentum flux, the temperature difference among $x$
561     and the other two dimensions was significant. This would result from the
562     algorithm of scaling method. From Eq. \ref{eq:fluxPlane}, after
563     momentum gradient is set up, $P_c^x$ would be roughly stable
564     ($<0$). Consequently, scaling factor $x$ would most probably larger
565     than 1. Therefore, the kinetic energy in $x$-dimension $K_c^x$ would
566     keep increase after most scaling steps. And if there is not enough time
567     for the kinetic energy to exchange among different dimensions and
568     different slabs, the system would finally build up temperature
569     (kinetic energy) difference among the three dimensions. Also, between
570     $y$ and $z$ dimensions in the scaled slabs, temperatures of $z$-axis
571     are closer to neighbor slabs. This is due to momentum transfer along
572     $z$ dimension between slabs.
573    
574     Although results between scaling and swapping methods are comparable,
575     the inherent temperature inhomogeneity even in relatively low imposed
576     exchange momentum flux simulations makes scaling RNEMD method less
577     attractive than swapping RNEMD in shear viscosity calculation.
578    
579    
580 gezelter 3609 \subsection{Bulk SPC/E water}
581    
582     We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
583     to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
584     the original swapping RNEMD method. Bedrov {\it et
585 gezelter 3594 al.}\cite{Bedrov:2000} argued that exchange of the molecule
586 skuang 3579 center-of-mass velocities instead of single atom velocities in a
587 gezelter 3609 molecule conserves the total kinetic energy and linear momentum. This
588     principle is also adopted in our simulations. Scaling was applied to
589     the center-of-mass velocities of the rigid bodies of SPC/E model water
590     molecules.
591 skuang 3563
592 gezelter 3609 To construct the simulations, a simulation box consisting of 1000
593     molecules were first equilibrated under ambient pressure and
594     temperature conditions using the isobaric-isothermal (NPT)
595     ensemble.\cite{melchionna93} A fixed volume was chosen to match the
596     average volume observed in the NPT simulations, and this was followed
597     by equilibration, first in the canonical (NVT) ensemble, followed by a
598     [XXX ps] period under constant-NVE conditions without any momentum
599     flux. [YYY ps] was allowed to stabilize the system with an imposed
600     momentum transfer to create a gradient, and [ZZZ ps] was alotted for
601     data collection under RNEMD.
602    
603     As shown in Figure \ref{spceGrad}, temperature gradients were
604     established similar to the previous work. However, the average
605     temperature of our system is 300K, while that in Bedrov {\it et al.}
606     is 318K, which would be attributed for part of the difference between
607     the final calculation results (Table \ref{spceThermal}). [WHY DIDN'T
608     WE DO 318 K?] Both methods yield values in reasonable agreement with
609     experiment [DONE AT WHAT TEMPERATURE?]
610    
611 skuang 3570 \begin{figure}
612 gezelter 3609 \includegraphics[width=\linewidth]{spceGrad}
613     \caption{Temperature gradients in SPC/E water thermal conductivity
614     simulations.}
615     \label{spceGrad}
616 skuang 3570 \end{figure}
617    
618     \begin{table*}
619 gezelter 3609 \begin{minipage}{\linewidth}
620     \begin{center}
621    
622     \caption{Thermal conductivity of SPC/E water under various
623     imposed thermal gradients. Uncertainties are indicated in
624     parentheses.}
625    
626     \begin{tabular}{|c|ccc|}
627     \hline
628 gezelter 3612 \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} & \multicolumn{3}{|c|}{$\lambda
629 gezelter 3609 (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
630     & This work (300K) & Previous simulations (318K)\cite{Bedrov:2000} &
631     Experiment\cite{WagnerKruse}\\
632     \hline
633     0.38 & 0.816(0.044) & & 0.64\\
634     0.81 & 0.770(0.008) & 0.784 & \\
635     1.54 & 0.813(0.007) & 0.730 & \\
636     \hline
637     \end{tabular}
638     \label{spceThermal}
639     \end{center}
640     \end{minipage}
641     \end{table*}
642 skuang 3570
643 gezelter 3609 \subsection{Crystalline Gold}
644 skuang 3570
645 gezelter 3609 To see how the method performed in a solid, we calculated thermal
646     conductivities using two atomistic models for gold. Several different
647     potential models have been developed that reasonably describe
648     interactions in transition metals. In particular, the Embedded Atom
649     Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc
650     sc})~\cite{Chen90} potential have been used to study a wide range of
651     phenomena in both bulk materials and
652     nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
653     Both potentials are based on a model of a metal which treats the
654     nuclei and core electrons as pseudo-atoms embedded in the electron
655     density due to the valence electrons on all of the other atoms in the
656     system. The {\sc sc} potential has a simple form that closely
657     resembles the Lennard Jones potential,
658     \begin{equation}
659     \label{eq:SCP1}
660     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] ,
661     \end{equation}
662     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
663     \begin{equation}
664     \label{eq:SCP2}
665     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}}.
666     \end{equation}
667     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
668     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
669     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
670     the interactions between the valence electrons and the cores of the
671     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
672     scale, $c_i$ scales the attractive portion of the potential relative
673     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
674     that assures a dimensionless form for $\rho$. These parameters are
675     tuned to various experimental properties such as the density, cohesive
676     energy, and elastic moduli for FCC transition metals. The quantum
677     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
678     including zero-point quantum corrections for different transition
679     metals.\cite{PhysRevB.59.3527} The {\sc eam} functional forms differ
680     slightly from {\sc sc} but the overall method is very similar.
681 skuang 3570
682 gezelter 3609 In this work, we have utilized both the {\sc eam} and the {\sc q-sc}
683     potentials to test the behavior of scaling RNEMD.
684 skuang 3570
685 gezelter 3609 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
686     atoms. [LxMxN UNIT CELLS]. Simulations were run both with and
687     without isobaric-isothermal (NPT)~\cite{melchionna93}
688     pre-equilibration at a target pressure of 1 atm. When equilibrated
689     under NPT conditions, our simulation box expanded by approximately 1\%
690 gezelter 3610 [IN VOLUME OR LINEAR DIMENSIONS ?]. Following adjustment of the box
691     volume, equilibrations in both the canonical and microcanonical
692     ensembles were carried out. With the simulation cell divided evenly
693     into 10 slabs, different thermal gradients were established by
694     applying a set of target thermal transfer fluxes.
695 skuang 3570
696 gezelter 3609 The results for the thermal conductivity of gold are shown in Table
697     \ref{AuThermal}. In these calculations, the end and middle slabs were
698 gezelter 3610 excluded in thermal gradient linear regession. {\sc eam} predicts
699 gezelter 3609 slightly larger thermal conductivities than {\sc q-sc}. However, both
700     values are smaller than experimental value by a factor of more than
701     200. This behavior has been observed previously by Richardson and
702     Clancy, and has been attributed to the lack of electronic effects in
703     these force fields.\cite{Clancy:1992} The non-equilibrium MD method
704 gezelter 3610 employed in their simulations produced comparable results to ours. It
705     should be noted that the density of the metal being simulated also
706     greatly affects the thermal conductivity. With an expanded lattice,
707     lower thermal conductance is expected (and observed). We also observed
708     a decrease in thermal conductance at higher temperatures, a trend that
709     agrees with experimental measurements [PAGE
710 gezelter 3609 NUMBERS?].\cite{AshcroftMermin}
711 skuang 3570
712 gezelter 3609 \begin{table*}
713     \begin{minipage}{\linewidth}
714     \begin{center}
715    
716     \caption{Calculated thermal conductivity of crystalline gold
717     using two related force fields. Calculations were done at both
718     experimental and equilibrated densities and at a range of
719     temperatures and thermal flux rates. Uncertainties are
720     indicated in parentheses. [CLANCY COMPARISON? SWAPPING
721     COMPARISON?]}
722    
723     \begin{tabular}{|c|c|c|cc|}
724     \hline
725     Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
726     $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
727     \hline
728     \multirow{7}{*}{\sc q-sc} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
729     & & & 2.86 & 1.08(0.05)\\
730     & & & 5.14 & 1.15(0.07)\\\cline{2-5}
731     & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
732     & & & 3.02 & 1.26(0.05)\\\cline{3-5}
733     & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
734     & & & 4.84 & 0.92(0.05)\\
735     \hline
736     \multirow{8}{*}{\sc eam} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
737     & & & 2.06 & 1.37(0.04)\\
738     & & & 2.55 & 1.41(0.07)\\\cline{2-5}
739     & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
740     & & & 2.04 & 1.41(0.07)\\
741     & & & 2.41 & 1.53(0.10)\\\cline{3-5}
742     & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
743     & & & 4.14 & 1.08(0.05)\\
744     \hline
745     \end{tabular}
746     \label{AuThermal}
747     \end{center}
748     \end{minipage}
749 skuang 3580 \end{table*}
750    
751 gezelter 3609 \subsection{Thermal Conductance at the Au/H$_2$O interface}
752     The most attractive aspect of the scaling approach for RNEMD is the
753     ability to use the method in non-homogeneous systems, where molecules
754     of different identities are segregated in different slabs. To test
755     this application, we simulated a Gold (111) / water interface. To
756     construct the interface, a box containing a lattice of 1188 Au atoms
757     (with the 111 surface in the +z and -z directions) was allowed to
758     relax under ambient temperature and pressure. A separate (but
759     identically sized) box of SPC/E water was also equilibrated at ambient
760     conditions. The two boxes were combined by removing all water
761     molecules withing 3 \AA radius of any gold atom. The final
762     configuration contained 1862 SPC/E water molecules.
763 skuang 3580
764 gezelter 3609 After simulations of bulk water and crystal gold, a mixture system was
765     constructed, consisting of 1188 Au atoms and 1862 H$_2$O
766     molecules. Spohr potential was adopted in depicting the interaction
767     between metal atom and water molecule.\cite{ISI:000167766600035} A
768     similar protocol of equilibration was followed. Several thermal
769     gradients was built under different target thermal flux. It was found
770     out that compared to our previous simulation systems, the two phases
771     could have large temperature difference even under a relatively low
772     thermal flux.
773    
774    
775 skuang 3581 After simulations of homogeneous water and gold systems using
776     NIVS-RNEMD method were proved valid, calculation of gold/water
777     interfacial thermal conductivity was followed. It is found out that
778     the low interfacial conductance is probably due to the hydrophobic
779 skuang 3595 surface in our system. Figure \ref{interface} (a) demonstrates mass
780 skuang 3581 density change along $z$-axis, which is perpendicular to the
781     gold/water interface. It is observed that water density significantly
782     decreases when approaching the surface. Under this low thermal
783     conductance, both gold and water phase have sufficient time to
784     eliminate temperature difference inside respectively (Figure
785 skuang 3595 \ref{interface} b). With indistinguishable temperature difference
786 skuang 3581 within respective phase, it is valid to assume that the temperature
787     difference between gold and water on surface would be approximately
788     the same as the difference between the gold and water phase. This
789     assumption enables convenient calculation of $G$ using
790     Eq. \ref{interfaceCalc} instead of measuring temperatures of thin
791     layer of water and gold close enough to surface, which would have
792     greater fluctuation and lower accuracy. Reported results (Table
793     \ref{interfaceRes}) are of two orders of magnitude smaller than our
794     calculations on homogeneous systems, and thus have larger relative
795     errors than our calculation results on homogeneous systems.
796 skuang 3573
797 skuang 3571 \begin{figure}
798 skuang 3595 \includegraphics[width=\linewidth]{interface}
799     \caption{Simulation results for Gold/Water interfacial thermal
800     conductivity: (a) Significant water density decrease is observed on
801 skuang 3597 crystalline gold surface, which indicates low surface contact and
802     leads to low thermal conductance. (b) Temperature profiles for a
803     series of simulations. Temperatures of different slabs in the same
804     phase show no significant differences.}
805 skuang 3595 \label{interface}
806 skuang 3571 \end{figure}
807    
808 skuang 3572 \begin{table*}
809 gezelter 3612 \begin{minipage}{\linewidth}
810     \begin{center}
811    
812     \caption{Computed interfacial thermal conductivity ($G$) values
813     for the Au(111) / water interface at ${\langle T\rangle \sim}$
814     300K using a range of energy fluxes. Uncertainties are
815     indicated in parentheses. }
816    
817     \begin{tabular}{|cccc|}
818     \hline
819     $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
820     T_{water} \rangle$ (K) & $G$
821     (MW/m$^2$/K)\\
822     \hline
823     98.0 & 355.2 & 295.8 & 1.65(0.21) \\
824     78.8 & 343.8 & 298.0 & 1.72(0.32) \\
825     73.6 & 344.3 & 298.0 & 1.59(0.24) \\
826     49.2 & 330.1 & 300.4 & 1.65(0.35) \\
827     \hline
828     \end{tabular}
829     \label{interfaceRes}
830     \end{center}
831     \end{minipage}
832 skuang 3572 \end{table*}
833    
834 skuang 3576
835 skuang 3574 \section{Conclusions}
836     NIVS-RNEMD simulation method is developed and tested on various
837 skuang 3581 systems. Simulation results demonstrate its validity in thermal
838     conductivity calculations, from Lennard-Jones fluid to multi-atom
839     molecule like water and metal crystals. NIVS-RNEMD improves
840     non-Boltzmann-Maxwell distributions, which exist in previous RNEMD
841     methods. Furthermore, it develops a valid means for unphysical thermal
842     transfer between different species of molecules, and thus extends its
843     applicability to interfacial systems. Our calculation of gold/water
844     interfacial thermal conductivity demonstrates this advantage over
845     previous RNEMD methods. NIVS-RNEMD has also limited application on
846     shear viscosity calculations, but could cause temperature difference
847     among different dimensions under high momentum flux. Modification is
848     necessary to extend the applicability of NIVS-RNEMD in shear viscosity
849     calculations.
850 skuang 3572
851 gezelter 3524 \section{Acknowledgments}
852     Support for this project was provided by the National Science
853     Foundation under grant CHE-0848243. Computational time was provided by
854     the Center for Research Computing (CRC) at the University of Notre
855     Dame. \newpage
856    
857 gezelter 3583 \bibliographystyle{aip}
858 gezelter 3524 \bibliography{nivsRnemd}
859 gezelter 3583
860 gezelter 3524 \end{doublespace}
861     \end{document}
862