ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3621
Committed: Wed Aug 4 15:32:37 2010 UTC (13 years, 10 months ago) by gezelter
Content type: application/x-tex
File size: 40552 byte(s)
Log Message:
more edits

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{setspace}
5 \usepackage{endfloat}
6 \usepackage{caption}
7 %\usepackage{tabularx}
8 \usepackage{graphicx}
9 \usepackage{multirow}
10 %\usepackage{booktabs}
11 %\usepackage{bibentry}
12 %\usepackage{mathrsfs}
13 %\usepackage[ref]{overcite}
14 \usepackage[square, comma, sort&compress]{natbib}
15 \usepackage{url}
16 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18 9.0in \textwidth 6.5in \brokenpenalty=10000
19
20 % double space list of tables and figures
21 \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22 \setlength{\abovecaptionskip}{20 pt}
23 \setlength{\belowcaptionskip}{30 pt}
24
25 %\renewcommand\citemid{\ } % no comma in optional referenc note
26 \bibpunct{[}{]}{,}{s}{}{;}
27 \bibliographystyle{aip}
28
29 \begin{document}
30
31 \title{A gentler approach to RNEMD: Non-isotropic Velocity Scaling for computing Thermal Conductivity and Shear Viscosity}
32
33 \author{Shenyu Kuang and J. Daniel
34 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
35 Department of Chemistry and Biochemistry,\\
36 University of Notre Dame\\
37 Notre Dame, Indiana 46556}
38
39 \date{\today}
40
41 \maketitle
42
43 \begin{doublespace}
44
45 \begin{abstract}
46 We present a new method for introducing stable non-equilibrium
47 velocity and temperature gradients in molecular dynamics simulations
48 of heterogeneous systems. This method extends earlier Reverse
49 Non-Equilibrium Molecular Dynamics (RNEMD) methods which use
50 momentum exchange swapping moves. The standard swapping moves can
51 create non-thermal velocity distributions and are difficult to use
52 for interfacial calculations. By using non-isotropic velocity
53 scaling (NIVS) on the molecules in specific regions of a system, it
54 is possible to impose momentum or thermal flux between regions of a
55 simulation while conserving the linear momentum and total energy of
56 the system. To test the methods, we have computed the thermal
57 conductivity of model liquid and solid systems as well as the
58 interfacial thermal conductivity of a metal-water interface. We
59 find that the NIVS-RNEMD improves the problematic velocity
60 distributions that develop in other RNEMD methods.
61 \end{abstract}
62
63 \newpage
64
65 %\narrowtext
66
67 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68 % BODY OF TEXT
69 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70
71 \section{Introduction}
72 The original formulation of Reverse Non-equilibrium Molecular Dynamics
73 (RNEMD) obtains transport coefficients (thermal conductivity and shear
74 viscosity) in a fluid by imposing an artificial momentum flux between
75 two thin parallel slabs of material that are spatially separated in
76 the simulation cell.\cite{MullerPlathe:1997xw,ISI:000080382700030} The
77 artificial flux is typically created by periodically ``swapping''
78 either the entire momentum vector $\vec{p}$ or single components of
79 this vector ($p_x$) between molecules in each of the two slabs. If
80 the two slabs are separated along the $z$ coordinate, the imposed flux
81 is either directional ($j_z(p_x)$) or isotropic ($J_z$), and the
82 response of a simulated system to the imposed momentum flux will
83 typically be a velocity or thermal gradient (Fig. \ref{thermalDemo}).
84 The shear viscosity ($\eta$) and thermal conductivity ($\lambda$) are
85 easily obtained by assuming linear response of the system,
86 \begin{eqnarray}
87 j_z(p_x) & = & -\eta \frac{\partial v_x}{\partial z}\\
88 J_z & = & \lambda \frac{\partial T}{\partial z}
89 \end{eqnarray}
90 RNEMD has been widely used to provide computational estimates of
91 thermal conductivities and shear viscosities in a wide range of
92 materials, from liquid copper to both monatomic and molecular fluids
93 (e.g. ionic
94 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}
95
96 \begin{figure}
97 \includegraphics[width=\linewidth]{thermalDemo}
98 \caption{RNEMD methods impose an unphysical transfer of momentum or
99 kinetic energy between a ``hot'' slab and a ``cold'' slab in the
100 simulation box. The molecular system responds to this imposed flux
101 by generating a momentum or temperature gradient. The slope of the
102 gradient can then be used to compute transport properties (e.g.
103 shear viscosity and thermal conductivity).}
104 \label{thermalDemo}
105 \end{figure}
106
107 RNEMD is preferable in many ways to the forward NEMD
108 methods\cite{ISI:A1988Q205300014,hess:209,Vasquez:2004fk,backer:154503,ISI:000266247600008}
109 because it imposes what is typically difficult to measure (a flux or
110 stress) and it is typically much easier to compute the response
111 (momentum gradients or strains). For similar reasons, RNEMD is also
112 preferable to slowly-converging equilibrium methods for measuring
113 thermal conductivity and shear viscosity (using Green-Kubo
114 relations\cite{daivis:541,mondello:9327} or the Helfand moment
115 approach of Viscardy {\it et
116 al.}\cite{Viscardy:2007bh,Viscardy:2007lq}) because these rely on
117 computing difficult to measure quantities.
118
119 Another attractive feature of RNEMD is that it conserves both total
120 linear momentum and total energy during the swaps (as long as the two
121 molecules have the same identity), so the swapped configurations are
122 typically samples from the same manifold of states in the
123 microcanonical ensemble.
124
125 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered some
126 problems with the original RNEMD swap technique. Notably, large
127 momentum fluxes (equivalent to frequent momentum swaps between the
128 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
129 momentum distributions in the two slabs, as well as non-linear thermal
130 and velocity distributions along the direction of the imposed flux
131 ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
132 and proposed self-adjusting metrics for retaining the usability of the
133 method.
134
135 In this paper, we develop and test a method for non-isotropic velocity
136 scaling (NIVS) which retains the desirable features of RNEMD
137 (conservation of linear momentum and total energy, compatibility with
138 periodic boundary conditions) while establishing true thermal
139 distributions in each of the two slabs. In the next section, we
140 present the method for determining the scaling constraints. We then
141 test the method on both liquids and solids as well as a non-isotropic
142 liquid-solid interface and show that it is capable of providing
143 reasonable estimates of the thermal conductivity and shear viscosity
144 in all of these cases.
145
146 \section{Methodology}
147 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
148 periodic system is partitioned into a series of thin slabs along one
149 axis ($z$). One of the slabs at the end of the periodic box is
150 designated the ``hot'' slab, while the slab in the center of the box
151 is designated the ``cold'' slab. The artificial momentum flux will be
152 established by transferring momentum from the cold slab and into the
153 hot slab.
154
155 Rather than using momentum swaps, we use a series of velocity scaling
156 moves. For molecules $\{i\}$ located within the cold slab,
157 \begin{equation}
158 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
159 x & 0 & 0 \\
160 0 & y & 0 \\
161 0 & 0 & z \\
162 \end{array} \right) \cdot \vec{v}_i
163 \end{equation}
164 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
165 of the three directions in the system. Likewise, the molecules
166 $\{j\}$ located in the hot slab will see a concomitant scaling of
167 velocities,
168 \begin{equation}
169 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
170 x^\prime & 0 & 0 \\
171 0 & y^\prime & 0 \\
172 0 & 0 & z^\prime \\
173 \end{array} \right) \cdot \vec{v}_j
174 \end{equation}
175
176 Conservation of linear momentum in each of the three directions
177 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
178 parameters together:
179 \begin{equation}
180 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
181 \end{equation}
182 where
183 \begin{eqnarray}
184 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i v_{i\alpha} \\
185 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j v_{j\alpha}
186 \label{eq:momentumdef}
187 \end{eqnarray}
188 Therefore, for each of the three directions, the hot scaling
189 parameters are a simple function of the cold scaling parameters and
190 the instantaneous linear momenta in each of the two slabs.
191 \begin{equation}
192 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
193 \label{eq:hotcoldscaling}
194 \end{equation}
195 where
196 \begin{equation}
197 p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
198 \end{equation}
199 for convenience.
200
201 Conservation of total energy also places constraints on the scaling:
202 \begin{equation}
203 \sum_{\alpha = x,y,z} \left\{ K_h^\alpha + K_c^\alpha \right\} = \sum_{\alpha = x,y,z}
204 \left\{ \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha \right\}
205 \end{equation}
206 where the translational kinetic energies, $K_h^\alpha$ and
207 $K_c^\alpha$, are computed for each of the three directions in a
208 similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
209 Substituting in the expressions for the hot scaling parameters
210 ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
211 {\it constraint ellipsoid}:
212 \begin{equation}
213 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
214 c_\alpha \right) = 0
215 \label{eq:constraintEllipsoid}
216 \end{equation}
217 where the constants are obtained from the instantaneous values of the
218 linear momenta and kinetic energies for the hot and cold slabs,
219 \begin{eqnarray}
220 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
221 \left(p_\alpha\right)^2\right) \\
222 b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
223 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
224 \label{eq:constraintEllipsoidConsts}
225 \end{eqnarray}
226 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
227 cold slab scaling parameters which, when applied, preserve the linear
228 momentum of the system in all three directions as well as total
229 kinetic energy.
230
231 The goal of using these velocity scaling variables is to transfer
232 kinetic energy from the cold slab to the hot slab. If the hot and
233 cold slabs are separated along the z-axis, the energy flux is given
234 simply by the decrease in kinetic energy of the cold bin:
235 \begin{equation}
236 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
237 \end{equation}
238 The expression for the energy flux can be re-written as another
239 ellipsoid centered on $(x,y,z) = 0$:
240 \begin{equation}
241 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
242 K_c^\alpha -J_z \Delta t
243 \label{eq:fluxEllipsoid}
244 \end{equation}
245 The spatial extent of the {\it thermal flux ellipsoid} is governed
246 both by the target flux, $J_z$ as well as the instantaneous values of
247 the kinetic energy components in the cold bin.
248
249 To satisfy an energetic flux as well as the conservation constraints,
250 we must determine the points ${x,y,z}$ that lie on both the constraint
251 ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
252 (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
253 ellipsoids in 3-dimensional space.
254
255 \begin{figure}
256 \includegraphics[width=\linewidth]{ellipsoids}
257 \caption{Velocity scaling coefficients which maintain both constant
258 energy and constant linear momentum of the system lie on the surface
259 of the {\it constraint ellipsoid} while points which generate the
260 target momentum flux lie on the surface of the {\it flux ellipsoid}.
261 The velocity distributions in the cold bin are scaled by only those
262 points which lie on both ellipsoids.}
263 \label{ellipsoids}
264 \end{figure}
265
266 Since ellipsoids can be expressed as polynomials up to second order in
267 each of the three coordinates, finding the the intersection points of
268 two ellipsoids is isomorphic to finding the roots a polynomial of
269 degree 16. There are a number of polynomial root-finding methods in
270 the literature,\cite{Hoffman:2001sf,384119} but numerically finding
271 the roots of high-degree polynomials is generally an ill-conditioned
272 problem.\cite{Hoffman:2001sf} One simplification is to maintain
273 velocity scalings that are {\it as isotropic as possible}. To do
274 this, we impose $x=y$, and treat both the constraint and flux
275 ellipsoids as 2-dimensional ellipses. In reduced dimensionality, the
276 intersecting-ellipse problem reduces to finding the roots of
277 polynomials of degree 4.
278
279 Depending on the target flux and current velocity distributions, the
280 ellipsoids can have between 0 and 4 intersection points. If there are
281 no intersection points, it is not possible to satisfy the constraints
282 while performing a non-equilibrium scaling move, and no change is made
283 to the dynamics.
284
285 With multiple intersection points, any of the scaling points will
286 conserve the linear momentum and kinetic energy of the system and will
287 generate the correct target flux. Although this method is inherently
288 non-isotropic, the goal is still to maintain the system as close to an
289 isotropic fluid as possible. With this in mind, we would like the
290 kinetic energies in the three different directions could become as
291 close as each other as possible after each scaling. Simultaneously,
292 one would also like each scaling as gentle as possible, i.e. ${x,y,z
293 \rightarrow 1}$, in order to avoid large perturbation to the system.
294 To do this, we pick the intersection point which maintains the three
295 scaling variables ${x, y, z}$ as well as the ratio of kinetic energies
296 ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to 1.
297
298 After the valid scaling parameters are arrived at by solving geometric
299 intersection problems in $x, y, z$ space in order to obtain cold slab
300 scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
301 determine the conjugate hot slab scaling variables.
302
303 \subsection{Introducing shear stress via velocity scaling}
304 It is also possible to use this method to magnify the random
305 fluctuations of the average momentum in each of the bins to induce a
306 momentum flux. Doing this repeatedly will create a shear stress on
307 the system which will respond with an easily-measured strain. The
308 momentum flux (say along the $x$-direction) may be defined as:
309 \begin{equation}
310 (1-x) P_c^x = j_z(p_x)\Delta t
311 \label{eq:fluxPlane}
312 \end{equation}
313 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
314 its position governed both by a target value, $j_z(p_x)$ as well as
315 the instantaneous value of the momentum along the $x$-direction.
316
317 In order to satisfy a momentum flux as well as the conservation
318 constraints, we must determine the points ${x,y,z}$ which lie on both
319 the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
320 flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
321 ellipsoid and a plane in 3-dimensional space.
322
323 In the case of momentum flux transfer, we also impose another
324 constraint to set the kinetic energy transfer as zero. In other
325 words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
326 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
327 set of quartic equations to the above kinetic energy transfer problem.
328
329 \section{Computational Details}
330
331 We have implemented this methodology in our molecular dynamics code,
332 OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
333 with a variable frequency after the molecular dynamics (MD) steps. We
334 have tested the method in a variety of different systems, including:
335 homogeneous fluids (Lennard-Jones and SPC/E water), crystalline
336 solids, using both the embedded atom method
337 (EAM)~\cite{PhysRevB.33.7983} and quantum Sutton-Chen
338 (QSC)~\cite{PhysRevB.59.3527} models for Gold, and heterogeneous
339 interfaces (QSC gold - SPC/E water). The last of these systems would
340 have been difficult to study using previous RNEMD methods, but the
341 current method can easily provide estimates of the interfacial thermal
342 conductivity ($G$).
343
344 \subsection{Simulation Cells}
345
346 In each of the systems studied, the dynamics was carried out in a
347 rectangular simulation cell using periodic boundary conditions in all
348 three dimensions. The cells were longer along the $z$ axis and the
349 space was divided into $N$ slabs along this axis (typically $N=20$).
350 The top slab ($n=1$) was designated the ``hot'' slab, while the
351 central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
352 cases, simulations were first thermalized in canonical ensemble (NVT)
353 using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
354 microcanonical ensemble (NVE) before introducing any non-equilibrium
355 method.
356
357 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
358
359 In order to compare our new methodology with the original
360 M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
361 first performed simulations using the original technique. MORE HERE
362
363 \subsection{RNEMD with NIVS scaling}
364
365 For each simulation utilizing the swapping method, a corresponding
366 NIVS-RNEMD simulation was carried out using a target momentum flux set
367 to produce the same flux experienced in the swapping simulation.
368
369 To test the temperature homogeneity, directional momentum and
370 temperature distributions were accumulated for molecules in each of
371 the slabs. Transport coefficients were computed using the temperature
372 (and momentum) gradients across the $z$-axis as well as the total
373 momentum flux the system experienced during data collection portion of
374 the simulation.
375
376 \subsection{Shear viscosities}
377
378 The momentum flux was calculated using the total non-physical momentum
379 transferred (${P_x}$) and the data collection time ($t$):
380 \begin{equation}
381 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
382 \end{equation}
383 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
384 box. The factor of two in the denominator is present because physical
385 momentum transfer between the slabs occurs in two directions ($+z$ and
386 $-z$). The velocity gradient ${\langle \partial v_x /\partial z
387 \rangle}$ was obtained using linear regression of the mean $x$
388 component of the velocity, $\langle v_x \rangle$, in each of the bins.
389 For Lennard-Jones simulations, shear viscosities are reported in
390 reduced units (${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$).
391
392 \subsection{Thermal Conductivities}
393
394 The energy flux was calculated in a similar manner to the momentum
395 flux, using the total non-physical energy transferred (${E_{total}}$)
396 and the data collection time $t$:
397 \begin{equation}
398 J_z = \frac{E_{total}}{2 t L_x L_y}
399 \end{equation}
400 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
401 obtained by a linear regression of the temperature profile. For
402 Lennard-Jones simulations, thermal conductivities are reported in
403 reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
404 k_B^{-1}\varepsilon^{-1/2}}$).
405
406 \subsection{Interfacial Thermal Conductivities}
407
408 For interfaces with a relatively low interfacial conductance, the bulk
409 regions on either side of an interface rapidly come to a state in
410 which the two phases have relatively homogeneous (but distinct)
411 temperatures. The interfacial thermal conductivity $G$ can therefore
412 be approximated as:
413
414 \begin{equation}
415 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_{gold}\rangle -
416 \langle T_{water}\rangle \right)}
417 \label{interfaceCalc}
418 \end{equation}
419 where ${E_{total}}$ is the imposed non-physical kinetic energy
420 transfer and ${\langle T_{gold}\rangle}$ and ${\langle
421 T_{water}\rangle}$ are the average observed temperature of gold and
422 water phases respectively. If the interfacial conductance is {\it
423 not} small, it is also be possible to compute the interfacial
424 thermal conductivity using this method utilizing the change in the
425 slope of the thermal gradient ($\partial^2 \langle T \rangle / \partial
426 z^2$) at the interface.
427
428 \section{Results}
429
430 \subsection{Lennard-Jones Fluid}
431 2592 Lennard-Jones atoms were placed in an orthorhombic cell
432 ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
433 reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
434 direct comparison between our results and previous methods. These
435 simulations were carried out with a reduced timestep ${\tau^* =
436 4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
437 temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
438 conductivity calculations, simulations were run under reduced
439 temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
440 ensemble. The simulations included $10^5$ steps of equilibration
441 without any momentum flux, $10^5$ steps of stablization with an
442 imposed momentum transfer to create a gradient, and $10^6$ steps of
443 data collection under RNEMD.
444
445 \subsubsection*{Thermal Conductivity}
446
447 Our thermal conductivity calculations show that the NIVS method agrees
448 well with the swapping method. Five different swap intervals were
449 tested (Table \ref{LJ}). Similar thermal gradients were observed with
450 similar thermal flux under the two different methods (Figure
451 \ref{thermalGrad}). Furthermore, with appropriate choice of scaling
452 variables, the temperatures along $x$, $y$ and $z$ axes showed
453 observable difference. WHAT DOES THIS MEAN? (Figure TO BE ADDED). The
454 system is able to maintain temperature homogeneity even under high
455 flux.
456
457 \begin{table*}
458 \begin{minipage}{\linewidth}
459 \begin{center}
460
461 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
462 ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
463 ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
464 at various momentum fluxes. The original swapping method and
465 the velocity scaling method give similar results.
466 Uncertainties are indicated in parentheses.}
467
468 \begin{tabular}{|cc|cc|cc|}
469 \hline
470 \multicolumn{2}{|c}{Momentum Exchange} &
471 \multicolumn{2}{|c}{Swapping RNEMD} &
472 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
473 \hline
474 \multirow{2}{*}{Swap Interval (fs)} & Equivalent $J_z^*$ or &
475 \multirow{2}{*}{$\lambda^*_{swap}$} &
476 \multirow{2}{*}{$\eta^*_{swap}$} &
477 \multirow{2}{*}{$\lambda^*_{scale}$} &
478 \multirow{2}{*}{$\eta^*_{scale}$} \\
479 & $j_z^*(p_x)$ (reduced units) & & & & \\
480 \hline
481 250 & 0.16 & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
482 500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
483 1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
484 2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
485 2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
486 \hline
487 \end{tabular}
488 \label{LJ}
489 \end{center}
490 \end{minipage}
491 \end{table*}
492
493 \begin{figure}
494 \includegraphics[width=\linewidth]{thermalGrad}
495 \caption{The NIVS-RNEMD method (b) creates similar temperature gradients
496 compared with the swapping method (a) under a variety of imposed
497 kinetic energy flux values. Furthermore, the implementation of
498 Non-Isotropic Velocity Scaling does not cause temperature
499 differences among the three dimensions (c).}
500 \label{thermalGrad}
501 \end{figure}
502
503 \subsubsection*{Velocity Distributions}
504
505 During these simulations, velocities were recorded every 1000 steps
506 and were used to produce distributions of both velocity and speed in
507 each of the slabs. From these distributions, we observed that under
508 relatively high non-physical kinetic energy flux, the speed of
509 molecules in slabs where swapping occured could deviate from the
510 Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
511 and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
512 distributions deviate from an ideal distribution. In the ``hot'' slab,
513 the probability density is notched at low speeds and has a substantial
514 shoulder at higher speeds relative to the ideal MB distribution. In
515 the cold slab, the opposite notching and shouldering occurs. This
516 phenomenon is more obvious at higher swapping rates.
517
518 The peak of the velocity distribution is substantially flattened in
519 the hot slab, and is overly sharp (with truncated wings) in the cold
520 slab. This problem is rooted in the mechanism of the swapping method.
521 Continually depleting low (high) speed particles in the high (low)
522 temperature slab is not complemented by diffusions of low (high) speed
523 particles from neighboring slabs, unless the swapping rate is
524 sufficiently small. Simutaneously, surplus low speed particles in the
525 low temperature slab do not have sufficient time to diffuse to
526 neighboring slabs. Since the thermal exchange rate must reach a
527 minimum level to produce an observable thermal gradient, the
528 swapping-method RNEMD has a relatively narrow choice of exchange times
529 that can be utilized.
530
531 For comparison, NIVS-RNEMD produces a speed distribution closer to the
532 Maxwell-Boltzmann curve (Figure \ref{thermalHist}). The reason for
533 this is simple; upon velocity scaling, a Gaussian distribution remains
534 Gaussian. Although a single scaling move is non-isotropic in three
535 dimensions, our criteria for choosing a set of scaling coefficients
536 helps maintain the distributions as close to isotropic as possible.
537 Consequently, NIVS-RNEMD is able to impose an unphysical thermal flux
538 as the previous RNEMD methods but without large perturbations to the
539 velocity distributions in the two slabs.
540
541 \begin{figure}
542 \includegraphics[width=\linewidth]{thermalHist}
543 \caption{Speed distribution for thermal conductivity using
544 ``swapping'' and NIVS-RNEMD methods. Shown is from simulations under
545 ${\langle T^* \rangle = 0.8}$ with an swapping interval of 200
546 time steps (equivalent ${J_z^*\sim 0.2}$). In circled areas,
547 distributions from ``swapping'' RNEMD simulation have deviations
548 from ideal Maxwell-Boltzmann distributions.}
549 \label{thermalHist}
550 \end{figure}
551
552
553 \subsubsection*{Shear Viscosity}
554 Our calculations (Table \ref{LJ}) show that velocity-scaling RNEMD
555 predicted comparable shear viscosities to swap RNEMD method. The
556 average molecular momentum gradients of these samples are shown in
557 Figure \ref{shear} (a) and (b).
558
559 \begin{figure}
560 \includegraphics[width=\linewidth]{shear}
561 \caption{Average momentum gradients in shear viscosity simulations,
562 using (a) ``swapping'' method and (b) NIVS-RNEMD method
563 respectively. (c) Temperature difference among $x$ and $y, z$
564 dimensions observed when using NIVS-RNEMD with ${j_z^*(p_x)\sim 0.09}$.}
565 \label{shear}
566 \end{figure}
567
568 Observations of the three one-dimensional temperatures in each of the
569 slabs shows that NIVS-RNEMD does produce some thermal anisotropy,
570 particularly in the hot and cold slabs. Figure \ref{shear} (c)
571 indicates that with a relatively large imposed momentum flux,
572 $j_z(p_x)$, the $x$ direction approaches a different temperature from
573 the $y$ and $z$ directions in both the hot and cold bins. This is an
574 artifact of the scaling constraints. After the momentum gradient has
575 been established, $P_c^x < 0$. Consequently, the scaling factor $x$
576 is nearly always greater than one in order to satisfy the constraints.
577 This will continually increase the kinetic energy in $x$-dimension,
578 $K_c^x$. If there is not enough time for the kinetic energy to
579 exchange among different directions and different slabs, the system
580 will exhibit the observed thermal anisotropy in the hot and cold bins.
581
582 Although results between scaling and swapping methods are comparable,
583 the inherent temperature anisotropy does make NIVS-RNEMD method less
584 attractive than swapping RNEMD for shear viscosity calculations. We
585 note that this problem appears only when momentum flux is applied, and
586 does not appear in thermal flux calculations.
587
588 \subsection{Bulk SPC/E water}
589
590 We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
591 to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
592 the original swapping RNEMD method. Bedrov {\it et
593 al.}\cite{Bedrov:2000} argued that exchange of the molecule
594 center-of-mass velocities instead of single atom velocities in a
595 molecule conserves the total kinetic energy and linear momentum. This
596 principle is also adopted Fin our simulations. Scaling was applied to
597 the center-of-mass velocities of the rigid bodies of SPC/E model water
598 molecules.
599
600 To construct the simulations, a simulation box consisting of 1000
601 molecules were first equilibrated under ambient pressure and
602 temperature conditions using the isobaric-isothermal (NPT)
603 ensemble.\cite{melchionna93} A fixed volume was chosen to match the
604 average volume observed in the NPT simulations, and this was followed
605 by equilibration, first in the canonical (NVT) ensemble, followed by a
606 100~ps period under constant-NVE conditions without any momentum flux.
607 Another 100~ps was allowed to stabilize the system with an imposed
608 momentum transfer to create a gradient, and 1~ns was allotted for data
609 collection under RNEMD.
610
611 In our simulations, the established temperature gradients were similar
612 to the previous work. Our simulation results at 318K are in good
613 agreement with those from Bedrov {\it et al.} (Table
614 \ref{spceThermal}). And both methods yield values in reasonable
615 agreement with experimental values.
616
617 \begin{figure}
618 \includegraphics[width=\linewidth]{spceGrad}
619 \caption{Temperature gradients in SPC/E water thermal conductivity
620 simulations.}
621 \label{spceGrad}
622 \end{figure}
623
624 \begin{table*}
625 \begin{minipage}{\linewidth}
626 \begin{center}
627
628 \caption{Thermal conductivity of SPC/E water under various
629 imposed thermal gradients. Uncertainties are indicated in
630 parentheses.}
631
632 \begin{tabular}{|c|c|ccc|}
633 \hline
634 \multirow{2}{*}{$\langle T\rangle$(K)} &
635 \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
636 \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
637 \mathrm{K}^{-1})$} \\
638 & & This work & Previous simulations\cite{Bedrov:2000} &
639 Experiment\cite{WagnerKruse}\\
640 \hline
641 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
642 \multirow{3}{*}{0.61}\\
643 & 0.81 & 0.770(0.008) & & \\
644 & 1.54 & 0.813(0.007) & & \\
645 \hline
646 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
647 \multirow{2}{*}{0.64}\\
648 & 1.59 & 0.778(0.019) & 0.730 & \\
649 \hline
650 \end{tabular}
651 \label{spceThermal}
652 \end{center}
653 \end{minipage}
654 \end{table*}
655
656 \subsection{Crystalline Gold}
657
658 To see how the method performed in a solid, we calculated thermal
659 conductivities using two atomistic models for gold. Several different
660 potential models have been developed that reasonably describe
661 interactions in transition metals. In particular, the Embedded Atom
662 Model (EAM)~\cite{PhysRevB.33.7983} and Sutton-Chen (SC)~\cite{Chen90}
663 potential have been used to study a wide range of phenomena in both
664 bulk materials and
665 nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
666 Both potentials are based on a model of a metal which treats the
667 nuclei and core electrons as pseudo-atoms embedded in the electron
668 density due to the valence electrons on all of the other atoms in the
669 system. The SC potential has a simple form that closely resembles the
670 Lennard Jones potential,
671 \begin{equation}
672 \label{eq:SCP1}
673 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] ,
674 \end{equation}
675 where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
676 \begin{equation}
677 \label{eq:SCP2}
678 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}}.
679 \end{equation}
680 $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
681 interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
682 Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
683 the interactions between the valence electrons and the cores of the
684 pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
685 scale, $c_i$ scales the attractive portion of the potential relative
686 to the repulsive interaction and $\alpha_{ij}$ is a length parameter
687 that assures a dimensionless form for $\rho$. These parameters are
688 tuned to various experimental properties such as the density, cohesive
689 energy, and elastic moduli for FCC transition metals. The quantum
690 Sutton-Chen (QSC) formulation matches these properties while including
691 zero-point quantum corrections for different transition
692 metals.\cite{PhysRevB.59.3527} The EAM functional forms differ
693 slightly from SC but the overall method is very similar.
694
695 In this work, we have utilized both the EAM and the QSC potentials to
696 test the behavior of scaling RNEMD.
697
698 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
699 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
700 both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
701 pre-equilibration at a target pressure of 1 atm. When equilibrated
702 under NPT conditions, our simulation box expanded by approximately 1\%
703 in volume. Following adjustment of the box volume, equilibrations in
704 both the canonical and microcanonical ensembles were carried out. With
705 the simulation cell divided evenly into 10 slabs, different thermal
706 gradients were established by applying a set of target thermal
707 transfer fluxes.
708
709 The results for the thermal conductivity of gold are shown in Table
710 \ref{AuThermal}. In these calculations, the end and middle slabs were
711 excluded in thermal gradient linear regession. EAM predicts slightly
712 larger thermal conductivities than QSC. However, both values are
713 smaller than experimental value by a factor of more than 200. This
714 behavior has been observed previously by Richardson and Clancy, and
715 has been attributed to the lack of electronic contribution in these
716 force fields.\cite{Clancy:1992} It should be noted that the density of
717 the metal being simulated has an effect on thermal conductance. With
718 an expanded lattice, lower thermal conductance is expected (and
719 observed). We also observed a decrease in thermal conductance at
720 higher temperatures, a trend that agrees with experimental
721 measurements.\cite{AshcroftMermin}
722
723 \begin{table*}
724 \begin{minipage}{\linewidth}
725 \begin{center}
726
727 \caption{Calculated thermal conductivity of crystalline gold
728 using two related force fields. Calculations were done at both
729 experimental and equilibrated densities and at a range of
730 temperatures and thermal flux rates. Uncertainties are
731 indicated in parentheses. Richardson {\it et
732 al.}\cite{Clancy:1992} give an estimate of 1.74 $\mathrm{W
733 m}^{-1}\mathrm{K}^{-1}$ for EAM gold
734 at a density of 19.263 g / cm$^3$.}
735
736 \begin{tabular}{|c|c|c|cc|}
737 \hline
738 Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
739 $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
740 \hline
741 \multirow{7}{*}{QSC} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
742 & & & 2.86 & 1.08(0.05)\\
743 & & & 5.14 & 1.15(0.07)\\\cline{2-5}
744 & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
745 & & & 3.02 & 1.26(0.05)\\\cline{3-5}
746 & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
747 & & & 4.84 & 0.92(0.05)\\
748 \hline
749 \multirow{8}{*}{EAM} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
750 & & & 2.06 & 1.37(0.04)\\
751 & & & 2.55 & 1.41(0.07)\\\cline{2-5}
752 & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
753 & & & 2.04 & 1.41(0.07)\\
754 & & & 2.41 & 1.53(0.10)\\\cline{3-5}
755 & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
756 & & & 4.14 & 1.08(0.05)\\
757 \hline
758 \end{tabular}
759 \label{AuThermal}
760 \end{center}
761 \end{minipage}
762 \end{table*}
763
764 \subsection{Thermal Conductance at the Au/H$_2$O interface}
765 The most attractive aspect of the scaling approach for RNEMD is the
766 ability to use the method in non-homogeneous systems, where molecules
767 of different identities are segregated in different slabs. To test
768 this application, we simulated a Gold (111) / water interface. To
769 construct the interface, a box containing a lattice of 1188 Au atoms
770 (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
771 relax under ambient temperature and pressure. A separate (but
772 identically sized) box of SPC/E water was also equilibrated at ambient
773 conditions. The two boxes were combined by removing all water
774 molecules within 3 \AA radius of any gold atom. The final
775 configuration contained 1862 SPC/E water molecules.
776
777 The Spohr potential was adopted in depicting the interaction between
778 metal atoms and water molecules.\cite{ISI:000167766600035} A similar
779 protocol of equilibration to our water simulations was followed. We
780 observed that the two phases developed large temperature differences
781 even under a relatively low thermal flux.
782
783 The low interfacial conductance is probably due to the hydrophobic
784 surface in our system. Figure \ref{interface} (a) demonstrates mass
785 density change along $z$-axis, which is perpendicular to the
786 gold/water interface. It is observed that water density significantly
787 decreases when approaching the surface. Under this low thermal
788 conductance, both gold and water phase have sufficient time to
789 eliminate temperature difference inside respectively (Figure
790 \ref{interface} b). With indistinguishable temperature difference
791 within respective phase, it is valid to assume that the temperature
792 difference between gold and water on surface would be approximately
793 the same as the difference between the gold and water phase. This
794 assumption enables convenient calculation of $G$ using Eq.
795 \ref{interfaceCalc} instead of measuring temperatures of thin layer of
796 water and gold close enough to surface, which would have greater
797 fluctuation and lower accuracy. Reported results (Table
798 \ref{interfaceRes}) are of two orders of magnitude smaller than our
799 calculations on homogeneous systems, and thus have larger relative
800 errors than our calculation results on homogeneous systems.
801
802 \begin{figure}
803 \includegraphics[width=\linewidth]{interface}
804 \caption{Simulation results for Gold/Water interfacial thermal
805 conductivity: (a) Significant water density decrease is observed on
806 crystalline gold surface, which indicates low surface contact and
807 leads to low thermal conductance. (b) Temperature profiles for a
808 series of simulations. Temperatures of different slabs in the same
809 phase show no significant differences.}
810 \label{interface}
811 \end{figure}
812
813 \begin{table*}
814 \begin{minipage}{\linewidth}
815 \begin{center}
816
817 \caption{Computed interfacial thermal conductivity ($G$) values
818 for the Au(111) / water interface at ${\langle T\rangle \sim}$
819 300K using a range of energy fluxes. Uncertainties are
820 indicated in parentheses. }
821
822 \begin{tabular}{|cccc| }
823 \hline
824 $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
825 T_{water} \rangle$ (K) & $G$
826 (MW/m$^2$/K)\\
827 \hline
828 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
829 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
830 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
831 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
832 \hline
833 \end{tabular}
834 \label{interfaceRes}
835 \end{center}
836 \end{minipage}
837 \end{table*}
838
839
840 \section{Conclusions}
841 NIVS-RNEMD simulation method is developed and tested on various
842 systems. Simulation results demonstrate its validity in thermal
843 conductivity calculations, from Lennard-Jones fluid to multi-atom
844 molecule like water and metal crystals. NIVS-RNEMD improves
845 non-Boltzmann-Maxwell distributions, which exist inb previous RNEMD
846 methods. Furthermore, it develops a valid means for unphysical thermal
847 transfer between different species of molecules, and thus extends its
848 applicability to interfacial systems. Our calculation of gold/water
849 interfacial thermal conductivity demonstrates this advantage over
850 previous RNEMD methods. NIVS-RNEMD has also limited application on
851 shear viscosity calculations, but could cause temperature difference
852 among different dimensions under high momentum flux. Modification is
853 necessary to extend the applicability of NIVS-RNEMD in shear viscosity
854 calculations.
855
856 \section{Acknowledgments}
857 Support for this project was provided by the National Science
858 Foundation under grant CHE-0848243. Computational time was provided by
859 the Center for Research Computing (CRC) at the University of Notre
860 Dame. \newpage
861
862 \bibliography{nivsRnemd}
863
864 \end{doublespace}
865 \end{document}
866