ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nivsRnemd/nivsRnemd.tex
Revision: 3645
Committed: Wed Sep 15 23:22:47 2010 UTC (13 years, 11 months ago) by skuang
Content type: application/x-tex
File size: 42860 byte(s)
Log Message:
some changes according to reviewers' comments.

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 method, 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 and integrating long-time correlation functions that are
118 subject to noise issues.
119
120 Another attractive feature of RNEMD is that it conserves both total
121 linear momentum and total energy during the swaps (as long as the two
122 molecules have the same identity), so the swapped configurations are
123 typically samples from the same manifold of states in the
124 microcanonical ensemble. Furthermore, the method is applicable with
125 different ensembles, unlike the heat-exchange algorithm proposed by
126 Hafskjold {\it et al.} \cite{HeX:1994,HeX:1993}, which is incompatible
127 with non-microcanonical ensemble.
128
129 Recently, Tenney and Maginn\cite{Maginn:2010} have discovered some
130 problems with the original RNEMD swap technique. Notably, large
131 momentum fluxes (equivalent to frequent momentum swaps between the
132 slabs) can result in ``notched'', ``peaked'' and generally non-thermal
133 momentum distributions in the two slabs, as well as non-linear thermal
134 and velocity distributions along the direction of the imposed flux
135 ($z$). Tenney and Maginn obtained reasonable limits on imposed flux
136 and proposed self-adjusting metrics for retaining the usability of the
137 method.
138
139 In this paper, we develop and test a method for non-isotropic velocity
140 scaling (NIVS) which retains the desirable features of RNEMD
141 (conservation of linear momentum and total energy, compatibility with
142 periodic boundary conditions) while establishing true thermal
143 distributions in each of the two slabs. In the next section, we
144 present the method for determining the scaling constraints. We then
145 test the method on both liquids and solids as well as a non-isotropic
146 liquid-solid interface and show that it is capable of providing
147 reasonable estimates of the thermal conductivity and shear viscosity
148 in all of these cases.
149
150 \section{Methodology}
151 We retain the basic idea of M\"{u}ller-Plathe's RNEMD method; the
152 periodic system is partitioned into a series of thin slabs along one
153 axis ($z$). One of the slabs at the end of the periodic box is
154 designated the ``hot'' slab, while the slab in the center of the box
155 is designated the ``cold'' slab. The artificial momentum flux will be
156 established by transferring momentum from the cold slab and into the
157 hot slab.
158
159 Rather than using momentum swaps, we use a series of velocity scaling
160 moves. For molecules $\{i\}$ located within the cold slab,
161 \begin{equation}
162 \vec{v}_i \leftarrow \left( \begin{array}{ccc}
163 x & 0 & 0 \\
164 0 & y & 0 \\
165 0 & 0 & z \\
166 \end{array} \right) \cdot \vec{v}_i
167 \end{equation}
168 where ${x, y, z}$ are a set of 3 velocity-scaling variables for each
169 of the three directions in the system. Likewise, the molecules
170 $\{j\}$ located in the hot slab will see a concomitant scaling of
171 velocities,
172 \begin{equation}
173 \vec{v}_j \leftarrow \left( \begin{array}{ccc}
174 x^\prime & 0 & 0 \\
175 0 & y^\prime & 0 \\
176 0 & 0 & z^\prime \\
177 \end{array} \right) \cdot \vec{v}_j
178 \end{equation}
179
180 Conservation of linear momentum in each of the three directions
181 ($\alpha = x,y,z$) ties the values of the hot and cold scaling
182 parameters together:
183 \begin{equation}
184 P_h^\alpha + P_c^\alpha = \alpha^\prime P_h^\alpha + \alpha P_c^\alpha
185 \end{equation}
186 where
187 \begin{eqnarray}
188 P_c^\alpha & = & \sum_{i = 1}^{N_c} m_i v_{i\alpha} \\
189 P_h^\alpha & = & \sum_{j = 1}^{N_h} m_j v_{j\alpha}
190 \label{eq:momentumdef}
191 \end{eqnarray}
192 Therefore, for each of the three directions, the hot scaling
193 parameters are a simple function of the cold scaling parameters and
194 the instantaneous linear momenta in each of the two slabs.
195 \begin{equation}
196 \alpha^\prime = 1 + (1 - \alpha) p_\alpha
197 \label{eq:hotcoldscaling}
198 \end{equation}
199 where
200 \begin{equation}
201 p_\alpha = \frac{P_c^\alpha}{P_h^\alpha}
202 \end{equation}
203 for convenience.
204
205 Conservation of total energy also places constraints on the scaling:
206 \begin{equation}
207 \sum_{\alpha = x,y,z} \left\{ K_h^\alpha + K_c^\alpha \right\} = \sum_{\alpha = x,y,z}
208 \left\{ \left(\alpha^\prime\right)^2 K_h^\alpha + \alpha^2 K_c^\alpha \right\}
209 \end{equation}
210 where the translational kinetic energies, $K_h^\alpha$ and
211 $K_c^\alpha$, are computed for each of the three directions in a
212 similar manner to the linear momenta (Eq. \ref{eq:momentumdef}).
213 Substituting in the expressions for the hot scaling parameters
214 ($\alpha^\prime$) from Eq. (\ref{eq:hotcoldscaling}), we obtain the
215 {\it constraint ellipsoid}:
216 \begin{equation}
217 \sum_{\alpha = x,y,z} \left( a_\alpha \alpha^2 + b_\alpha \alpha +
218 c_\alpha \right) = 0
219 \label{eq:constraintEllipsoid}
220 \end{equation}
221 where the constants are obtained from the instantaneous values of the
222 linear momenta and kinetic energies for the hot and cold slabs,
223 \begin{eqnarray}
224 a_\alpha & = &\left(K_c^\alpha + K_h^\alpha
225 \left(p_\alpha\right)^2\right) \\
226 b_\alpha & = & -2 K_h^\alpha p_\alpha \left( 1 + p_\alpha\right) \\
227 c_\alpha & = & K_h^\alpha p_\alpha^2 + 2 K_h^\alpha p_\alpha - K_c^\alpha
228 \label{eq:constraintEllipsoidConsts}
229 \end{eqnarray}
230 This ellipsoid (Eq. \ref{eq:constraintEllipsoid}) defines the set of
231 cold slab scaling parameters which, when applied, preserve the linear
232 momentum of the system in all three directions as well as total
233 kinetic energy.
234
235 The goal of using these velocity scaling variables is to transfer
236 kinetic energy from the cold slab to the hot slab. If the hot and
237 cold slabs are separated along the z-axis, the energy flux is given
238 simply by the decrease in kinetic energy of the cold bin:
239 \begin{equation}
240 (1-x^2) K_c^x + (1-y^2) K_c^y + (1-z^2) K_c^z = J_z\Delta t
241 \end{equation}
242 The expression for the energy flux can be re-written as another
243 ellipsoid centered on $(x,y,z) = 0$:
244 \begin{equation}
245 \sum_{\alpha = x,y,z} K_c^\alpha \alpha^2 = \sum_{\alpha = x,y,z}
246 K_c^\alpha -J_z \Delta t
247 \label{eq:fluxEllipsoid}
248 \end{equation}
249 The spatial extent of the {\it thermal flux ellipsoid} is governed
250 both by the target flux, $J_z$ as well as the instantaneous values of
251 the kinetic energy components in the cold bin.
252
253 To satisfy an energetic flux as well as the conservation constraints,
254 we must determine the points ${x,y,z}$ that lie on both the constraint
255 ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the flux ellipsoid
256 (Eq. \ref{eq:fluxEllipsoid}), i.e. the intersection of the two
257 ellipsoids in 3-dimensional space.
258
259 \begin{figure}
260 \includegraphics[width=\linewidth]{ellipsoids}
261 \caption{Velocity scaling coefficients which maintain both constant
262 energy and constant linear momentum of the system lie on the surface
263 of the {\it constraint ellipsoid} while points which generate the
264 target momentum flux lie on the surface of the {\it flux ellipsoid}.
265 The velocity distributions in the cold bin are scaled by only those
266 points which lie on both ellipsoids.}
267 \label{ellipsoids}
268 \end{figure}
269
270 Since ellipsoids can be expressed as polynomials up to second order in
271 each of the three coordinates, finding the the intersection points of
272 two ellipsoids is isomorphic to finding the roots a polynomial of
273 degree 16. There are a number of polynomial root-finding methods in
274 the literature,\cite{Hoffman:2001sf,384119} but numerically finding
275 the roots of high-degree polynomials is generally an ill-conditioned
276 problem.\cite{Hoffman:2001sf} One simplification is to maintain
277 velocity scalings that are {\it as isotropic as possible}. To do
278 this, we impose $x=y$, and treat both the constraint and flux
279 ellipsoids as 2-dimensional ellipses. In reduced dimensionality, the
280 intersecting-ellipse problem reduces to finding the roots of
281 polynomials of degree 4.
282
283 Depending on the target flux and current velocity distributions, the
284 ellipsoids can have between 0 and 4 intersection points. If there are
285 no intersection points, it is not possible to satisfy the constraints
286 while performing a non-equilibrium scaling move, and no change is made
287 to the dynamics.
288
289 With multiple intersection points, any of the scaling points will
290 conserve the linear momentum and kinetic energy of the system and will
291 generate the correct target flux. Although this method is inherently
292 non-isotropic, the goal is still to maintain the system as close to an
293 isotropic fluid as possible. With this in mind, we would like the
294 kinetic energies in the three different directions could become as
295 close as each other as possible after each scaling. Simultaneously,
296 one would also like each scaling to be as gentle as possible, i.e.
297 ${x,y,z \rightarrow 1}$, in order to avoid large perturbations to the
298 system. To do this, we pick the intersection point which maintains
299 the three scaling variables ${x, y, z}$ as well as the ratio of
300 kinetic energies ${K_c^z/K_c^x, K_c^z/K_c^y}$ as close as possible to
301 1.
302
303 After the valid scaling parameters are arrived at by solving geometric
304 intersection problems in $x, y, z$ space in order to obtain cold slab
305 scaling parameters, Eq. (\ref{eq:hotcoldscaling}) is used to
306 determine the conjugate hot slab scaling variables.
307
308 \subsection{Introducing shear stress via velocity scaling}
309 It is also possible to use this method to magnify the random
310 fluctuations of the average momentum in each of the bins to induce a
311 momentum flux. Doing this repeatedly will create a shear stress on
312 the system which will respond with an easily-measured strain. The
313 momentum flux (say along the $x$-direction) may be defined as:
314 \begin{equation}
315 (1-x) P_c^x = j_z(p_x)\Delta t
316 \label{eq:fluxPlane}
317 \end{equation}
318 This {\it momentum flux plane} is perpendicular to the $x$-axis, with
319 its position governed both by a target value, $j_z(p_x)$ as well as
320 the instantaneous value of the momentum along the $x$-direction.
321
322 In order to satisfy a momentum flux as well as the conservation
323 constraints, we must determine the points ${x,y,z}$ which lie on both
324 the constraint ellipsoid (Eq. \ref{eq:constraintEllipsoid}) and the
325 flux plane (Eq. \ref{eq:fluxPlane}), i.e. the intersection of an
326 ellipsoid and a plane in 3-dimensional space.
327
328 In the case of momentum flux transfer, we also impose another
329 constraint to set the kinetic energy transfer as zero. In other
330 words, we apply Eq. \ref{eq:fluxEllipsoid} and let ${J_z = 0}$. With
331 one variable fixed by Eq. \ref{eq:fluxPlane}, one now have a similar
332 set of quartic equations to the above kinetic energy transfer problem.
333
334 \section{Computational Details}
335
336 We have implemented this methodology in our molecular dynamics code,
337 OpenMD,\cite{Meineke:2005gd,openmd} performing the NIVS scaling moves
338 with a variable frequency after the molecular dynamics (MD) steps. We
339 have tested the method in a variety of different systems, including:
340 homogeneous fluids (Lennard-Jones and SPC/E water), crystalline
341 solids, using both the embedded atom method
342 (EAM)~\cite{PhysRevB.33.7983} and quantum Sutton-Chen
343 (QSC)~\cite{PhysRevB.59.3527} models for Gold, and heterogeneous
344 interfaces (QSC gold - SPC/E water). Even though previous RNEMD
345 methods might remain usable for the last of these systems, energy
346 transfer from imaginary elastic collisions would be less effective
347 when the two particles involved have larger mass difference, and thus
348 affect the actuall implementation of these methods. However, our
349 current method can easily provide estimates of the interfacial thermal
350 conductivity ($G$).
351
352 \subsection{Simulation Cells}
353
354 In each of the systems studied, the dynamics was carried out in a
355 rectangular simulation cell using periodic boundary conditions in all
356 three dimensions. The cells were longer along the $z$ axis and the
357 space was divided into $N$ slabs along this axis (typically $N=20$).
358 The top slab ($n=1$) was designated the ``hot'' slab, while the
359 central slab ($n= N/2 + 1$) was designated the ``cold'' slab. In all
360 cases, simulations were first thermalized in canonical ensemble (NVT)
361 using a Nos\'{e}-Hoover thermostat.\cite{Hoover85} then equilibrated in
362 microcanonical ensemble (NVE) before introducing any non-equilibrium
363 method.
364
365 \subsection{RNEMD with M\"{u}ller-Plathe swaps}
366
367 In order to compare our new methodology with the original
368 M\"{u}ller-Plathe swapping algorithm,\cite{ISI:000080382700030} we
369 first performed simulations using the original technique. At fixed
370 intervals, kinetic energy or momentum exchange moves were performed
371 between the hot and the cold slabs. The interval between exchange
372 moves governs the effective momentum flux ($j_z(p_x)$) or energy flux
373 ($J_z$) between the two slabs so to vary this quantity, we performed
374 simulations with a variety of delay intervals between the swapping moves.
375
376 For thermal conductivity measurements, the particle with smallest
377 speed in the hot slab and the one with largest speed in the cold slab
378 had their entire momentum vectors swapped. In the test cases run
379 here, all particles had the same chemical identity and mass, so this
380 move preserves both total linear momentum and total energy. It is
381 also possible to exchange energy by assuming an elastic collision
382 between the two particles which are exchanging energy.
383
384 For shear stress simulations, the particle with the most negative
385 $p_x$ in the hot slab and the one with the most positive $p_x$ in the
386 cold slab exchanged only this component of their momentum vectors.
387
388 \subsection{RNEMD with NIVS scaling}
389
390 For each simulation utilizing the swapping method, a corresponding
391 NIVS-RNEMD simulation was carried out using a target momentum flux set
392 to produce the same flux experienced in the swapping simulation.
393
394 To test the temperature homogeneity, momentum and temperature
395 distributions (for all three dimensions) were accumulated for
396 molecules in each of the slabs. Transport coefficients were computed
397 using the temperature (and momentum) gradients across the $z$-axis as
398 well as the total momentum flux the system experienced during data
399 collection portion of the simulation.
400
401 \subsection{Shear viscosities}
402
403 The momentum flux was calculated using the total non-physical momentum
404 transferred (${P_x}$) and the data collection time ($t$):
405 \begin{equation}
406 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
407 \end{equation}
408 where $L_x$ and $L_y$ denote the $x$ and $y$ lengths of the simulation
409 box. The factor of two in the denominator is present because physical
410 momentum transfer between the slabs occurs in two directions ($+z$ and
411 $-z$). The velocity gradient ${\langle \partial v_x /\partial z
412 \rangle}$ was obtained using linear regression of the mean $x$
413 component of the velocity, $\langle v_x \rangle$, in each of the bins.
414 For Lennard-Jones simulations, shear viscosities are reported in
415 reduced units (${\eta^* = \eta \sigma^2 (\varepsilon m)^{-1/2}}$).
416
417 \subsection{Thermal Conductivities}
418
419 The energy flux was calculated in a similar manner to the momentum
420 flux, using the total non-physical energy transferred (${E_{total}}$)
421 and the data collection time $t$:
422 \begin{equation}
423 J_z = \frac{E_{total}}{2 t L_x L_y}
424 \end{equation}
425 The temperature gradient ${\langle\partial T/\partial z\rangle}$ was
426 obtained by a linear regression of the temperature profile. For
427 Lennard-Jones simulations, thermal conductivities are reported in
428 reduced units (${\lambda^*=\lambda \sigma^2 m^{1/2}
429 k_B^{-1}\varepsilon^{-1/2}}$).
430
431 \subsection{Interfacial Thermal Conductivities}
432
433 For interfaces with a relatively low interfacial conductance, the bulk
434 regions on either side of an interface rapidly come to a state in
435 which the two phases have relatively homogeneous (but distinct)
436 temperatures. The interfacial thermal conductivity $G$ can therefore
437 be approximated as:
438
439 \begin{equation}
440 G = \frac{E_{total}}{2 t L_x L_y \left( \langle T_\mathrm{hot}\rangle -
441 \langle T_\mathrm{cold}\rangle \right)}
442 \label{interfaceCalc}
443 \end{equation}
444 where ${E_{total}}$ is the imposed non-physical kinetic energy
445 transfer and ${\langle T_\mathrm{hot}\rangle}$ and ${\langle
446 T_\mathrm{cold}\rangle}$ are the average observed temperature of the
447 two separated phases. If the interfacial conductance is {\it not}
448 small, it would also be possible to compute the interfacial thermal
449 conductivity using this method by computing the change in the slope of
450 the thermal gradient ($\partial^2 \langle T \rangle /
451 \partial z^2$) at the interface.
452
453 \section{Results}
454
455 \subsection{Lennard-Jones Fluid}
456 2592 Lennard-Jones atoms were placed in an orthorhombic cell
457 ${10.06\sigma \times 10.06\sigma \times 30.18\sigma}$ on a side. The
458 reduced density ${\rho^* = \rho\sigma^3}$ was thus 0.85, which enabled
459 direct comparison between our results and previous methods. These
460 simulations were carried out with a reduced timestep ${\tau^* =
461 4.6\times10^{-4}}$. For the shear viscosity calculations, the mean
462 temperature was ${T^* = k_B T/\varepsilon = 0.72}$. For thermal
463 conductivity calculations, simulations were run under reduced
464 temperature ${\langle T^*\rangle = 0.72}$ in the microcanonical
465 ensemble. The simulations included $10^5$ steps of equilibration
466 without any momentum flux, $10^5$ steps of stablization with an
467 imposed momentum transfer to create a gradient, and $10^6$ steps of
468 data collection under RNEMD.
469
470 \subsubsection*{Thermal Conductivity}
471
472 Our thermal conductivity calculations show that the NIVS method agrees
473 well with the swapping method. Five different swap intervals were
474 tested (Table \ref{LJ}). Similar thermal gradients were observed with
475 similar thermal flux under the two different methods (Fig.
476 \ref{thermalGrad}). Furthermore, the 1-d temperature profiles showed
477 no observable differences between the $x$, $y$ and $z$ axes (lower
478 panel of Fig. \ref{thermalGrad}), so even though we are using a
479 non-isotropic scaling method, none of the three directions are
480 experience disproportionate heating due to the velocity scaling.
481
482 \begin{table*}
483 \begin{minipage}{\linewidth}
484 \begin{center}
485
486 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
487 ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
488 ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
489 at various momentum fluxes. The original swapping method and
490 the velocity scaling method give similar results.
491 Uncertainties are indicated in parentheses.}
492
493 \begin{tabular}{|cc|cc|cc|}
494 \hline
495 \multicolumn{2}{|c}{Momentum Exchange} &
496 \multicolumn{2}{|c}{Swapping RNEMD} &
497 \multicolumn{2}{|c|}{NIVS-RNEMD} \\
498 \hline
499 \multirow{2}{*}{Swap Interval (timesteps)} & Equivalent $J_z^*$ or &
500 \multirow{2}{*}{$\lambda^*_{swap}$} &
501 \multirow{2}{*}{$\eta^*_{swap}$} &
502 \multirow{2}{*}{$\lambda^*_{scale}$} &
503 \multirow{2}{*}{$\eta^*_{scale}$} \\
504 & $j_z^*(p_x)$ (reduced units) & & & & \\
505 \hline
506 250 & 0.16 & 7.03(0.34) & 3.57(0.06) & 7.30(0.10) & 3.54(0.04)\\
507 500 & 0.09 & 7.03(0.14) & 3.64(0.05) & 6.95(0.09) & 3.76(0.09)\\
508 1000 & 0.047 & 6.91(0.42) & 3.52(0.16) & 7.19(0.07) & 3.66(0.06)\\
509 2000 & 0.024 & 7.52(0.15) & 3.72(0.05) & 7.19(0.28) & 3.32(0.18)\\
510 2500 & 0.019 & 7.41(0.29) & 3.42(0.06) & 7.98(0.33) & 3.43(0.08)\\
511 \hline
512 \end{tabular}
513 \label{LJ}
514 \end{center}
515 \end{minipage}
516 \end{table*}
517
518 \begin{figure}
519 \includegraphics[width=\linewidth]{thermalGrad}
520 \caption{The NIVS-RNEMD method creates similar temperature gradients
521 compared with the swapping method under a variety of imposed
522 kinetic energy flux values. Furthermore, the implementation of
523 Non-Isotropic Velocity Scaling does not cause temperature
524 anisotropy to develop in thermal conductivity calculations.}
525 \label{thermalGrad}
526 \end{figure}
527
528 \subsubsection*{Velocity Distributions}
529
530 To test the effects on the velocity distributions, we accumulated
531 velocities every 100 steps and produced distributions of both velocity
532 and speed in each of the slabs. From these distributions, we observed
533 that under high non-physical kinetic energy flux, the speed of
534 molecules in slabs where {\it swapping} occured could deviate from the
535 Maxwell-Boltzmann distribution. This behavior was also noted by Tenney
536 and Maginn.\cite{Maginn:2010} Figure \ref{thermalHist} shows how these
537 distributions deviate from an ideal distribution. In the ``hot'' slab,
538 the probability density is notched at low speeds and has a substantial
539 shoulder at higher speeds relative to the ideal distribution. In the
540 cold slab, the opposite notching and shouldering occurs. This
541 phenomenon is more obvious at high swapping rates.
542
543 The peak of the velocity distribution is substantially flattened in
544 the hot slab, and is overly sharp (with truncated wings) in the cold
545 slab. This problem is rooted in the mechanism of the swapping method.
546 Continually depleting low (high) speed particles in the high (low)
547 temperature slab is not complemented by diffusion of low (high) speed
548 particles from neighboring slabs unless the swapping rate is
549 sufficiently small. Simutaneously, surplus low speed particles in the
550 cold slab do not have sufficient time to diffuse to neighboring slabs.
551 Since the thermal exchange rate must reach a minimum level to produce
552 an observable thermal gradient, the swapping-method RNEMD has a
553 relatively narrow choice of exchange times that can be utilized.
554
555 For comparison, NIVS-RNEMD produces a speed distribution closer to the
556 Maxwell-Boltzmann curve (Fig. \ref{thermalHist}). The reason for this
557 is simple; upon velocity scaling, a Gaussian distribution remains
558 Gaussian. Although a single scaling move is non-isotropic in three
559 dimensions, our criteria for choosing a set of scaling coefficients
560 helps maintain the distributions as close to isotropic as possible.
561 Consequently, NIVS-RNEMD is able to impose a non-physical thermal flux
562 without large perturbations to the velocity distributions in the two
563 slabs.
564
565 \begin{figure}
566 \includegraphics[width=\linewidth]{thermalHist}
567 \caption{Velocity and speed distributions that develop under the
568 swapping and NIVS-RNEMD methods at high flux. The distributions for
569 the hot bins (upper panels) and cold bins (lower panels) were
570 obtained from Lennard-Jones simulations with $\langle T^* \rangle =
571 4.5$ with a flux of $J_z^* \sim 5$ (equivalent to a swapping interval
572 of 10 time steps). This is a relatively large flux which shows the
573 non-thermal distributions that develop under the swapping method.
574 NIVS does a better job of producing near-thermal distributions in
575 the bins.}
576 \label{thermalHist}
577 \end{figure}
578
579
580 \subsubsection*{Shear Viscosity}
581 Our calculations (Table \ref{LJ}) show that velocity-scaling RNEMD
582 predicted similar values for shear viscosities to the swapping RNEMD
583 method. The average molecular momentum gradients of these samples are
584 shown in the upper two panels of Fig. \ref{shear}.
585
586 \begin{figure}
587 \includegraphics[width=\linewidth]{shear}
588 \caption{Average momentum gradients in shear viscosity simulations,
589 using ``swapping'' method (top panel) and NIVS-RNEMD method
590 (middle panel). NIVS-RNEMD produces a thermal anisotropy artifact
591 in the hot and cold bins when used for shear viscosity. This
592 artifact does not appear in thermal conductivity calculations.}
593 \label{shear}
594 \end{figure}
595
596 Observations of the three one-dimensional temperatures in each of the
597 slabs shows that NIVS-RNEMD does produce some thermal anisotropy,
598 particularly in the hot and cold slabs. Note that these temperature
599 measurements have been taken into account of the kinetic energy
600 contributed by the slab field velocity. However, this contribution has
601 only a comparable order of magnitude to the errors of data, and does
602 not significantly affect our observation. The lower panel of Fig.
603 \ref{shear} indicates that with a relatively large imposed momentum
604 flux, $j_z(p_x)$, the $x$ direction approaches a different temperature
605 from the $y$ and $z$ directions in both the hot and cold bins. This
606 is an artifact of the scaling constraints. After a momentum gradient
607 has been established, $P_c^x$ is always less than zero. Consequently,
608 the scaling factor $x$ is always greater than one in order to satisfy
609 the constraints. This will continually increase the kinetic energy in
610 $x$-dimension, $K_c^x$. If there is not enough time for the kinetic
611 energy to exchange among different directions and different slabs, the
612 system will exhibit the observed thermal anisotropy in the hot and
613 cold bins.
614
615 Although results between scaling and swapping methods are comparable,
616 the inherent temperature anisotropy does make NIVS-RNEMD method less
617 attractive than swapping RNEMD for shear viscosity calculations. We
618 note that this problem appears only when a large {\it linear} momentum
619 flux is applied, and does not appear in {\it thermal} flux
620 calculations.
621
622 \subsection{Bulk SPC/E water}
623
624 We compared the thermal conductivity of SPC/E water using NIVS-RNEMD
625 to the work of Bedrov {\it et al.}\cite{Bedrov:2000}, which employed
626 the original swapping RNEMD method. Bedrov {\it et
627 al.}\cite{Bedrov:2000} argued that exchange of the molecular
628 center-of-mass velocities instead of single atom velocities conserves
629 the total kinetic energy and linear momentum. This principle is also
630 adopted in our simulations. Scaling was applied to the center-of-mass
631 velocities of SPC/E water molecules.
632
633 To construct the simulations, a simulation box consisting of 1000
634 molecules were first equilibrated under ambient pressure and
635 temperature conditions using the isobaric-isothermal (NPT)
636 ensemble.\cite{melchionna93} A fixed volume was chosen to match the
637 average volume observed in the NPT simulations, and this was followed
638 by equilibration, first in the canonical (NVT) ensemble, followed by a
639 100~ps period under constant-NVE conditions without any momentum flux.
640 Another 100~ps was allowed to stabilize the system with an imposed
641 momentum transfer to create a gradient, and 1~ns was allotted for data
642 collection under RNEMD. Total system energy is recorded to ensure that
643 it is not drifted noticeably without a thermostat although
644 electrostatic interactions are involved.
645
646 In our simulations, the established temperature gradients were similar
647 to the previous work. Our simulation results at 318K are in good
648 agreement with those from Bedrov {\it et al.} (Table
649 \ref{spceThermal}). And both methods yield values in reasonable
650 agreement with experimental values.
651
652 \begin{table*}
653 \begin{minipage}{\linewidth}
654 \begin{center}
655
656 \caption{Thermal conductivity of SPC/E water under various
657 imposed thermal gradients. Uncertainties are indicated in
658 parentheses.}
659
660 \begin{tabular}{|c|c|ccc|}
661 \hline
662 \multirow{2}{*}{$\langle T\rangle$(K)} &
663 \multirow{2}{*}{$\langle dT/dz\rangle$(K/\AA)} &
664 \multicolumn{3}{|c|}{$\lambda (\mathrm{W m}^{-1}
665 \mathrm{K}^{-1})$} \\
666 & & This work & Previous simulations\cite{Bedrov:2000} &
667 Experiment\cite{WagnerKruse}\\
668 \hline
669 \multirow{3}{*}{300} & 0.38 & 0.816(0.044) & &
670 \multirow{3}{*}{0.61}\\
671 & 0.81 & 0.770(0.008) & & \\
672 & 1.54 & 0.813(0.007) & & \\
673 \hline
674 \multirow{2}{*}{318} & 0.75 & 0.750(0.032) & 0.784 &
675 \multirow{2}{*}{0.64}\\
676 & 1.59 & 0.778(0.019) & 0.730 & \\
677 \hline
678 \end{tabular}
679 \label{spceThermal}
680 \end{center}
681 \end{minipage}
682 \end{table*}
683
684 \subsection{Crystalline Gold}
685
686 To see how the method performed in a solid, we calculated thermal
687 conductivities using two atomistic models for gold. Several different
688 potential models have been developed that reasonably describe
689 interactions in transition metals. In particular, the Embedded Atom
690 Model (EAM)~\cite{PhysRevB.33.7983} and Sutton-Chen (SC)~\cite{Chen90}
691 potential have been used to study a wide range of phenomena in both
692 bulk materials and
693 nanoparticles.\cite{ISI:000207079300006,Clancy:1992,Vardeman:2008fk,Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq}
694 Both potentials are based on a model of a metal which treats the
695 nuclei and core electrons as pseudo-atoms embedded in the electron
696 density due to the valence electrons on all of the other atoms in the
697 system. The SC potential has a simple form that closely resembles the
698 Lennard Jones potential,
699 \begin{equation}
700 \label{eq:SCP1}
701 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] ,
702 \end{equation}
703 where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
704 \begin{equation}
705 \label{eq:SCP2}
706 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}}.
707 \end{equation}
708 $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
709 interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
710 Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
711 the interactions between the valence electrons and the cores of the
712 pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
713 scale, $c_i$ scales the attractive portion of the potential relative
714 to the repulsive interaction and $\alpha_{ij}$ is a length parameter
715 that assures a dimensionless form for $\rho$. These parameters are
716 tuned to various experimental properties such as the density, cohesive
717 energy, and elastic moduli for FCC transition metals. The quantum
718 Sutton-Chen (QSC) formulation matches these properties while including
719 zero-point quantum corrections for different transition
720 metals.\cite{PhysRevB.59.3527} The EAM functional forms differ
721 slightly from SC but the overall method is very similar.
722
723 In this work, we have utilized both the EAM and the QSC potentials to
724 test the behavior of scaling RNEMD.
725
726 A face-centered-cubic (FCC) lattice was prepared containing 2880 Au
727 atoms (i.e. ${6\times 6\times 20}$ unit cells). Simulations were run
728 both with and without isobaric-isothermal (NPT)~\cite{melchionna93}
729 pre-equilibration at a target pressure of 1 atm. When equilibrated
730 under NPT conditions, our simulation box expanded by approximately 1\%
731 in volume. Following adjustment of the box volume, equilibrations in
732 both the canonical and microcanonical ensembles were carried out. With
733 the simulation cell divided evenly into 10 slabs, different thermal
734 gradients were established by applying a set of target thermal fluxes.
735
736 The results for the thermal conductivity of gold are shown in Table
737 \ref{AuThermal}. In these calculations, the end and middle slabs were
738 excluded from the thermal gradient linear regession. EAM predicts
739 slightly larger thermal conductivities than QSC. However, both values
740 are smaller than experimental value by a factor of more than 200. This
741 behavior has been observed previously by Richardson and Clancy, and
742 has been attributed to the lack of electronic contribution in these
743 force fields.\cite{Clancy:1992} It should be noted that the density of
744 the metal being simulated has an effect on thermal conductance. With
745 an expanded lattice, lower thermal conductance is expected (and
746 observed). We also observed a decrease in thermal conductance at
747 higher temperatures, a trend that agrees with experimental
748 measurements.\cite{AshcroftMermin}
749
750 \begin{table*}
751 \begin{minipage}{\linewidth}
752 \begin{center}
753
754 \caption{Calculated thermal conductivity of crystalline gold
755 using two related force fields. Calculations were done at both
756 experimental and equilibrated densities and at a range of
757 temperatures and thermal flux rates. Uncertainties are
758 indicated in parentheses. Richardson {\it et
759 al.}\cite{Clancy:1992} give an estimate of 1.74 $\mathrm{W
760 m}^{-1}\mathrm{K}^{-1}$ for EAM gold
761 at a density of 19.263 g / cm$^3$.}
762
763 \begin{tabular}{|c|c|c|cc|}
764 \hline
765 Force Field & $\rho$ (g/cm$^3$) & ${\langle T\rangle}$ (K) &
766 $\langle dT/dz\rangle$ (K/\AA) & $\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$\\
767 \hline
768 \multirow{7}{*}{QSC} & \multirow{3}{*}{19.188} & \multirow{3}{*}{300} & 1.44 & 1.10(0.06)\\
769 & & & 2.86 & 1.08(0.05)\\
770 & & & 5.14 & 1.15(0.07)\\\cline{2-5}
771 & \multirow{4}{*}{19.263} & \multirow{2}{*}{300} & 2.31 & 1.25(0.06)\\
772 & & & 3.02 & 1.26(0.05)\\\cline{3-5}
773 & & \multirow{2}{*}{575} & 3.02 & 1.02(0.07)\\
774 & & & 4.84 & 0.92(0.05)\\
775 \hline
776 \multirow{8}{*}{EAM} & \multirow{3}{*}{19.045} & \multirow{3}{*}{300} & 1.24 & 1.24(0.16)\\
777 & & & 2.06 & 1.37(0.04)\\
778 & & & 2.55 & 1.41(0.07)\\\cline{2-5}
779 & \multirow{5}{*}{19.263} & \multirow{3}{*}{300} & 1.06 & 1.45(0.13)\\
780 & & & 2.04 & 1.41(0.07)\\
781 & & & 2.41 & 1.53(0.10)\\\cline{3-5}
782 & & \multirow{2}{*}{575} & 2.82 & 1.08(0.03)\\
783 & & & 4.14 & 1.08(0.05)\\
784 \hline
785 \end{tabular}
786 \label{AuThermal}
787 \end{center}
788 \end{minipage}
789 \end{table*}
790
791 \subsection{Thermal Conductance at the Au/H$_2$O interface}
792 The most attractive aspect of the scaling approach for RNEMD is the
793 ability to use the method in non-homogeneous systems, where molecules
794 of different identities are segregated in different slabs. To test
795 this application, we simulated a Gold (111) / water interface. To
796 construct the interface, a box containing a lattice of 1188 Au atoms
797 (with the 111 surface in the $+z$ and $-z$ directions) was allowed to
798 relax under ambient temperature and pressure. A separate (but
799 identically sized) box of SPC/E water was also equilibrated at ambient
800 conditions. The two boxes were combined by removing all water
801 molecules within 3 \AA~ radius of any gold atom. The final
802 configuration contained 1862 SPC/E water molecules.
803
804 The Spohr potential was adopted in depicting the interaction between
805 metal atoms and water molecules.\cite{ISI:000167766600035} A similar
806 protocol of equilibration to our water simulations was followed. We
807 observed that the two phases developed large temperature differences
808 even under a relatively low thermal flux.
809
810 The low interfacial conductance is probably due to an acoustic
811 impedance mismatch between the solid and the liquid
812 phase.\cite{Cahill:793,RevModPhys.61.605} Experiments on the thermal
813 conductivity of gold nanoparticles and nanorods in solvent and in
814 glass cages have predicted values for $G$ between 100 and 350
815 (MW/m$^2$/K), two orders of magnitude larger than the value reported
816 here. The experiments typically have multiple surfaces that have been
817 protected by ionic surfactants, either
818 citrate\cite{Wilson:2002uq,plech:195423} or cetyltrimethylammonium
819 bromide (CTAB), or which are in direct contact with various glassy
820 solids. In these cases, the acoustic impedance mismatch would be
821 substantially reduced, leading to much higher interfacial
822 conductances. It is also possible, however, that the lack of
823 electronic effects that gives rise to the low thermal conductivity of
824 EAM gold is also causing a low reading for this particular interface.
825
826 Under this low thermal conductance, both gold and water phases have
827 sufficient time to eliminate local temperature differences (Fig.
828 \ref{interface}). With flat thermal profiles within each phase, it is
829 valid to assume that the temperature difference between gold and water
830 surfaces would be approximately the same as the difference between the
831 gold and water bulk regions. This assumption enables convenient
832 calculation of $G$ using Eq. \ref{interfaceCalc}.
833
834 \begin{figure}
835 \includegraphics[width=\linewidth]{interface}
836 \caption{Temperature profiles of the Gold / Water interface at four
837 different values for the thermal flux. Temperatures for slabs
838 either in the gold or in the water show no significant differences,
839 although there is a large discontinuity between the materials
840 because of the relatively low interfacial thermal conductivity.}
841 \label{interface}
842 \end{figure}
843
844 \begin{table*}
845 \begin{minipage}{\linewidth}
846 \begin{center}
847
848 \caption{Computed interfacial thermal conductivity ($G$) values
849 for the Au(111) / water interface at ${\langle T\rangle \sim}$
850 300K using a range of energy fluxes. Uncertainties are
851 indicated in parentheses. }
852
853 \begin{tabular}{|cccc| }
854 \hline
855 $J_z$ (MW/m$^2$) & $\langle T_{gold} \rangle$ (K) & $\langle
856 T_{water} \rangle$ (K) & $G$
857 (MW/m$^2$/K)\\
858 \hline
859 98.0 & 355.2 & 295.8 & 1.65(0.21) \\
860 78.8 & 343.8 & 298.0 & 1.72(0.32) \\
861 73.6 & 344.3 & 298.0 & 1.59(0.24) \\
862 49.2 & 330.1 & 300.4 & 1.65(0.35) \\
863 \hline
864 \end{tabular}
865 \label{interfaceRes}
866 \end{center}
867 \end{minipage}
868 \end{table*}
869
870
871 \section{Conclusions}
872
873 Our simulations demonstrate that validity of non-isotropic velocity
874 scaling (NIVS) in RNEMD calculations of thermal conductivity in atomic
875 and molecular liquids and solids. NIVS-RNEMD improves the problematic
876 velocity distributions which can develop in other RNEMD methods.
877 Furthermore, it provides a means for carrying out non-physical thermal
878 transfer between different species of molecules, and thus extends its
879 applicability to interfacial systems. Our calculation of the gold /
880 water interfacial thermal conductivity demonstrates this advantage
881 over previous RNEMD methods. NIVS-RNEMD also has limited applications
882 for shear viscosity calculations, but has the potential to cause
883 temperature anisotropy under high momentum fluxes. Further work will
884 be necessary to eliminate the one-dimensional heating if shear
885 viscosities are required.
886
887 \section{Acknowledgments}
888 The authors would like to thank Craig Tenney and Ed Maginn for many
889 helpful discussions. Support for this project was provided by the
890 National Science Foundation under grant CHE-0848243. Computational
891 time was provided by the Center for Research Computing (CRC) at the
892 University of Notre Dame.
893 \newpage
894
895 \bibliography{nivsRnemd}
896
897 \end{doublespace}
898 \end{document}
899