ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3785
Committed: Tue Dec 13 23:31:35 2011 UTC (12 years, 6 months ago) by skuang
Content type: application/x-tex
File size: 41347 byte(s)
Log Message:
a revised version ready

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 reference note
26 \bibpunct{[}{]}{,}{n}{}{;}
27 \bibliographystyle{achemso}
28
29 \begin{document}
30
31 \title{A minimal perturbation approach to RNEMD able to simultaneously
32 impose thermal and momentum gradients}
33
34 \author{Shenyu Kuang and J. Daniel
35 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
36 Department of Chemistry and Biochemistry,\\
37 University of Notre Dame\\
38 Notre Dame, Indiana 46556}
39
40 \date{\today}
41
42 \maketitle
43
44 \begin{doublespace}
45
46 \begin{abstract}
47 We present a new method for introducing stable nonequilibrium
48 velocity and temperature gradients in molecular dynamics simulations
49 of heterogeneous systems. This method conserves the linear momentum
50 and total energy of the system and improves previous Reverse
51 Non-Equilibrium Molecular Dynamics (RNEMD) methods while maintaining
52 thermal velocity distributions. It also avoid thermal anisotropy
53 occured in previous NIVS simulations by using isotropic velocity
54 scaling on the molecules in specific regions of a system. To test
55 the method, we have computed the thermal conductivity and shear
56 viscosity of model liquid systems as well as the interfacial
57 frictions of a series of metal/liquid interfaces. Its ability to
58 combine the thermal and momentum gradients allows us to obtain shear
59 viscosity data for a range of temperatures in only one trajectory.
60
61 \end{abstract}
62
63 \newpage
64
65 %\narrowtext
66
67 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68 % BODY OF TEXT
69 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70
71 \section{Introduction}
72 Imposed-flux methods in Molecular Dynamics (MD)
73 simulations\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
74 can establish steady state systems with an applied flux set vs a
75 corresponding gradient that can be measured. These methods does not
76 need many trajectories to provide information of transport properties
77 of a given system. Thus, they are utilized in computing thermal and
78 mechanical transfer of homogeneous bulk systems as well as
79 heterogeneous systems such as solid-liquid
80 interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
81
82 The Reverse Non-Equilibrium MD (RNEMD) methods adopt constraints that
83 satisfy linear momentum and total energy conservation of a system when
84 imposing fluxes in a simulation. Thus they are compatible with various
85 ensembles, including the micro-canonical (NVE) ensemble, without the
86 need of an external thermostat. The original approaches proposed by
87 M\"{u}ller-Plathe {\it et
88 al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
89 momentum swapping for generating energy/momentum fluxes, which can
90 also be compatible with particles of different identities. Although
91 simple to implement in a simulation, this approach can create
92 nonthermal velocity distributions, as discovered by Tenney and
93 Maginn\cite{Maginn:2010}. Furthermore, this approach is less efficient
94 for kinetic energy transfer between particles of different identities,
95 especially when the mass difference between the particles becomes
96 significant. This also limits its applications on heterogeneous
97 interfacial systems.
98
99 Recently, we developed a different approach, using Non-Isotropic
100 Velocity Scaling (NIVS) \cite{kuang:164101} algorithm to impose
101 fluxes. Compared to the momentum swapping move, it scales the velocity
102 vectors in two separate regions of a simulated system with respective
103 diagonal scaling matrices. These matrices are determined by solving a
104 set of equations including linear momentum and kinetic energy
105 conservation constraints and target flux satisfaction. This method is
106 able to effectively impose a wide range of kinetic energy fluxes
107 without obvious perturbation to the velocity distributions of the
108 simulated systems, regardless of the presence of heterogeneous
109 interfaces. We have successfully applied this approach in studying the
110 interfacial thermal conductance at metal-solvent
111 interfaces.\cite{kuang:AuThl}
112
113 However, the NIVS approach has limited applications in imposing
114 momentum fluxes. Temperature anisotropy could happen under high
115 momentum fluxes due to the implementation of this algorithm. Thus,
116 combining thermal and momentum flux is also difficult to obtain with
117 this approach. However, such combination may provide a means to
118 simulate thermal/momentum gradient coupled processes such as Soret
119 effect in liquid flows. Therefore, developing improved approaches to
120 extend the applications of the imposed-flux method is desirable.
121
122 In this paper, we improve the RNEMD methods by proposing a novel
123 approach to impose fluxes. This approach separate the means of applying
124 momentum and thermal flux with operations in one time step and thus is
125 able to simutaneously impose thermal and momentum flux. Furthermore,
126 the approach retains desirable features of previous RNEMD approaches
127 and is simpler to implement compared to the NIVS method. In what
128 follows, we first present the method and its implementation in a
129 simulation. Then we compare the method on bulk fluids to previous
130 methods. Also, interfacial frictions are computed for a series of
131 interfaces.
132
133 \section{Methodology}
134 Similar to the NIVS method,\cite{kuang:164101} we consider a
135 periodic system divided into a series of slabs along a certain axis
136 (e.g. $z$). The unphysical thermal and/or momentum flux is designated
137 from the center slab to one of the end slabs, and thus the thermal
138 flux results in a lower temperature of the center slab than the end
139 slab, and the momentum flux results in negative center slab momentum
140 with positive end slab momentum (unless these fluxes are set
141 negative). Therefore, the center slab is denoted as ``$c$'', while the
142 end slab as ``$h$''.
143
144 To impose these fluxes, we periodically apply different set of
145 operations on velocities of particles {$i$} within the center slab and
146 those of particles {$j$} within the end slab:
147 \begin{eqnarray}
148 \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
149 \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
150 \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
151 \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
152 \end{eqnarray}
153 where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denotes
154 the instantaneous bulk velocity of slabs $c$ and $h$ respectively
155 before an operation is applied. When a momentum flux $\vec{j}_z(\vec{p})$
156 presents, these bulk velocities would have a corresponding change
157 ($\vec{a}_c$ and $\vec{a}_h$ respectively) according to Newton's
158 second law:
159 \begin{eqnarray}
160 M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \\
161 M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
162 \end{eqnarray}
163 where $M$ denotes total mass of particles within a slab:
164 \begin{eqnarray}
165 M_c & = & \sum_{i = 1}^{N_c} m_i \\
166 M_h & = & \sum_{j = 1}^{N_h} m_j
167 \end{eqnarray}
168 and $\Delta t$ is the interval between two separate operations.
169
170 The above operations already conserve the linear momentum of a
171 periodic system. To further satisfy total energy conservation as well
172 as to impose the thermal flux $J_z$, the following equations are
173 included as well:
174 [MAY PUT EXTRA MATH IN SUPPORT INFO OR APPENDIX]
175 \begin{eqnarray}
176 K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
177 \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
178 K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
179 \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2
180 \end{eqnarray}
181 where $K_c$ and $K_h$ denotes translational kinetic energy of slabs
182 $c$ and $h$ respectively before an operation is applied. These
183 translational kinetic energy conservation equations are sufficient to
184 ensure total energy conservation, as the operations applied in our
185 method do not change the kinetic energy related to other degrees of
186 freedom or the potential energy of a system, given that its potential
187 energy does not depend on particle velocity.
188
189 The above sets of equations are sufficient to determine the velocity
190 scaling coefficients ($c$ and $h$) as well as $\vec{a}_c$ and
191 $\vec{a}_h$. Note that there are two roots respectively for $c$ and
192 $h$. However, the positive roots (which are closer to 1) are chosen so
193 that the perturbations to a system can be reduced to a minimum. Figure
194 \ref{method} illustrates the implementation sketch of this algorithm
195 in an individual step.
196
197 \begin{figure}
198 \includegraphics[width=\linewidth]{method}
199 \caption{Illustration of the implementation of the algorithm in a
200 single step. Starting from an ideal velocity distribution, the
201 transformation is used to apply the effect of both a thermal and a
202 momentum flux from the ``c'' slab to the ``h'' slab. As the figure
203 shows, thermal distributions can preserve after this operation.}
204 \label{method}
205 \end{figure}
206
207 By implementing these operations at a certain frequency, a steady
208 thermal and/or momentum flux can be applied and the corresponding
209 temperature and/or momentum gradients can be established.
210
211 Compared to the previous NIVS method, this approach is computationally
212 more efficient in that only quadratic equations are involved to
213 determine a set of scaling coefficients, while the NIVS method needs
214 to solve quartic equations. Furthermore, this method implements
215 isotropic scaling of velocities in respective slabs, unlike the NIVS,
216 where an extra criteria function is necessary to choose a set of
217 coefficients that performs a scaling as isotropic as possible. More
218 importantly, separating the means of momentum flux imposing from
219 velocity scaling avoids the underlying cause to thermal anisotropy in
220 NIVS when applying a momentum flux. And later sections will
221 demonstrate that this can improve the performance in shear viscosity
222 simulations.
223
224 This approach is advantageous over the original momentum swapping in
225 many aspects. In one swapping, the velocity vectors involved are
226 usually very different (or the generated flux is trivial to obtain
227 gradients), thus the swapping tends to incur perturbations to the
228 neighbors of the particles involved. Comparatively, our approach
229 disperse the flux to every selected particle in a slab so that
230 perturbations in the flux generating region could be
231 minimized. Additionally, because the momentum swapping steps tend to
232 result in a nonthermal distribution, when an imposed flux is
233 relatively large and diffusions from the neighboring slabs could no
234 longer remedy this effect, problematic distributions would be
235 observed. In comparison, the operations of our approach has the nature
236 of preserving the equilibrium velocity distributions (commonly
237 Maxwell-Boltzmann), and results in later section will illustrate that
238 this is helpful to retain thermal distributions in a simulation.
239
240 \section{Computational Details}
241 The algorithm has been implemented in our MD simulation code,
242 OpenMD\cite{Meineke:2005gd,openmd}. We compare the method with
243 previous RNEMD methods or equilibrium MD (EMD) methods in homogeneous fluids
244 (Lennard-Jones and SPC/E water). And taking advantage of the method,
245 we simulate the interfacial friction of different heterogeneous
246 interfaces (gold-organic solvent and gold-SPC/E water and ice-liquid
247 water).
248
249 \subsection{Simulation Protocols}
250 The systems to be investigated are set up in orthorhombic simulation
251 cells with periodic boundary conditions in all three dimensions. The
252 $z$ axis of these cells were longer and set as the temperature and/or
253 momentum gradient axis. And the cells were evenly divided into $N$
254 slabs along this axis, with various $N$ depending on individual
255 system. The $x$ and $y$ axis were of the same length in homogeneous
256 systems or had length scale close to each other where heterogeneous
257 interfaces presents. In all cases, before introducing a nonequilibrium
258 method to establish steady thermal and/or momentum gradients for
259 further measurements and calculations, canonical ensemble with a
260 Nos\'e-Hoover thermostat\cite{hoover85} and microcanonical ensemble
261 equilibrations were used before data collections. For SPC/E water
262 simulations, isobaric-isothermal equilibrations\cite{melchionna93} are
263 performed before the above to reach normal pressure (1 bar); for
264 interfacial systems, similar equilibrations are used to relax the
265 surface tensions of the $xy$ plane.
266
267 While homogeneous fluid systems can be set up with rather random
268 configurations, our interfacial systems needs a series of steps to
269 ensure the interfaces be established properly for computations. The
270 preparation and equilibration of butanethiol covered gold (111)
271 surface and further solvation and equilibration process is described
272 in details as in reference \cite{kuang:AuThl}.
273
274 As for the ice/liquid water interfaces, the basal surface of ice
275 lattice was first constructed. Hirsch {\it et
276 al.}\cite{doi:10.1021/jp048434u} explored the energetics of ice
277 lattices with all possible proton order configurations. We refer to
278 their results and choose the configuration of the lowest energy after
279 geometry optimization as the unit cell for our ice lattices. Although
280 experimental solid/liquid coexistant temperature under normal pressure
281 should be close to 273K, Bryk and Haymet's simulations of ice/liquid
282 water interfaces with different models suggest that for SPC/E, the
283 most stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
284 Therefore, our ice/liquid water simulations were carried out at
285 225K. To have extra protection of the ice lattice during initial
286 equilibration (when the randomly generated liquid phase configuration
287 could release large amount of energy in relaxation), restraints were
288 applied to the ice lattice to avoid inadvertent melting by the heat
289 dissipated from the high enery configurations.
290 [MAY ADD A SNAPSHOT FOR BASAL PLANE]
291
292 \subsection{Force Field Parameters}
293 For comparison of our new method with previous work, we retain our
294 force field parameters consistent with previous simulations. Argon is
295 the Lennard-Jones fluid used here, and its results are reported in
296 reduced unit for direct comparison purpose.
297
298 As for our water simulations, SPC/E model is used throughout this work
299 for consistency. Previous work for transport properties of SPC/E water
300 model is available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so
301 that unnecessary repetition of previous methods can be avoided.
302
303 The Au-Au interaction parameters in all simulations are described by
304 the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
305 QSC potentials include zero-point quantum corrections and are
306 reparametrized for accurate surface energies compared to the
307 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
308 Spohr potential was adopted\cite{ISI:000167766600035} to depict
309 Au-H$_2$O interactions.
310
311 For our gold/organic liquid interfaces, the small organic molecules
312 included in our simulations are the Au surface capping agent
313 butanethiol and liquid hexane and toluene. The United-Atom
314 models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
315 for these components were used in this work for better computational
316 efficiency, while maintaining good accuracy. We refer readers to our
317 previous work\cite{kuang:AuThl} for further details of these models,
318 as well as the interactions between Au and the above organic molecule
319 components.
320
321 \subsection{Thermal conductivities}
322 When $\vec{j}_z(\vec{p})$ is set to zero and a target $J_z$ is set to
323 impose kinetic energy transfer, the method can be used for thermal
324 conductivity computations. Similar to previous RNEMD methods, we
325 assume linear response of the temperature gradient with respect to the
326 thermal flux in general case. And the thermal conductivity ($\lambda$)
327 can be obtained with the imposed kinetic energy flux and the measured
328 thermal gradient:
329 \begin{equation}
330 J_z = -\lambda \frac{\partial T}{\partial z}
331 \end{equation}
332 Like other imposed-flux methods, the energy flux was calculated using
333 the total non-physical energy transferred (${E_{total}}$) from slab
334 ``c'' to slab ``h'', which is recorded throughout a simulation, and
335 the time for data collection $t$:
336 \begin{equation}
337 J_z = \frac{E_{total}}{2 t L_x L_y}
338 \end{equation}
339 where $L_x$ and $L_y$ denotes the dimensions of the plane in a
340 simulation cell perpendicular to the thermal gradient, and a factor of
341 two in the denominator is necessary for the heat transport occurs in
342 both $+z$ and $-z$ directions. The average temperature gradient
343 ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
344 regression of the temperature profile, which is recorded during a
345 simulation for each slab in a cell. For Lennard-Jones simulations,
346 thermal conductivities are reported in reduced units
347 (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
348
349 \subsection{Shear viscosities}
350 Alternatively, the method can carry out shear viscosity calculations
351 by specify a momentum flux. In our algorithm, one can specify the
352 three components of the flux vector $\vec{j}_z(\vec{p})$
353 respectively. For shear viscosity simulations, $j_z(p_z)$ is usually
354 set to zero. For isotropic systems, the direction of
355 $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
356 ability of arbitarily specifying the vector direction in our method
357 could provide convenience in anisotropic simulations.
358
359 Similar to thermal conductivity computations, for a homogeneous
360 system, linear response of the momentum gradient with respect to the
361 shear stress is assumed, and the shear viscosity ($\eta$) can be
362 obtained with the imposed momentum flux (e.g. in $x$ direction) and
363 the measured gradient:
364 \begin{equation}
365 j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
366 \end{equation}
367 where the flux is similarly defined:
368 \begin{equation}
369 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
370 \end{equation}
371 with $P_x$ being the total non-physical momentum transferred within
372 the data collection time. Also, the averaged velocity gradient
373 ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
374 regression of the $x$ component of the mean velocity ($\langle
375 v_x\rangle$) in each of the bins. For Lennard-Jones simulations, shear
376 viscosities are also reported in reduced units
377 (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
378
379 Although $J_z$ may be switched off for shear viscosity simulations at
380 a certain temperature, our method's ability to impose both a thermal
381 and a momentum flux in one simulation allows the combination of a
382 temperature and a velocity gradient. In this case, since viscosity is
383 generally a function of temperature, the local viscosity also depends
384 on the local temperature. Therefore, in one such simulation, viscosity
385 at $z$ (corresponding to a certain $T$) can be computed with the
386 applied shear flux and the local velocity gradient (which can be
387 obtained by finite difference approximation). As a whole, the
388 viscosity can be mapped out as the function of temperature in one
389 single trajectory of simulation. Results for shear viscosity
390 computations of SPC/E water will demonstrate its effectiveness in
391 detail.
392
393 \subsection{Interfacial friction and Slip length}
394 While the shear stress results in a velocity gradient within bulk
395 fluid phase, its effect at a solid-liquid interface could vary due to
396 the interaction strength between the two phases. The interfacial
397 friction coefficient $\kappa$ is defined to relate the shear stress
398 (e.g. along $x$-axis) with the relative fluid velocity tangent to the
399 interface:
400 \begin{equation}
401 j_z(p_x)|_{interface} = \kappa\Delta v_x|_{interface}
402 \end{equation}
403 Under ``stick'' boundary condition, $\Delta v_x|_{interface}
404 \rightarrow 0$, which leads to $\kappa\rightarrow\infty$. However, for
405 ``slip'' boundary conditions at the solid-liquid interfaces, $\kappa$
406 becomes finite. To characterize the interfacial boundary conditions,
407 slip length ($\delta$) is defined using $\kappa$ and the shear
408 viscocity of liquid phase ($\eta$):
409 \begin{equation}
410 \delta = \frac{\eta}{\kappa}
411 \end{equation}
412 so that $\delta\rightarrow 0$ in the ``no-slip'' boundary condition,
413 and depicts how ``slippery'' an interface is. Figure \ref{slipLength}
414 illustrates how this quantity is defined and computed for a
415 solid-liquid interface. [MAY INCLUDE SNAPSHOT IN FIGURE]
416
417 \begin{figure}
418 \includegraphics[width=\linewidth]{defDelta}
419 \caption{The slip length $\delta$ can be obtained from a velocity
420 profile of a solid-liquid interface simulation, when a momentum flux
421 is applied. An example of Au/hexane interfaces is shown, and the
422 calculation for the left side is illustrated. The calculation for
423 the right side is similar to the left.}
424 \label{slipLength}
425 \end{figure}
426
427 In our method, a shear stress can be applied similar to shear
428 viscosity computations by applying an unphysical momentum flux
429 (e.g. $j_z(p_x)$). A corresponding velocity profile can be obtained as
430 shown in Figure \ref{slipLength}, in which the velocity gradients
431 within liquid phase and velocity difference at the liquid-solid
432 interface can be measured respectively. Further calculations and
433 characterizations of the interface can be carried out using these
434 data.
435
436 \section{Results and Discussions}
437 \subsection{Lennard-Jones fluid}
438 Our orthorhombic simulation cell of Lennard-Jones fluid has identical
439 parameters to our previous work\cite{kuang:164101} to facilitate
440 comparison. Thermal conductivitis and shear viscosities were computed
441 with the algorithm applied to the simulations. The results of thermal
442 conductivity are compared with our previous NIVS algorithm. However,
443 since the NIVS algorithm could produce temperature anisotropy for
444 shear viscocity computations, these results are instead compared to
445 the momentum swapping approaches. Table \ref{LJ} lists these
446 calculations with various fluxes in reduced units.
447
448 \begin{table*}
449 \begin{minipage}{\linewidth}
450 \begin{center}
451
452 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
453 ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
454 ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
455 at various momentum fluxes. The new method yields similar
456 results to previous RNEMD methods. All results are reported in
457 reduced unit. Uncertainties are indicated in parentheses.}
458
459 \begin{tabular}{cccccc}
460 \hline\hline
461 \multicolumn{2}{c}{Momentum Exchange} &
462 \multicolumn{2}{c}{$\lambda^*$} &
463 \multicolumn{2}{c}{$\eta^*$} \\
464 \hline
465 Swap Interval $t^*$ & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
466 NIVS\cite{kuang:164101} & This work & Swapping & This work \\
467 \hline
468 0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
469 0.232 & 0.09 & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
470 0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
471 0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
472 1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
473 \hline\hline
474 \end{tabular}
475 \label{LJ}
476 \end{center}
477 \end{minipage}
478 \end{table*}
479
480 \subsubsection{Thermal conductivity}
481 Our thermal conductivity calculations with this method yields
482 comparable results to the previous NIVS algorithm. This indicates that
483 the thermal gradients introduced using this method are also close to
484 previous RNEMD methods. Simulations with moderately higher thermal
485 fluxes tend to yield more reliable thermal gradients and thus avoid
486 large errors, while overly high thermal fluxes could introduce side
487 effects such as non-linear temperature gradient response or
488 inadvertent phase transitions.
489
490 Since the scaling operation is isotropic in this method, one does not
491 need extra care to ensure temperature isotropy between the $x$, $y$
492 and $z$ axes, while for NIVS, thermal anisotropy might happen if the
493 criteria function for choosing scaling coefficients does not perform
494 as expected. Furthermore, this method avoids inadvertent concomitant
495 momentum flux when only thermal flux is imposed, which could not be
496 achieved with swapping or NIVS approaches. The thermal energy exchange
497 in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
498 or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
499 P^\alpha$) would not achieve this effect unless thermal flux vanishes
500 (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which do not contribute to
501 applying a thermal flux). In this sense, this method aids to achieve
502 minimal perturbation to a simulation while imposing a thermal flux.
503
504 \subsubsection{Shear viscosity}
505 Table \ref{LJ} also compares our shear viscosity results with the
506 momentum swapping approach. Our calculations show that our method
507 predicted similar values of shear viscosities to the momentum swapping
508 approach, as well as the velocity gradient profiles. Moderately larger
509 momentum fluxes are helpful to reduce the errors of measured velocity
510 gradients and thus the final result. However, it is pointed out that
511 the momentum swapping approach tends to produce nonthermal velocity
512 distributions.\cite{Maginn:2010}
513
514 To examine that temperature isotropy holds in simulations using our
515 method, we measured the three one-dimensional temperatures in each of
516 the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
517 temperatures were calculated after subtracting the contribution from
518 bulk velocities of the slabs. The one-dimensional temperature profiles
519 showed no observable difference between the three dimensions. This
520 ensures that isotropic scaling automatically preserves temperature
521 isotropy and that our method is useful in shear viscosity
522 computations.
523
524 \begin{figure}
525 \includegraphics[width=\linewidth]{tempXyz}
526 \caption{Unlike the previous NIVS algorithm, the new method does not
527 produce a thermal anisotropy. No temperature difference between
528 different dimensions were observed beyond the magnitude of the error
529 bars. Note that the two ``hotter'' regions are caused by the shear
530 stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
531 an effect that only observed in our methods.}
532 \label{tempXyz}
533 \end{figure}
534
535 Furthermore, the velocity distribution profiles are tested by imposing
536 a large shear stress into the simulations. Figure \ref{vDist}
537 demonstrates how our method is able to maintain thermal velocity
538 distributions against the momentum swapping approach even under large
539 imposed fluxes. Previous swapping methods tend to deplete particles of
540 positive velocities in the negative velocity slab (``c'') and vice
541 versa in slab ``h'', where the distributions leave notchs. This
542 problematic profiles become significant when the imposed-flux becomes
543 larger and diffusions from neighboring slabs could not offset the
544 depletions. Simutaneously, abnormal peaks appear corresponding to
545 excessive particles having velocity swapped from the other slab. These
546 nonthermal distributions limit applications of the swapping approach
547 in shear stress simulations. Our method avoids the above problematic
548 distributions by altering the means of applying momentum
549 fluxes. Comparatively, velocity distributions recorded from
550 simulations with our method is so close to the ideal thermal
551 prediction that no obvious difference is shown in Figure
552 \ref{vDist}. Conclusively, our method avoids problems that occurs in
553 previous RNEMD methods and provides a useful means for shear viscosity
554 computations.
555
556 \begin{figure}
557 \includegraphics[width=\linewidth]{velDist}
558 \caption{Velocity distributions that develop under the swapping and
559 our methods at a large flux. These distributions were obtained from
560 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
561 swapping interval of 20 time steps). This is a relatively large flux
562 to demonstrate the nonthermal distributions that develop under the
563 swapping method. In comparison, distributions produced by our method
564 are very close to the ideal thermal situations.}
565 \label{vDist}
566 \end{figure}
567
568 \subsection{Bulk SPC/E water}
569 We extend our applications of thermal conductivity and shear viscosity
570 computations to a complex fluid model of SPC/E water. A simulation
571 cell with 1000 molecules was set up in the similar manner as in
572 \cite{kuang:164101}. For thermal conductivity simulations,
573 measurements were taken to compare with previous RNEMD methods; for
574 shear viscosity computations, simulations were run under a series of
575 temperatures (with corresponding pressure relaxation using the
576 isobaric-isothermal ensemble\cite{melchionna93}), and results were
577 compared to available data from EMD
578 methods\cite{10.1063/1.3330544,Medina2011}. Besides, a simulation with
579 both thermal and momentum gradient was carried out to map out shear
580 viscosity as a function of temperature to see the effectiveness and
581 accuracy our method could reach.
582
583 \subsubsection{Thermal conductivity}
584 Table \ref{spceThermal} summarizes our thermal conductivity
585 computations under different temperatures and thermal gradients, in
586 comparison to the previous NIVS results\cite{kuang:164101} and
587 experimental measurements\cite{WagnerKruse}. Note that no appreciable
588 drift of total system energy or temperature was observed when our
589 method is applied, which indicates that our algorithm conserves total
590 energy well for systems involving electrostatic interactions.
591
592 Measurements using our method established similar temperature
593 gradients to the previous NIVS method. Our simulation results are in
594 good agreement with those from previous simulations. And both methods
595 yield values in reasonable agreement with experimental
596 values. Simulations using moderately higher thermal gradient or those
597 with longer gradient axis ($z$) for measurement seem to have better
598 accuracy, from our results.
599
600 \begin{table*}
601 \begin{minipage}{\linewidth}
602 \begin{center}
603
604 \caption{Thermal conductivity of SPC/E water under various
605 imposed thermal gradients. Uncertainties are indicated in
606 parentheses.}
607
608 \begin{tabular}{ccccc}
609 \hline\hline
610 $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
611 {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
612 (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
613 Experiment\cite{WagnerKruse} \\
614 \hline
615 300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\
616 318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
617 & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
618 & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
619 twice as long.} & & \\
620 \hline\hline
621 \end{tabular}
622 \label{spceThermal}
623 \end{center}
624 \end{minipage}
625 \end{table*}
626
627 \subsubsection{Shear viscosity}
628 The improvement our method achieves for shear viscosity computations
629 enables us to apply it on SPC/E water models. The series of
630 temperatures under which our shear viscosity calculations were carried
631 out covers the liquid range under normal pressure. Our simulations
632 predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
633 (Table \ref{spceShear}). Considering subtlties such as temperature or
634 pressure/density errors in these two series of measurements, our
635 results show no significant difference from those with EMD
636 methods. Since each value reported using our method takes only one
637 single trajectory of simulation, instead of average from many
638 trajectories when using EMD, our method provides an effective means
639 for shear viscosity computations.
640
641 \begin{table*}
642 \begin{minipage}{\linewidth}
643 \begin{center}
644
645 \caption{Computed shear viscosity of SPC/E water under different
646 temperatures. Results are compared to those obtained with EMD
647 method[CITATION]. Uncertainties are indicated in parentheses.}
648
649 \begin{tabular}{cccc}
650 \hline\hline
651 $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
652 {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
653 (K) & (10$^{10}$s$^{-1}$) & This work & Previous
654 simulations\cite{Medina2011} \\
655 \hline
656 273 & 1.12 & 1.218(0.004) & 1.282(0.048) \\
657 & 1.79 & 1.140(0.012) & \\
658 303 & 2.09 & 0.646(0.008) & 0.643(0.019) \\
659 318 & 2.50 & 0.536(0.007) & \\
660 & 5.25 & 0.510(0.007) & \\
661 & 2.82 & 0.474(0.003)\footnote{Simulation with $L_z$ twice
662 as long.} & \\
663 333 & 3.10 & 0.428(0.002) & 0.421(0.008) \\
664 363 & 2.34 & 0.279(0.014) & 0.291(0.005) \\
665 & 4.26 & 0.306(0.001) & \\
666 \hline\hline
667 \end{tabular}
668 \label{spceShear}
669 \end{center}
670 \end{minipage}
671 \end{table*}
672
673 A more effective way to map out $\eta$ vs $T$ is to combine a momentum
674 flux with a thermal flux. Figure \ref{Tvxdvdz} shows the thermal and
675 velocity gradient in one such simulation. At different positions with
676 different temperatures, the velocity gradient is not a constant but
677 can be computed locally. With the data provided in Figure
678 \ref{Tvxdvdz}, a series of $\eta$ is calculated as in Figure
679 \ref{etaT} and a linear fit was performed to $\partial v_x/\partial z$
680 vs. $z$ so that the resulted $\eta$ can be present as a curve as
681 well. For comparison, other results are also mapped in the figure.
682
683 \begin{figure}
684 \includegraphics[width=\linewidth]{tvxdvdz}
685 \caption{With a combination of a thermal and a momentum flux, a
686 simulation can have both a temperature (top) and a velocity (middle)
687 gradient. Due to the thermal gradient, $\partial v_x/\partial z$ is
688 not constant but can be computed using finite difference
689 approximations (lower). These data can be used further to calculate
690 $\eta$ vs $T$ (Figure \ref{etaT}).}
691 \label{Tvxdvdz}
692 \end{figure}
693
694 From Figure \ref{etaT}, one can see that the generated curve agrees
695 well with the above RNEMD simulations at different temperatures, as
696 well as results reported using EMD
697 methods\cite{10.1063/1.3330544,Medina2011} in much of the temperature
698 range simulated. However, this curve has relatively large error in
699 lower temperature regions and has some difference in predicting $\eta$
700 near 273K. Provided that this curve only takes one trajectory to
701 generate, these results are of satisfactory efficiency and
702 accuracy. Since previous work already pointed out that the SPC/E model
703 tends to predict lower viscosity compared to experimental
704 data,\cite{Medina2011} experimental comparison are not given here.
705
706 \begin{figure}
707 \includegraphics[width=\linewidth]{etaT}
708 \caption{The curve generated by single simulation with thermal and
709 momentum gradient predicts satisfatory values in much of the
710 temperature range under test.}
711 \label{etaT}
712 \end{figure}
713
714 \subsection{Interfacial frictions and slip lengths}
715 Another attractive aspect of our method is the ability to apply
716 momentum and/or thermal flux in nonhomogeneous systems, where
717 molecules of different identities (or phases) are segregated in
718 different regions. We have previously studied the interfacial thermal
719 transport of a series of metal gold-liquid
720 surfaces\cite{kuang:164101,kuang:AuThl}, and would like to further
721 investigate the relationship between this phenomenon and the
722 interfacial frictions.
723
724 Table \ref{etaKappaDelta} includes these computations and previous
725 calculations of corresponding interfacial thermal conductance. For
726 bare Au(111) surfaces, slip boundary conditions were observed for both
727 organic and aqueous liquid phases, corresponding to previously
728 computed low interfacial thermal conductance. In comparison, the
729 butanethiol covered Au(111) surface appeared to be sticky to the
730 organic liquid layers in our simulations. We have reported conductance
731 enhancement effect for this surface capping agent,\cite{kuang:AuThl}
732 and these observations have a qualitative agreement with the thermal
733 conductance results. This agreement also supports discussions on the
734 relationship between surface wetting and slip effect and thermal
735 conductance of the
736 interface.\cite{PhysRevLett.82.4671,doi:10.1080/0026897031000068578,garde:PhysRevLett2009}
737
738 \begin{table*}
739 \begin{minipage}{\linewidth}
740 \begin{center}
741
742 \caption{Computed interfacial friction coefficient values for
743 interfaces with various components for liquid and solid
744 phase. Error estimates are indicated in parentheses.}
745
746 \begin{tabular}{llcccccc}
747 \hline\hline
748 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
749 & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
750 \cite{kuang:164101}.} \\
751 surface & molecules & K & MPa & mPa$\cdot$s &
752 10$^4$Pa$\cdot$s/m & nm & MW/m$^2$/K \\
753 \hline
754 Au(111) & hexane & 200 & 1.08 & 0.197(0.009) & 5.30(0.36) &
755 3.72 & 46.5 \\
756 & & & 2.15 & 0.141(0.002) & 5.31(0.26) &
757 2.76 & \\
758 Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.286(0.019) & $\infty$
759 & 0 & 131 \\
760 & & & 5.39 & 0.320(0.006) & $\infty$
761 & 0 & \\
762 \hline
763 Au(111) & toluene & 200 & 1.08 & 0.722(0.035) & 15.7(0.7) &
764 4.60 & 70.1 \\
765 & & & 2.16 & 0.544(0.030) & 11.2(0.5) &
766 4.86 & \\
767 Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.980(0.057) &
768 $\infty$ & 0 & 187 \\
769 & & & 10.8 & 0.995(0.005) &
770 $\infty$ & 0 & \\
771 \hline
772 Au(111) & water & 300 & 1.08 & 0.399(0.050) & 1.928(0.022) &
773 20.7 & 1.65 \\
774 & & & 2.16 & 0.794(0.255) & 1.895(0.003) &
775 41.9 & \\
776 \hline
777 ice(basal) & water & 225 & 19.4 & 15.8(0.2) & $\infty$ & 0 & \\
778 \hline\hline
779 \end{tabular}
780 \label{etaKappaDelta}
781 \end{center}
782 \end{minipage}
783 \end{table*}
784
785 An interesting effect alongside the surface friction change is
786 observed on the shear viscosity of liquids in the regions close to the
787 solid surface. In our results, $\eta$ measured near a ``slip'' surface
788 tends to be smaller than that near a ``stick'' surface. This may
789 suggest the influence from an interface on the dynamic properties of
790 liquid within its neighbor regions. It is known that diffusions of
791 solid particles in liquid phase is affected by their surface
792 conditions (stick or slip boundary).\cite{10.1063/1.1610442} Our
793 observations could provide a support to this phenomenon.
794
795 In addition to these previously studied interfaces, we attempt to
796 construct ice-water interfaces and the basal plane of ice lattice was
797 studied here. In contrast to the Au(111)/water interface, where the
798 friction coefficient is substantially small and large slip effect
799 presents, the ice/liquid water interface demonstrates strong
800 solid-liquid interactions and appears to be sticky. The supercooled
801 liquid phase is an order of magnitude more viscous than measurements
802 in previous section. It would be of interst to investigate the effect
803 of different ice lattice planes (such as prism and other surfaces) on
804 interfacial friction and the corresponding liquid viscosity.
805
806 \section{Conclusions}
807 Our simulations demonstrate the validity of our method in RNEMD
808 computations of thermal conductivity and shear viscosity in atomic and
809 molecular liquids. Our method maintains thermal velocity distributions
810 and avoids thermal anisotropy in previous NIVS shear stress
811 simulations, as well as retains attractive features of previous RNEMD
812 methods. There is no {\it a priori} restrictions to the method to be
813 applied in various ensembles, so prospective applications to
814 extended-system methods are possible.
815
816 Our method is capable of effectively imposing thermal and/or momentum
817 flux accross an interface. This facilitates studies that relates
818 dynamic property measurements to the chemical details of an
819 interface. Therefore, investigations can be carried out to
820 characterize interfacial interactions using the method.
821
822 Another attractive feature of our method is the ability of
823 simultaneously introducing thermal and momentum gradients in a
824 system. This facilitates us to effectively map out the shear viscosity
825 with respect to a range of temperature in single trajectory of
826 simulation with satisafactory accuracy. Complex systems that involve
827 thermal and momentum gradients might potentially benefit from
828 this. For example, the Soret effects under a velocity gradient might
829 be models of interest to purification and separation researches.
830
831 \section{Acknowledgments}
832 Support for this project was provided by the National Science
833 Foundation under grant CHE-0848243. Computational time was provided by
834 the Center for Research Computing (CRC) at the University of Notre
835 Dame.
836
837 \newpage
838
839 \bibliography{stokes}
840
841 \end{doublespace}
842 \end{document}