ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3786
Committed: Wed Dec 14 21:59:12 2011 UTC (12 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 44054 byte(s)
Log Message:

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{Velocity Shearing and Scaling RNEMD: a minimally perturbing
32 method for producing temperature 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 both the linear
50 momentum and total energy of the system and improves previous
51 reverse non-equilibrium molecular dynamics (RNEMD) methods while
52 retaining equilibrium thermal velocity distributions in each region
53 of the system. The new method avoids the thermal anisotropy
54 produced by previous methods by using isotropic velocity scaling and
55 shearing on all of the molecules in specific regions. To test the
56 method, we have computed the thermal conductivity and shear
57 viscosity of model liquid systems as well as the interfacial
58 friction coeefficients of a series of solid / liquid interfaces. The
59 method's ability to combine the thermal and momentum gradients
60 allows us to obtain shear viscosity data for a range of temperatures
61 from a single trajectory.
62 \end{abstract}
63
64 \newpage
65
66 %\narrowtext
67
68 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
69 % BODY OF TEXT
70 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71
72 \section{Introduction}
73 One of the standard ways to compute transport coefficients such as the
74 viscosities and thermal conductivities of liquids is to use
75 imposed-flux non-equilibrium molecular dynamics
76 methods.\cite{MullerPlathe:1997xw,ISI:000080382700030,kuang:164101}
77 These methods establish stable non-equilibrium conditions in a
78 simulation box using an applied momentum or thermal flux between
79 different regions of the box. The corresponding temperature or
80 velocity gradients which develop in response to the applied flux is
81 related (via linear response theory) to the transport coefficient of
82 interest. These methods are quite efficient, in that they do not need
83 many trajectories to provide information about transport properties. To
84 date, they have been utilized in computing thermal and mechanical
85 transfer of both homogeneous liquids as well as heterogeneous systems
86 such as solid-liquid
87 interfaces.\cite{garde:nl2005,garde:PhysRevLett2009,kuang:AuThl}
88
89 The reverse non-equilibrium molecular dynamics (RNEMD) methods utilize
90 additional constraints that ensure conservation of linear momentum and
91 total energy of the system while imposing the desired flux. The RNEMD
92 methods are therefore capable of sampling various thermodynamically
93 relevent ensembles, including the micro-canonical (NVE) ensemble,
94 without resorting to an external thermostat. The original approaches
95 proposed by M\"{u}ller-Plathe {\it et
96 al.}\cite{MullerPlathe:1997xw,ISI:000080382700030} utilize simple
97 momentum swapping moves for generating energy/momentum fluxes. The
98 swapping moves can also be made compatible with particles of different
99 identities. Although the swapping moves are simple to implement in
100 molecular simulations, Tenney and Maginn have shown that they create
101 nonthermal velocity distributions.\cite{Maginn:2010} Furthermore, this
102 approach is not particularly efficient for kinetic energy transfer
103 between particles of different identities, particularly when the mass
104 difference between the particles becomes significant. This problem
105 makes applying swapping-move RNEMD methods on heterogeneous
106 interfacial systems somewhat difficult.
107
108 Recently, we developed a somewhat different approach to applying
109 thermal fluxes in RNEMD simulation using a Non-Isotropic Velocity
110 Scaling (NIVS) algorithm.\cite{kuang:164101} This algorithm scales all
111 atomic velocity vectors in two separate regions of a simulated system
112 using two diagonal scaling matrices. The scaling matrices are
113 determined by solving single quartic equation which includes linear
114 momentum and kinetic energy conservation constraints as well as the
115 target thermal flux between the regions. The NIVS method is able to
116 effectively impose a wide range of kinetic energy fluxes without
117 significant perturbation to the velocity distributions away from ideal
118 Maxwell-Boltzmann distributions, even in the presence of heterogeneous
119 interfaces. We successfully applied this approach in studying the
120 interfacial thermal conductance at chemically-capped metal-solvent
121 interfaces.\cite{kuang:AuThl}
122
123 The NIVS approach works very well for preparing stable {\it thermal}
124 gradients. However, as we pointed out in the original paper, it has
125 limited application in imposing {\it linear} momentum fluxes (which
126 are required for measuring shear viscosities). The reason for this is
127 that linear momentum flux was being imposed by scaling random
128 fluctuations of the center of the velocity distributions. Repeated
129 application of the original NIVS approach resulted in temperature
130 anisotropy, i.e. the width of the velocity distributions depended on
131 coordinates perpendicular to the desired gradient direction. For this
132 reason, combining thermal and momentum fluxes was difficult with the
133 original NIVS algorithm. However, combinations of thermal and
134 velocity gradients would provide a means to simulate thermal-linear
135 coupled processes such as Soret effect in liquid flows. Therefore,
136 developing improved approaches to the scaling imposed-flux methods
137 would be useful.
138
139 In this paper, we improve the RNEMD methods by introducing a novel
140 approach to creating imposed fluxes. This approach separates the
141 shearing and scaling of the velocity distributions in different
142 spatial regions and can apply both transformations within a single
143 time step. The approach retains desirable features of previous RNEMD
144 approaches and is simpler to implement compared to the earlier NIVS
145 method. In what follows, we first present the Shearing-and-Scaling
146 (SS) RNEMD method and its implementation in a simulation. Then we
147 compare the SS-RNEMD method in bulk fluids to previous methods. We
148 also compute interfacial frictions are computed for a series of
149 heterogeneous interfaces.
150
151 \section{Methodology}
152 In an approach similar to the earlier NIVS method,\cite{kuang:164101}
153 we consider a periodic system which has been divided into a series of
154 slabs along a single axis (e.g. $z$). The unphysical thermal and
155 momentum fluxes are applied from one of the end slabs to the center
156 slab, and thus the thermal flux produces a higher temperature in the
157 center slab than in the end slab, and the momentum flux results in a
158 center slab moving along the positive $y$ axis (see Fig. \ref{cartoon}).
159 The applied fluxes can be set to negative values if the reverse
160 gradients are desired. For convenience the center slab is denoted as
161 the {\it hot} or {\it ``H''} slab, while the end slab is denoted {\it
162 ``C''} (or {\it cold}).
163
164 \begin{figure}
165 \includegraphics[width=\linewidth]{cartoon}
166 \caption{The SS-RNEMD approach we are introducing imposes unphysical
167 transfer of both momentum and kinetic energy between a ``hot'' slab
168 and a ``cold'' slab in the simulation box. The molecular system
169 responds to this imposed flux by generating both momentum and
170 temperature gradients. The slope of the gradients can then be used
171 to compute transport properties (e.g. shear viscosity and thermal
172 conductivity).}
173 \label{cartoon}
174 \end{figure}
175
176 To impose these fluxes, we periodically apply a set of operations on
177 the velocities of particles {$i$} within the cold slab and a separate
178 operation on particles {$j$} within the hot slab.
179 \begin{eqnarray}
180 \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
181 \rangle\right) + \left(\langle\vec{v}_c\rangle + \vec{a}_c\right) \\
182 \vec{v}_j & \leftarrow & h\cdot\left(\vec{v}_j - \langle\vec{v}_h
183 \rangle\right) + \left(\langle\vec{v}_h\rangle + \vec{a}_h\right)
184 \end{eqnarray}
185 where $\langle\vec{v}_c\rangle$ and $\langle\vec{v}_h\rangle$ denote
186 the instantaneous average velocity of the molecules within slabs $C$
187 and $H$ respectively. When a momentum flux $\vec{j}_z(\vec{p})$ is
188 present, these slab-averaged velocities also get corresponding
189 incremental changes ($\vec{a}_c$ and $\vec{a}_h$ respectively) that
190 are applied to all particles within each slab. The incremental
191 changes are obtained using Newton's second law:
192 \begin{eqnarray}
193 M_c \vec{a}_c & = & -\vec{j}_z(\vec{p}) \Delta t \label{eq:newton1} \\
194 M_h \vec{a}_h & = & \vec{j}_z(\vec{p}) \Delta t
195 \label{eq:newton2}
196 \end{eqnarray}
197 where $M$ denotes total mass of particles within a slab:
198 \begin{eqnarray}
199 M_c & = & \sum_{i = 1}^{N_c} m_i \\
200 M_h & = & \sum_{j = 1}^{N_h} m_j
201 \end{eqnarray}
202 and $\Delta t$ is the interval between two separate operations.
203
204 The operations in Eqs. \ref{eq:newton1} and \ref{eq:newton2} already
205 conserve the linear momentum of a periodic system. To further satisfy
206 total energy conservation as well as to impose the thermal flux $J_z$,
207 the following constraint equations must be solved for the two scaling
208 variables $c$ and $h$:
209 \begin{eqnarray}
210 K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\vec{v}_c
211 \rangle^2) + \frac{1}{2}M_c (\langle \vec{v}_c \rangle + \vec{a}_c)^2 \\
212 K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\vec{v}_h
213 \rangle^2) + \frac{1}{2}M_h (\langle \vec{v}_h \rangle + \vec{a}_h)^2.
214 \label{constraint}
215 \end{eqnarray}
216 Here $K_c$ and $K_h$ denote the translational kinetic energy of slabs
217 $C$ and $H$ respectively. These conservation equations are sufficient
218 to ensure total energy conservation, as the operations applied in our
219 method do not change the kinetic energy related to orientational
220 degrees of freedom or the potential energy of a system (as long as the
221 potential energy is independent of particle velocity).
222
223 Equations \ref{eq:newton1}-\ref{constraint} are sufficient to
224 determine the velocity scaling coefficients ($c$ and $h$) as well as
225 $\vec{a}_c$ and $\vec{a}_h$. Note that there are usually two roots
226 respectively for $c$ and $h$. However, the positive roots (which are
227 closer to 1) are chosen so that the perturbations to the system are
228 minimal. Figure \ref{method} illustrates the implementation of this
229 algorithm in an individual step.
230
231 \begin{figure}
232 \centering
233 \includegraphics[width=5in]{method}
234 \caption{Illustration of a single step implementation of the
235 algorithm. Starting with the velocity distributions for the two
236 slabs in a shearing fluid, the transformation is used to apply the
237 effect of both a thermal and a momentum flux from the ``c'' slab to
238 the ``h'' slab. As the figure shows, gaussian distributions are
239 preserved by both the scaling and shearing operations.}
240 \label{method}
241 \end{figure}
242
243 By implementing these operations at a fixed frequency, stable thermal
244 and momentum fluxes can both be applied and the corresponding
245 temperature and momentum gradients can be established.
246
247 Compared to the previous NIVS method, the SS-RNEMD approach is
248 computationally simpler in that only quadratic equations are involved
249 to determine a set of scaling coefficients, while the NIVS method
250 required solution of quartic equations. Furthermore, this method
251 implements {\it isotropic} scaling of velocities in respective slabs,
252 unlike NIVS, which required extra flexibility in the choice of scaling
253 coefficients to allow for the energy and momentum constraints. Most
254 importantly, separating the linear momentum flux from velocity scaling
255 avoids the underlying cause of the thermal anisotropy in NIVS. In
256 later sections, we demonstrate that this can improve the calculated
257 shear viscosities from RNEMD simulations.
258
259 The SS-RNEMD approach has advantages over the original momentum
260 swapping in many respects. In the swapping method, the velocity
261 vectors involved are usually very different (or the generated flux
262 would be quite small), thus the swapping tends to create strong
263 perturbations in the neighborhood of the particles involved. In
264 comparison, the SS approach distributes the flux widely to all
265 particles in a slab so that perturbations in the flux generating
266 region are minimized. Additionally, momentum swapping steps tend to
267 result in nonthermal velocity distributions when the imposed flux is
268 large and diffusion from the neighboring slabs cannot carry momentum
269 away quickly enough. In comparison, the scaling and shearing moves
270 made in the SS approach preserve the shapes of the equilibrium
271 velocity disributions (e.g. Maxwell-Boltzmann). The results presented
272 in later sections will illustrate that this is quite helpful in
273 retaining reasonable thermal distributions in a simulation.
274
275 \section{Computational Details}
276 The algorithm has been implemented in our MD simulation code,
277 OpenMD.\cite{Meineke:2005gd,openmd} We will first compare the method
278 with previous RNEMD methods and equilibrium MD in homogeneous
279 fluids (Lennard-Jones and SPC/E water). We have also used the new
280 method to simulate the interfacial friction of different heterogeneous
281 interfaces (Au (111) with organic solvents, Au(111) with SPC/E water,
282 and the Ice Ih - liquid water interface).
283
284 \subsection{Simulation Protocols}
285 The systems we investigated were set up in orthorhombic simulation
286 cells with periodic boundary conditions in all three dimensions. The
287 $z$-axes of these cells were typically quite long and served as the
288 temperature and/or momentum gradient axes. The cells were evenly
289 divided into $N$ slabs along this axis, with $N$ varying for the
290 individual system. The $x$ and $y$ axes were of similar lengths in all
291 simulations. In all cases, before introducing a nonequilibrium method
292 to establish steady thermal and/or momentum gradients, equilibration
293 simulations were run under the canonical ensemble with a Nos\'e-Hoover
294 thermostat\cite{hoover85} followed by further equilibration using
295 standard constant energy (NVE) conditions. For SPC/E water
296 simulations, isobaric-isothermal equilibrations\cite{melchionna93}
297 were performed before equilibration to reach standard densities at
298 atmospheric pressure (1 bar); for interfacial systems, similar
299 equilibrations with anisotropic box relaxations are used to relax the
300 surface tension in the $xy$ plane.
301
302 While homogeneous fluid systems can be set up with random
303 configurations, interfacial systems are typically prepared with a
304 single crystal face presented perpendicular to the $z$-axis. This
305 crystal face is aligned in the x-y plane of the periodic cell, and the
306 solvent occupies the region directly above and below a crystalline
307 slab. The preparation and equilibration of butanethiol covered
308 Au(111) surfaces, as well as the solvation and equilibration processes
309 used for these interfaces are described in detail in reference
310 \cite{kuang:AuThl}.
311
312 For the ice / liquid water interfaces, the basal surface of ice
313 lattice was first constructed. Hirsch and Ojam\"{a}e
314 \cite{doi:10.1021/jp048434u} explored the energetics of ice lattices
315 with all possible proton ordered configurations. We utilized Hirsch
316 and Ojam\"{a}e's structure 6 ($P2_12_12_1$) which is an orthorhombic
317 cell giving a proton-ordered version of Ice Ih. The basal face of ice
318 in this unit cell geometry is the $\{0~0~1\}$ face. Although
319 experimental solid/liquid coexistant temperature under normal pressure
320 are close to 273K, Bryk and Haymet's simulations of ice/liquid water
321 interfaces with different models suggest that for SPC/E, the most
322 stable interface is observed at 225$\pm$5K.\cite{bryk:10258}
323 Therefore, our ice/liquid water simulations were carried out at an
324 average temperature of 225K. Molecular translation and orientational
325 restraints were applied in the early stages of equilibration to
326 prevent melting of the ice slab. These restraints were removed during
327 NVT equilibration, well before data collection was carried out.
328
329 \subsection{Force Field Parameters}
330 For comparing the SS-RNEMD method with previous work, we retained
331 force field parameters consistent with previous simulations. Argon is
332 the Lennard-Jones fluid used here, and its results are reported in
333 reduced units for purposes of direct comparison with previous
334 calculations.
335
336 For our water simulations, we utilized the SPC/E model throughout this
337 work. Previous work for transport properties of SPC/E water model is
338 available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct
339 comparison with previous calculation methods is possible.
340
341 The Au-Au interaction parameters in all simulations are described by
342 the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
343 QSC potentials include zero-point quantum corrections and are
344 reparametrized for accurate surface energies compared to the
345 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
346 Spohr potential was adopted\cite{ISI:000167766600035} to depict
347 Au-H$_2$O interactions.
348
349 For our gold/organic liquid interfaces, the small organic molecules
350 included in our simulations are the Au surface capping agent
351 butanethiol as well as liquid hexane and liquid toluene. The
352 United-Atom
353 models\cite{TraPPE-UA.thiols,TraPPE-UA.alkanes,TraPPE-UA.alkylbenzenes}
354 for these components were used in this work for computational
355 efficiency, while maintaining good accuracy. We refer readers to our
356 previous work\cite{kuang:AuThl} for further details of these models,
357 as well as the interactions between Au and the above organic molecule
358 components.
359
360 \subsection{Thermal conductivities}
361 When the linear momentum flux $\vec{j}_z(\vec{p})$ is set to zero and
362 the target $J_z$ is non-zero, SS-RNEMD imposes kinetic energy transfer
363 between the slabs, which can be used for computation of thermal
364 conductivities. Similar to previous RNEMD methods, we assume that we
365 are in the linear response regime of the temperature gradient with
366 respect to the thermal flux. The thermal conductivity ($\lambda$) can
367 be calculated using the imposed kinetic energy flux and the measured
368 thermal gradient:
369 \begin{equation}
370 J_z = -\lambda \frac{\partial T}{\partial z}
371 \end{equation}
372 Like other imposed-flux methods, the energy flux was calculated using
373 the total non-physical energy transferred (${E_{total}}$) from slab
374 ``c'' to slab ``h'', which was recorded throughout the simulation, and
375 the total data collection time $t$:
376 \begin{equation}
377 J_z = \frac{E_{total}}{2 t L_x L_y}.
378 \end{equation}
379 Here, $L_x$ and $L_y$ denote the dimensions of the plane in a
380 simulation cell perpendicular to the thermal gradient, and a factor of
381 two in the denominator is necessary as the heat transport occurs in
382 both the $+z$ and $-z$ directions. The average temperature gradient
383 ${\langle\partial T/\partial z\rangle}$ can be obtained by a linear
384 regression of the temperature profile, which is recorded during a
385 simulation for each slab in a cell. For Lennard-Jones simulations,
386 thermal conductivities are reported in reduced units
387 (${\lambda^*=\lambda \sigma^2 m^{1/2} k_B^{-1}\varepsilon^{-1/2}}$).
388
389 \subsection{Shear viscosities}
390 Alternatively, when the linear momentum flux $\vec{j}_z(\vec{p})$ is
391 non-zero in either the $x$ or $y$ directions, the method can be used
392 to compute the shear viscosity. For isotropic systems, the direction
393 of $\vec{j}_z(\vec{p})$ on the $xy$ plane should not matter, but the
394 ability to arbitarily specify the vector direction in our method could
395 provide convenience when working with anisotropic interfaces.
396
397 In a manner similar to the thermal conductivity calculations, a linear
398 response of the momentum gradient with respect to the shear stress is
399 assumed, and the shear viscosity ($\eta$) can be obtained with the
400 imposed momentum flux (e.g. in $x$ direction) and the measured
401 velocity gradient:
402 \begin{equation}
403 j_z(p_x) = -\eta \frac{\partial v_x}{\partial z}
404 \end{equation}
405 where the flux is similarly defined:
406 \begin{equation}
407 j_z(p_x) = \frac{P_x}{2 t L_x L_y}
408 \end{equation}
409 with $P_x$ being the total non-physical momentum transferred within
410 the data collection time. Also, the averaged velocity gradient
411 ${\langle\partial v_x/\partial z\rangle}$ can be obtained using linear
412 regression of the $x$ component of the mean velocity ($\langle
413 v_x\rangle$) in each of the bins. For Lennard-Jones simulations, shear
414 viscosities are also reported in reduced units
415 (${\eta^*=\eta\sigma^2(\varepsilon m)^{-1/2}}$).
416
417 Although $J_z$ may be switched off for shear viscosity simulations,
418 the SS-RNEMD method allows the user the ability to simultaneously
419 impose both a thermal and a momentum flux during a single
420 simulation. This can create system with coincident temperature and a
421 velocity gradients. Since the viscosity is generally a function of
422 temperature, the local viscosity depends on the local temperature in
423 the fluid. Therefore, in a single simulation, viscosity at $z$
424 (corresponding to a certain $T$) can be computed with the applied
425 shear flux and the local velocity gradient (which can be obtained by
426 finite difference approximation). This means that the temperature
427 dependence of the viscosity can be mapped out in only one
428 trajectory. Results for shear viscosity computations of SPC/E water
429 will demonstrate SS-RNEMD's efficiency in this respect.
430
431 \subsection{Interfacial friction and slip length}
432 Shear stress creates a velocity gradient within bulk fluid phases, but
433 at solid-liquid interfaces, the effects of a shear stress depend on
434 the molecular details of the interface. The interfacial friction
435 coefficient, $\kappa$, relates the shear stress (e.g. along the
436 $x$-axis) with the relative fluid velocity tangent to the interface:
437 \begin{equation}
438 j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right]
439 \end{equation}
440 where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured
441 directly adjacent to the interface. Under ``stick'' boundary
442 condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads
443 to $\kappa\rightarrow\infty$. However, for ``slip'' boundary
444 conditions at a solid-liquid interface, $\kappa$ becomes finite. To
445 characterize the interfacial boundary conditions, the slip length,
446 $\delta$, is defined by the ratio of the fluid-phase viscosity to the
447 friction coefficient of the interface:
448 \begin{equation}
449 \delta = \frac{\eta}{\kappa}
450 \end{equation}
451 In ``no-slip'' or ``stick'' boundary conditions, $\delta\rightarrow
452 0$, and $\delta$ is a measure of how ``slippery'' an interface is.
453 Figure \ref{slipLength} illustrates how this quantity is defined and
454 computed for a solid-liquid interface.
455
456 \begin{figure}
457 \includegraphics[width=\linewidth]{defDelta}
458 \caption{The slip length $\delta$ can be obtained from a velocity
459 profile of a solid-liquid interface simulation, when a momentum flux
460 is applied. The data shown is for a simulated Au/hexane interface.
461 The Au crystalline region is moving as a block (lower dots), while
462 the measured velocity gradient in the hexane phase is discontinuous
463 a the interface.}
464 \label{slipLength}
465 \end{figure}
466
467 Since the method can be applied for interfaces as well as for bulk
468 materials, the shear stress is applied in an identical manner to the
469 shear viscosity computations, e.g. by applying an unphysical momentum
470 flux, $j_z(\vec{p})$. With the correct choice of $\vec{p}$ in the
471 $x-y$ plane, one can compute friction coefficients and slip lengths
472 for a number of different dragging vectors on a given slab. The
473 corresponding velocity profiles can be obtained as shown in Figure
474 \ref{slipLength}, in which the velocity gradients within the liquid
475 phase and the velocity difference at the liquid-solid interface can be
476 easily measured from saved simulation data.
477
478 \section{Results and Discussions}
479 \subsection{Lennard-Jones fluid}
480 Our orthorhombic simulation cell for the Lennard-Jones fluid has
481 identical parameters to our previous work\cite{kuang:164101} to
482 facilitate comparison. Thermal conductivities and shear viscosities
483 were computed with the new algorithm applied to the simulations. The
484 results of thermal conductivity are compared with our previous NIVS
485 algorithm. However, since the NIVS algorithm produced temperature
486 anisotropy for shear viscocity computations, these results are instead
487 compared to the previous momentum swapping approaches. Table \ref{LJ}
488 lists these values with various fluxes in reduced units.
489
490 \begin{table*}
491 \begin{minipage}{\linewidth}
492 \begin{center}
493
494 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
495 ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
496 ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
497 at various momentum fluxes. The new method yields similar
498 results to previous RNEMD methods. All results are reported in
499 reduced unit. Uncertainties are indicated in parentheses.}
500
501 \begin{tabular}{cccccc}
502 \hline\hline
503 \multicolumn{2}{c}{Momentum Exchange} &
504 \multicolumn{2}{c}{$\lambda^*$} &
505 \multicolumn{2}{c}{$\eta^*$} \\
506 \hline
507 Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
508 NIVS\cite{kuang:164101} & This work & Swapping & This work \\
509 \hline
510 0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
511 0.232 & 0.09 & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
512 0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
513 0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
514 1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
515 \hline\hline
516 \end{tabular}
517 \label{LJ}
518 \end{center}
519 \end{minipage}
520 \end{table*}
521
522 \subsubsection{Thermal conductivity}
523 Our thermal conductivity calculations with this method yields
524 comparable results to the previous NIVS algorithm. This indicates that
525 the thermal gradients introduced using this method are also close to
526 previous RNEMD methods. Simulations with moderately higher thermal
527 fluxes tend to yield more reliable thermal gradients and thus avoid
528 large errors, while overly high thermal fluxes could introduce side
529 effects such as non-linear temperature gradient response or
530 inadvertent phase transitions.
531
532 Since the scaling operation is isotropic in this method, one does not
533 need extra care to ensure temperature isotropy between the $x$, $y$
534 and $z$ axes, while for NIVS, thermal anisotropy might happen if the
535 criteria function for choosing scaling coefficients does not perform
536 as expected. Furthermore, this method avoids inadvertent concomitant
537 momentum flux when only thermal flux is imposed, which could not be
538 achieved with swapping or NIVS approaches. The thermal energy exchange
539 in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
540 or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
541 P^\alpha$) would not achieve this effect unless thermal flux vanishes
542 (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which do not contribute to
543 applying a thermal flux). In this sense, this method aids in achieving
544 minimal perturbation to a simulation while imposing a thermal flux.
545
546 \subsubsection{Shear viscosity}
547 Table \ref{LJ} also compares our shear viscosity results with the
548 momentum swapping approach. Our calculations show that our method
549 predicted similar values of shear viscosities to the momentum swapping
550 approach, as well as the velocity gradient profiles. Moderately larger
551 momentum fluxes are helpful to reduce the errors of measured velocity
552 gradients and thus the final result. However, it is pointed out that
553 the momentum swapping approach tends to produce nonthermal velocity
554 distributions.\cite{Maginn:2010}
555
556 To examine that temperature isotropy holds in simulations using our
557 method, we measured the three one-dimensional temperatures in each of
558 the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
559 temperatures were calculated after subtracting the contribution from
560 bulk velocities of the slabs. The one-dimensional temperature profiles
561 showed no observable difference between the three dimensions. This
562 ensures that isotropic scaling automatically preserves temperature
563 isotropy and that our method is useful in shear viscosity
564 computations.
565
566 \begin{figure}
567 \includegraphics[width=\linewidth]{tempXyz}
568 \caption{Unlike the previous NIVS algorithm, the new method does not
569 produce a thermal anisotropy. No temperature difference between
570 different dimensions were observed beyond the magnitude of the error
571 bars. Note that the two ``hotter'' regions are caused by the shear
572 stress, as reported by Tenney and Maginn\cite{Maginn:2010}, but not
573 an effect that only observed in our methods.}
574 \label{tempXyz}
575 \end{figure}
576
577 Furthermore, the velocity distribution profiles are tested by imposing
578 a large shear stress into the simulations. Figure \ref{vDist}
579 demonstrates how our method is able to maintain thermal velocity
580 distributions against the momentum swapping approach even under large
581 imposed fluxes. Previous swapping methods tend to deplete particles of
582 positive velocities in the negative velocity slab (``c'') and vice
583 versa in slab ``h'', where the distributions leave notchs. This
584 problematic profiles become significant when the imposed-flux becomes
585 larger and diffusions from neighboring slabs could not offset the
586 depletions. Simutaneously, abnormal peaks appear corresponding to
587 excessive particles having velocity swapped from the other slab. These
588 nonthermal distributions limit applications of the swapping approach
589 in shear stress simulations. Our method avoids the above problematic
590 distributions by altering the means of applying momentum
591 fluxes. Comparatively, velocity distributions recorded from
592 simulations with our method is so close to the ideal thermal
593 prediction that no obvious difference is shown in Figure
594 \ref{vDist}. Conclusively, our method avoids problems that occurs in
595 previous RNEMD methods and provides a useful means for shear viscosity
596 computations.
597
598 \begin{figure}
599 \includegraphics[width=\linewidth]{velDist}
600 \caption{Velocity distributions that develop under the swapping and
601 our methods at a large flux. These distributions were obtained from
602 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
603 swapping interval of 20 time steps). This is a relatively large flux
604 to demonstrate the nonthermal distributions that develop under the
605 swapping method. In comparison, distributions produced by our method
606 are very close to the ideal thermal situations.}
607 \label{vDist}
608 \end{figure}
609
610 \subsection{Bulk SPC/E water}
611 We extend our applications of thermal conductivity and shear viscosity
612 computations to a complex fluid model of SPC/E water. A simulation
613 cell with 1000 molecules was set up in the similar manner as in
614 \cite{kuang:164101}. For thermal conductivity simulations,
615 measurements were taken to compare with previous RNEMD methods; for
616 shear viscosity computations, simulations were run under a series of
617 temperatures (with corresponding pressure relaxation using the
618 isobaric-isothermal ensemble\cite{melchionna93}), and results were
619 compared to available data from EMD
620 methods\cite{10.1063/1.3330544,Medina2011}. Besides, a simulation with
621 both thermal and momentum gradient was carried out to map out shear
622 viscosity as a function of temperature to see the effectiveness and
623 accuracy our method could reach.
624
625 \subsubsection{Thermal conductivity}
626 Table \ref{spceThermal} summarizes our thermal conductivity
627 computations under different temperatures and thermal gradients, in
628 comparison to the previous NIVS results\cite{kuang:164101} and
629 experimental measurements\cite{WagnerKruse}. Note that no appreciable
630 drift of total system energy or temperature was observed when our
631 method is applied, which indicates that our algorithm conserves total
632 energy well for systems involving electrostatic interactions.
633
634 Measurements using our method established similar temperature
635 gradients to the previous NIVS method. Our simulation results are in
636 good agreement with those from previous simulations. And both methods
637 yield values in reasonable agreement with experimental
638 values. Simulations using moderately higher thermal gradient or those
639 with longer gradient axis ($z$) for measurement seem to have better
640 accuracy, from our results.
641
642 \begin{table*}
643 \begin{minipage}{\linewidth}
644 \begin{center}
645
646 \caption{Thermal conductivity of SPC/E water under various
647 imposed thermal gradients. Uncertainties are indicated in
648 parentheses.}
649
650 \begin{tabular}{ccccc}
651 \hline\hline
652 $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
653 {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
654 (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
655 Experiment\cite{WagnerKruse} \\
656 \hline
657 300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\
658 318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
659 & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
660 & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
661 twice as long.} & & \\
662 \hline\hline
663 \end{tabular}
664 \label{spceThermal}
665 \end{center}
666 \end{minipage}
667 \end{table*}
668
669 \subsubsection{Shear viscosity}
670 The improvement our method achieves for shear viscosity computations
671 enables us to apply it on SPC/E water models. The series of
672 temperatures under which our shear viscosity calculations were carried
673 out covers the liquid range under normal pressure. Our simulations
674 predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
675 (Table \ref{spceShear}). Considering subtlties such as temperature or
676 pressure/density errors in these two series of measurements, our
677 results show no significant difference from those with EMD
678 methods. Since each value reported using our method takes only one
679 single trajectory of simulation, instead of average from many
680 trajectories when using EMD, our method provides an effective means
681 for shear viscosity computations.
682
683 \begin{table*}
684 \begin{minipage}{\linewidth}
685 \begin{center}
686
687 \caption{Computed shear viscosity of SPC/E water under different
688 temperatures. Results are compared to those obtained with EMD
689 method[CITATION]. Uncertainties are indicated in parentheses.}
690
691 \begin{tabular}{cccc}
692 \hline\hline
693 $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
694 {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
695 (K) & (10$^{10}$s$^{-1}$) & This work & Previous
696 simulations\cite{Medina2011} \\
697 \hline
698 273 & 1.12 & 1.218(0.004) & 1.282(0.048) \\
699 & 1.79 & 1.140(0.012) & \\
700 303 & 2.09 & 0.646(0.008) & 0.643(0.019) \\
701 318 & 2.50 & 0.536(0.007) & \\
702 & 5.25 & 0.510(0.007) & \\
703 & 2.82 & 0.474(0.003)\footnote{Simulation with $L_z$ twice
704 as long.} & \\
705 333 & 3.10 & 0.428(0.002) & 0.421(0.008) \\
706 363 & 2.34 & 0.279(0.014) & 0.291(0.005) \\
707 & 4.26 & 0.306(0.001) & \\
708 \hline\hline
709 \end{tabular}
710 \label{spceShear}
711 \end{center}
712 \end{minipage}
713 \end{table*}
714
715 A more effective way to map out $\eta$ vs $T$ is to combine a momentum
716 flux with a thermal flux. Figure \ref{Tvxdvdz} shows the thermal and
717 velocity gradient in one such simulation. At different positions with
718 different temperatures, the velocity gradient is not a constant but
719 can be computed locally. With the data provided in Figure
720 \ref{Tvxdvdz}, a series of $\eta$ is calculated as in Figure
721 \ref{etaT} and a linear fit was performed to $\partial v_x/\partial z$
722 vs. $z$ so that the resulted $\eta$ can be present as a curve as
723 well. For comparison, other results are also mapped in the figure.
724
725 \begin{figure}
726 \includegraphics[width=\linewidth]{tvxdvdz}
727 \caption{With a combination of a thermal and a momentum flux, a
728 simulation can have both a temperature (top) and a velocity (middle)
729 gradient. Due to the thermal gradient, $\partial v_x/\partial z$ is
730 not constant but can be computed using finite difference
731 approximations (lower). These data can be used further to calculate
732 $\eta$ vs $T$ (Figure \ref{etaT}).}
733 \label{Tvxdvdz}
734 \end{figure}
735
736 From Figure \ref{etaT}, one can see that the generated curve agrees
737 well with the above RNEMD simulations at different temperatures, as
738 well as results reported using EMD
739 methods\cite{10.1063/1.3330544,Medina2011} in much of the temperature
740 range simulated. However, this curve has relatively large error in
741 lower temperature regions and has some difference in predicting $\eta$
742 near 273K. Provided that this curve only takes one trajectory to
743 generate, these results are of satisfactory efficiency and
744 accuracy. Since previous work already pointed out that the SPC/E model
745 tends to predict lower viscosity compared to experimental
746 data,\cite{Medina2011} experimental comparison are not given here.
747
748 \begin{figure}
749 \includegraphics[width=\linewidth]{etaT}
750 \caption{The curve generated by single simulation with thermal and
751 momentum gradient predicts satisfatory values in much of the
752 temperature range under test.}
753 \label{etaT}
754 \end{figure}
755
756 \subsection{Interfacial frictions and slip lengths}
757 Another attractive aspect of our method is the ability to apply
758 momentum and/or thermal flux in nonhomogeneous systems, where
759 molecules of different identities (or phases) are segregated in
760 different regions. We have previously studied the interfacial thermal
761 transport of a series of metal gold-liquid
762 surfaces\cite{kuang:164101,kuang:AuThl}, and would like to further
763 investigate the relationship between this phenomenon and the
764 interfacial frictions.
765
766 Table \ref{etaKappaDelta} includes these computations and previous
767 calculations of corresponding interfacial thermal conductance. For
768 bare Au(111) surfaces, slip boundary conditions were observed for both
769 organic and aqueous liquid phases, corresponding to previously
770 computed low interfacial thermal conductance. In comparison, the
771 butanethiol covered Au(111) surface appeared to be sticky to the
772 organic liquid layers in our simulations. We have reported conductance
773 enhancement effect for this surface capping agent,\cite{kuang:AuThl}
774 and these observations have a qualitative agreement with the thermal
775 conductance results. This agreement also supports discussions on the
776 relationship between surface wetting and slip effect and thermal
777 conductance of the
778 interface.\cite{PhysRevLett.82.4671,doi:10.1080/0026897031000068578,garde:PhysRevLett2009}
779
780 \begin{table*}
781 \begin{minipage}{\linewidth}
782 \begin{center}
783
784 \caption{Computed interfacial friction coefficient values for
785 interfaces with various components for liquid and solid
786 phase. Error estimates are indicated in parentheses.}
787
788 \begin{tabular}{llcccccc}
789 \hline\hline
790 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
791 & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
792 \cite{kuang:164101}.} \\
793 surface & molecules & K & MPa & mPa$\cdot$s &
794 10$^4$Pa$\cdot$s/m & nm & MW/m$^2$/K \\
795 \hline
796 Au(111) & hexane & 200 & 1.08 & 0.197(0.009) & 5.30(0.36) &
797 3.72 & 46.5 \\
798 & & & 2.15 & 0.141(0.002) & 5.31(0.26) &
799 2.76 & \\
800 Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.286(0.019) & $\infty$
801 & 0 & 131 \\
802 & & & 5.39 & 0.320(0.006) & $\infty$
803 & 0 & \\
804 \hline
805 Au(111) & toluene & 200 & 1.08 & 0.722(0.035) & 15.7(0.7) &
806 4.60 & 70.1 \\
807 & & & 2.16 & 0.544(0.030) & 11.2(0.5) &
808 4.86 & \\
809 Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.980(0.057) &
810 $\infty$ & 0 & 187 \\
811 & & & 10.8 & 0.995(0.005) &
812 $\infty$ & 0 & \\
813 \hline
814 Au(111) & water & 300 & 1.08 & 0.399(0.050) & 1.928(0.022) &
815 20.7 & 1.65 \\
816 & & & 2.16 & 0.794(0.255) & 1.895(0.003) &
817 41.9 & \\
818 \hline
819 ice(basal) & water & 225 & 19.4 & 15.8(0.2) & $\infty$ & 0 & \\
820 \hline\hline
821 \end{tabular}
822 \label{etaKappaDelta}
823 \end{center}
824 \end{minipage}
825 \end{table*}
826
827 An interesting effect alongside the surface friction change is
828 observed on the shear viscosity of liquids in the regions close to the
829 solid surface. In our results, $\eta$ measured near a ``slip'' surface
830 tends to be smaller than that near a ``stick'' surface. This may
831 suggest the influence from an interface on the dynamic properties of
832 liquid within its neighbor regions. It is known that diffusions of
833 solid particles in liquid phase is affected by their surface
834 conditions (stick or slip boundary).\cite{10.1063/1.1610442} Our
835 observations could provide a support to this phenomenon.
836
837 In addition to these previously studied interfaces, we attempt to
838 construct ice-water interfaces and the basal plane of ice lattice was
839 studied here. In contrast to the Au(111)/water interface, where the
840 friction coefficient is substantially small and large slip effect
841 presents, the ice/liquid water interface demonstrates strong
842 solid-liquid interactions and appears to be sticky. The supercooled
843 liquid phase is an order of magnitude more viscous than measurements
844 in previous section. It would be of interst to investigate the effect
845 of different ice lattice planes (such as prism and other surfaces) on
846 interfacial friction and the corresponding liquid viscosity.
847
848 \section{Conclusions}
849 Our simulations demonstrate the validity of our method in RNEMD
850 computations of thermal conductivity and shear viscosity in atomic and
851 molecular liquids. Our method maintains thermal velocity distributions
852 and avoids thermal anisotropy in previous NIVS shear stress
853 simulations, as well as retains attractive features of previous RNEMD
854 methods. There is no {\it a priori} restrictions to the method to be
855 applied in various ensembles, so prospective applications to
856 extended-system methods are possible.
857
858 Our method is capable of effectively imposing thermal and/or momentum
859 flux accross an interface. This facilitates studies that relates
860 dynamic property measurements to the chemical details of an
861 interface. Therefore, investigations can be carried out to
862 characterize interfacial interactions using the method.
863
864 Another attractive feature of our method is the ability of
865 simultaneously introducing thermal and momentum gradients in a
866 system. This facilitates us to effectively map out the shear viscosity
867 with respect to a range of temperature in single trajectory of
868 simulation with satisafactory accuracy. Complex systems that involve
869 thermal and momentum gradients might potentially benefit from
870 this. For example, the Soret effects under a velocity gradient might
871 be models of interest to purification and separation researches.
872
873 \section{Acknowledgments}
874 Support for this project was provided by the National Science
875 Foundation under grant CHE-0848243. Computational time was provided by
876 the Center for Research Computing (CRC) at the University of Notre
877 Dame.
878
879 \newpage
880
881 \bibliography{stokes}
882
883 \end{doublespace}
884 \end{document}