ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/stokes/stokes.tex
Revision: 3788
Committed: Thu Dec 15 16:19:36 2011 UTC (12 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 44224 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 simulating 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 and interfacial systems. This method conserves both
50 the linear momentum and total energy of the system and improves
51 previous reverse non-equilibrium molecular dynamics (RNEMD) methods
52 while retaining equilibrium thermal velocity distributions in each
53 region of the system. The new method avoids the thermal anisotropy
54 produced by previous methods by using isotropic velocity scaling and
55 shearing (VSS) on all of the molecules in specific regions. To test
56 the 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 velocity shearing and
146 scaling (VSS) RNEMD method and its implementation in a
147 simulation. Then we compare the VSS-RNEMD method in bulk fluids to
148 previous methods. We also compute interfacial frictions are computed
149 for a series of 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 VSS-RNEMD approach imposes unphysical transfer of both
167 linear momentum and kinetic energy between a ``hot'' slab and a
168 ``cold'' slab in the simulation box. The system responds to this
169 imposed flux by generating both momentum and temperature gradients.
170 The slope of the gradients can then be used to compute transport
171 properties (e.g. shear viscosity and thermal conductivity).}
172 \label{cartoon}
173 \end{figure}
174
175 To impose these fluxes, we periodically apply a set of operations on
176 the velocities of particles {$i$} within the cold slab and a separate
177 operation on particles {$j$} within the hot slab.
178 \begin{eqnarray}
179 \vec{v}_i & \leftarrow & c\cdot\left(\vec{v}_i - \langle\vec{v}_c
180 \rangle\right) + \left(\langle\vec{v}_c\rangle +
181 \vec{a}_c\right) \label{eq:xformc} \\
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) \label{eq:xformh}
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{A single step implementation of the VSS algorithm starts with
235 the velocity distributions for the two slabs in a shearing fluid
236 (solid lines). Equations \ref{eq:xformc} and \ref{eq:xformh} are
237 used to scale and shear velocities in both slabs to mimic the effect
238 of both a thermal and a momentum flux. 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 VSS-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 VSS-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 VSS 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 VSS 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 that produces a proton-ordered version of ice Ih. The basal face
318 of ice 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$5K.\cite{bryk:10258}
323 Therefore, our ice/liquid water simulations were carried out at an
324 average temperature of 225K. Molecular translation and orientational
325 restraints were applied in the early stages of equilibration to
326 prevent melting of the ice slab. These restraints were removed during
327 NVT equilibration, well before data collection was carried out.
328
329 \subsection{Force Field Parameters}
330 For comparing the VSS-RNEMD method with previous work, we retained
331 force field parameters consistent with previous simulations. Argon is
332 the Lennard-Jones fluid used here, and its results are reported in
333 reduced units for purposes of direct comparison with previous
334 calculations.
335
336 For our water simulations, we utilized the SPC/E model throughout this
337 work. Previous work for transport properties of SPC/E water model is
338 available\cite{Bedrov:2000,10.1063/1.3330544,Medina2011} so direct
339 comparison with previous calculation methods is possible.
340
341 The Au-Au interaction parameters in all simulations are described by
342 the quantum Sutton-Chen (QSC) formulation.\cite{PhysRevB.59.3527} The
343 QSC potentials include zero-point quantum corrections and are
344 reparametrized for accurate surface energies compared to the
345 Sutton-Chen potentials.\cite{Chen90} For gold/water interfaces, the
346 Spohr potential was adopted\cite{ISI:000167766600035} to depict
347 Au-H$_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, VSS-RNEMD imposes kinetic energy
363 transfer between the slabs, which can be used for computation of
364 thermal conductivities. As with other 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 then be calculated using the imposed kinetic energy flux and the
368 measured 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 {\it C} to slab {\it 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 the
380 simulation cell that is perpendicular to the thermal gradient, and a
381 factor of two in the denominator is necessary as the heat transport
382 occurs in both the $+z$ and $-z$ directions. The average temperature
383 gradient ${\langle\partial T/\partial z\rangle}$ can be obtained by a
384 linear regression of the temperature profile, which is recorded during
385 a 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 be convenient 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 VSS-RNEMD method allows the user the ability to {\it
419 simultaneously} impose both a thermal and a momentum flux during a
420 single simulation. This creates configurations with coincident
421 temperature and a velocity gradients. Since the viscosity is
422 generally a function of temperature, the local viscosity depends on
423 the local temperature in the fluid. Therefore, in a single simulation,
424 viscosity at $z$ (corresponding to a certain $T$) can be computed with
425 the applied shear flux and the local velocity gradient (which can be
426 obtained by finite difference approximation). This means that the
427 temperature 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 VSS-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 velocity of the fluid tangent to the
437 interface:
438 \begin{equation}
439 j_z(p_x) = \kappa \left[v_x(fluid) - v_x(solid)\right]
440 \end{equation}
441 where $v_x(fluid)$ and $v_x(solid)$ are the velocities measured
442 directly adjacent to the interface. Under ``stick'' boundary
443 condition, $\Delta v_x|_\mathrm{interface} \rightarrow 0$, which leads
444 to $\kappa\rightarrow\infty$. However, for ``slip'' boundary
445 conditions at a solid-liquid interface, $\kappa$ becomes finite. To
446 characterize the interfacial boundary conditions, the slip length,
447 $\delta$, is defined by the ratio of the fluid-phase viscosity to the
448 friction coefficient of the interface:
449 \begin{equation}
450 \delta = \frac{\eta}{\kappa}
451 \end{equation}
452 In ``stick'' boundary conditions, $\delta\rightarrow 0$, so $\delta$
453 is a measure of how ``slippery'' an interface is. Figure
454 \ref{slipLength} illustrates how this quantity is defined and computed
455 for a solid-liquid interface.
456
457 \begin{figure}
458 \includegraphics[width=\linewidth]{defDelta}
459 \caption{The slip length ($\delta$) can be obtained from a velocity
460 profile along an axis perpendicular to the interface. The data shown
461 is for an Au / hexane interface -- the crystalline region (Au) is
462 moving as a block (lower dots), while the measured velocity gradient
463 in the hexane phase is discontinuous at the interface.}
464 \label{slipLength}
465 \end{figure}
466
467 Since the VSS-RNEMD method can be applied for interfaces as well as
468 for bulk materials, the shear stress is applied in an identical manner
469 to the shear viscosity computations, e.g. by applying an unphysical
470 momentum flux, $j_z(\vec{p})$. With the correct choice of $\vec{p}$
471 in the $x-y$ plane, one can compute friction coefficients and slip
472 lengths for a number of different dragging vectors on a given slab.
473 The 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 previous work to facilitate
482 comparison.\cite{kuang:164101} Thermal conductivities and shear
483 viscosities were computed with the new algorithm applied to the
484 simulations. The results of thermal conductivity are compared with our
485 previous NIVS algorithm. However, since the NIVS algorithm produces
486 temperature anisotropy for shear viscocity computations, these results
487 are instead compared to the previous momentum swapping
488 approaches. Table \ref{LJ} lists these values with various fluxes in
489 reduced units.
490
491 \begin{table*}
492 \begin{minipage}{\linewidth}
493 \begin{center}
494
495 \caption{Thermal conductivity ($\lambda^*$) and shear viscosity
496 ($\eta^*$) (in reduced units) of a Lennard-Jones fluid at
497 ${\langle T^* \rangle = 0.72}$ and ${\rho^* = 0.85}$ computed
498 at various momentum fluxes. The new method yields similar
499 results to previous RNEMD methods. All results are reported in
500 reduced unit. Uncertainties are indicated in parentheses.}
501
502 \begin{tabular}{cccccc}
503 \hline\hline
504 \multicolumn{2}{c}{Momentum Exchange} &
505 \multicolumn{2}{c}{$\lambda^*$} &
506 \multicolumn{2}{c}{$\eta^*$} \\
507 \hline
508 Swap Interval & Equivalent $J_z^*$ or $j_z^*(p_x)$ &
509 NIVS\cite{kuang:164101} & This work & Swapping & This work \\
510 \hline
511 0.116 & 0.16 & 7.30(0.10) & 6.25(0.23) & 3.57(0.06) & 3.53(0.16)\\
512 0.232 & 0.09 & 6.95(0.09) & 8.02(0.56) & 3.64(0.05) & 3.43(0.17)\\
513 0.463 & 0.047 & 7.19(0.07) & 6.48(0.15) & 3.52(0.16) & 3.51(0.08)\\
514 0.926 & 0.024 & 7.19(0.28) & 7.02(0.13) & 3.72(0.05) & 3.79(0.11)\\
515 1.158 & 0.019 & 7.98(0.33) & 7.37(0.23) & 3.42(0.06) & 3.53(0.06)\\
516 \hline\hline
517 \end{tabular}
518 \label{LJ}
519 \end{center}
520 \end{minipage}
521 \end{table*}
522
523 \subsubsection{Thermal conductivity}
524 Our thermal conductivity calculations with this method yields
525 comparable results to the previous NIVS algorithm. This indicates that
526 the thermal gradients introduced using this method are also close to
527 previous RNEMD methods. Simulations with moderately higher thermal
528 fluxes tend to yield more reliable thermal gradients and thus avoid
529 large errors, while overly high thermal fluxes could introduce side
530 effects such as non-linear temperature gradient response or
531 inadvertent phase transitions.
532
533 Since the scaling operation is isotropic in this method, one does not
534 need extra care to ensure temperature isotropy between the $x$, $y$
535 and $z$ axes, while for NIVS, thermal anisotropy might happen if the
536 criteria function for choosing scaling coefficients does not perform
537 as expected. Furthermore, this method avoids inadvertent concomitant
538 momentum flux when only thermal flux is imposed, which could not be
539 achieved with swapping or NIVS approaches. The thermal energy exchange
540 in swapping ($\vec{p}_i$ in slab ``c'' with $\vec{p}_j$ in slab ``h'')
541 or NIVS (total slab momentum conponemts $P^\alpha$ scaled to $\alpha
542 P^\alpha$) would not achieve this effect unless thermal flux vanishes
543 (i.e. $\vec{p}_i=\vec{p}_j$ or $\alpha=1$ which do not contribute to
544 applying a thermal flux). In this sense, this method aids in achieving
545 minimal perturbation to a simulation while imposing a thermal flux.
546
547 \subsubsection{Shear viscosity}
548 Table \ref{LJ} also compares our shear viscosity results with the
549 momentum swapping approach. Our calculations show that our method
550 predicted similar values of shear viscosities to the momentum swapping
551 approach, as well as the velocity gradient profiles. Moderately larger
552 momentum fluxes are helpful to reduce the errors of measured velocity
553 gradients and thus the final result. However, it is pointed out that
554 the momentum swapping approach tends to produce nonthermal velocity
555 distributions.\cite{Maginn:2010}
556
557 To examine that temperature isotropy holds in simulations using our
558 method, we measured the three one-dimensional temperatures in each of
559 the slabs (Figure \ref{tempXyz}). Note that the $x$-dimensional
560 temperatures were calculated after subtracting the contribution from
561 bulk velocities of the slabs. The one-dimensional temperature profiles
562 showed no observable difference between the three dimensions. This
563 ensures that isotropic scaling automatically preserves temperature
564 isotropy and that our method is useful in shear viscosity
565 computations.
566
567 \begin{figure}
568 \includegraphics[width=\linewidth]{tempXyz}
569 \caption{Unlike the previous NIVS algorithm, the VSS method does not
570 produce thermal anisotropy under shearing stress. Temperature
571 differences along the different box axes were significantly smaller
572 than the error bars. Note that the two regions of elevated
573 temperature are caused by the shear stress. This is the same
574 frictional heating reported previously by Tenney and
575 Maginn.\cite{Maginn:2010}}
576 \label{tempXyz}
577 \end{figure}
578
579 Furthermore, the velocity distribution profiles are tested by imposing
580 a large shear stress into the simulations. Figure \ref{vDist}
581 demonstrates how our method is able to maintain thermal velocity
582 distributions against the momentum swapping approach even under large
583 imposed fluxes. Previous swapping methods tend to deplete particles of
584 positive velocities in the negative velocity slab (``c'') and vice
585 versa in slab ``h'', where the distributions leave notchs. This
586 problematic profiles become significant when the imposed-flux becomes
587 larger and diffusions from neighboring slabs could not offset the
588 depletions. Simutaneously, abnormal peaks appear corresponding to
589 excessive particles having velocity swapped from the other slab. These
590 nonthermal distributions limit applications of the swapping approach
591 in shear stress simulations. Our method avoids the above problematic
592 distributions by altering the means of applying momentum
593 fluxes. Comparatively, velocity distributions recorded from
594 simulations with our method is so close to the ideal thermal
595 prediction that no obvious difference is shown in Figure
596 \ref{vDist}. Conclusively, our method avoids problems that occurs in
597 previous RNEMD methods and provides a useful means for shear viscosity
598 computations.
599
600 \begin{figure}
601 \centering
602 \includegraphics[width=5.5in]{velDist}
603 \caption{Velocity distributions that develop under VSS more closely
604 resemble ideal Maxwell-Boltzmann distributions than earlier
605 momentum-swapping RNEMD approaches. This data was obtained from
606 Lennard-Jones simulations with $j_z^*(p_x)\sim 0.4$ (equivalent to a
607 swapping interval of 20 time steps). This is a relatively large flux
608 which demonstrates some of the non-thermal velocity distributions
609 that can develop under swapping RNEMD.}
610 \label{vDist}
611 \end{figure}
612
613 \subsection{Bulk SPC/E water}
614 We extend our applications of thermal conductivity and shear viscosity
615 computations to a complex fluid model of SPC/E water. A simulation
616 cell with 1000 molecules was set up in the similar manner as in
617 \cite{kuang:164101}. For thermal conductivity simulations,
618 measurements were taken to compare with previous RNEMD methods; for
619 shear viscosity computations, simulations were run under a series of
620 temperatures (with corresponding pressure relaxation using the
621 isobaric-isothermal ensemble\cite{melchionna93}), and results were
622 compared to available data from EMD
623 methods\cite{10.1063/1.3330544,Medina2011}. Besides, a simulation with
624 both thermal and momentum gradient was carried out to map out shear
625 viscosity as a function of temperature to see the effectiveness and
626 accuracy our method could reach.
627
628 \subsubsection{Thermal conductivity}
629 Table \ref{spceThermal} summarizes our thermal conductivity
630 computations under different temperatures and thermal gradients, in
631 comparison to the previous NIVS results\cite{kuang:164101} and
632 experimental measurements\cite{WagnerKruse}. Note that no appreciable
633 drift of total system energy or temperature was observed when our
634 method is applied, which indicates that our algorithm conserves total
635 energy well for systems involving electrostatic interactions.
636
637 Measurements using our method established similar temperature
638 gradients to the previous NIVS method. Our simulation results are in
639 good agreement with those from previous simulations. And both methods
640 yield values in reasonable agreement with experimental
641 values. Simulations using moderately higher thermal gradient or those
642 with longer gradient axis ($z$) for measurement seem to have better
643 accuracy, from our results.
644
645 \begin{table*}
646 \begin{minipage}{\linewidth}
647 \begin{center}
648
649 \caption{Thermal conductivity of SPC/E water under various
650 imposed thermal gradients. Uncertainties are indicated in
651 parentheses.}
652
653 \begin{tabular}{ccccc}
654 \hline\hline
655 $\langle T\rangle$ & $\langle dT/dz\rangle$ & \multicolumn{3}{c}
656 {$\lambda (\mathrm{W m}^{-1} \mathrm{K}^{-1})$} \\
657 (K) & (K/\AA) & This work & Previous NIVS\cite{kuang:164101} &
658 Experiment\cite{WagnerKruse} \\
659 \hline
660 300 & 0.8 & 0.815(0.027) & 0.770(0.008) & 0.61 \\
661 318 & 0.8 & 0.801(0.024) & 0.750(0.032) & 0.64 \\
662 & 1.6 & 0.766(0.007) & 0.778(0.019) & \\
663 & 0.8 & 0.786(0.009)\footnote{Simulation with $L_z$
664 twice as long.} & & \\
665 \hline\hline
666 \end{tabular}
667 \label{spceThermal}
668 \end{center}
669 \end{minipage}
670 \end{table*}
671
672 \subsubsection{Shear viscosity}
673 The improvement our method achieves for shear viscosity computations
674 enables us to apply it on SPC/E water models. The series of
675 temperatures under which our shear viscosity calculations were carried
676 out covers the liquid range under normal pressure. Our simulations
677 predict a similar trend of $\eta$ vs. $T$ to EMD results we refer to
678 (Table \ref{spceShear}). Considering subtlties such as temperature or
679 pressure/density errors in these two series of measurements, our
680 results show no significant difference from those with EMD
681 methods. Since each value reported using our method takes only one
682 single trajectory of simulation, instead of average from many
683 trajectories when using EMD, our method provides an effective means
684 for shear viscosity computations.
685
686 \begin{table*}
687 \begin{minipage}{\linewidth}
688 \begin{center}
689
690 \caption{Computed shear viscosity of SPC/E water under different
691 temperatures. Results are compared to those obtained with EMD
692 method[CITATION]. Uncertainties are indicated in parentheses.}
693
694 \begin{tabular}{cccc}
695 \hline\hline
696 $T$ & $\partial v_x/\partial z$ & \multicolumn{2}{c}
697 {$\eta (\mathrm{mPa}\cdot\mathrm{s})$} \\
698 (K) & (10$^{10}$s$^{-1}$) & This work & Previous
699 simulations\cite{Medina2011} \\
700 \hline
701 273 & 1.12 & 1.218(0.004) & 1.282(0.048) \\
702 & 1.79 & 1.140(0.012) & \\
703 303 & 2.09 & 0.646(0.008) & 0.643(0.019) \\
704 318 & 2.50 & 0.536(0.007) & \\
705 & 5.25 & 0.510(0.007) & \\
706 & 2.82 & 0.474(0.003)\footnote{Simulation with $L_z$ twice
707 as long.} & \\
708 333 & 3.10 & 0.428(0.002) & 0.421(0.008) \\
709 363 & 2.34 & 0.279(0.014) & 0.291(0.005) \\
710 & 4.26 & 0.306(0.001) & \\
711 \hline\hline
712 \end{tabular}
713 \label{spceShear}
714 \end{center}
715 \end{minipage}
716 \end{table*}
717
718 A more effective way to map out $\eta$ vs $T$ is to combine a momentum
719 flux with a thermal flux. Figure \ref{Tvxdvdz} shows the thermal and
720 velocity gradient in one such simulation. At different positions with
721 different temperatures, the velocity gradient is not a constant but
722 can be computed locally. With the data provided in Figure
723 \ref{Tvxdvdz}, a series of $\eta$ is calculated as in Figure
724 \ref{etaT} and a linear fit was performed to $\partial v_x/\partial z$
725 vs. $z$ so that the resulted $\eta$ can be present as a curve as
726 well. For comparison, other results are also mapped in the figure.
727
728 \begin{figure}
729 \centering
730 \includegraphics[width=5.5in]{tvxdvdz}
731 \caption{With a combination of a thermal and a momentum flux, a
732 simulation can exhibit both a temperature (top) and a velocity
733 (middle) gradient. $\partial v_x/\partial z$ depends on the local
734 temperature, so it varies along the gradient axis. This derivative
735 can be computed using finite difference approximations (lower) and
736 can be used to calculate $\eta$ vs $T$ (Figure \ref{etaT}).}
737 \label{Tvxdvdz}
738 \end{figure}
739
740 From Figure \ref{etaT}, one can see that the generated curve agrees
741 well with the above RNEMD simulations at different temperatures, as
742 well as results reported using EMD
743 methods\cite{10.1063/1.3330544,Medina2011} in much of the temperature
744 range simulated. However, this curve has relatively large error in
745 lower temperature regions and has some difference in predicting $\eta$
746 near 273K. Provided that this curve only takes one trajectory to
747 generate, these results are of satisfactory efficiency and
748 accuracy. Since previous work already pointed out that the SPC/E model
749 tends to predict lower viscosity compared to experimental
750 data,\cite{Medina2011} experimental comparison are not given here.
751
752 \begin{figure}
753 \includegraphics[width=\linewidth]{etaT}
754 \caption{The temperature dependence of the shear viscosity generated
755 by a {\it single} VSS-RNEMD simulation. Applying both thermal and
756 momentum gradient predicts reasonable values in much of the
757 temperature range being tested.}
758 \label{etaT}
759 \end{figure}
760
761 \subsection{Interfacial frictions and slip lengths}
762 Another attractive aspect of our method is the ability to apply
763 momentum and/or thermal flux in nonhomogeneous systems, where
764 molecules of different identities (or phases) are segregated in
765 different regions. We have previously studied the interfacial thermal
766 transport of a series of metal gold-liquid
767 surfaces\cite{kuang:164101,kuang:AuThl}, and would like to further
768 investigate the relationship between this phenomenon and the
769 interfacial frictions.
770
771 Table \ref{etaKappaDelta} includes these computations and previous
772 calculations of corresponding interfacial thermal conductance. For
773 bare Au(111) surfaces, slip boundary conditions were observed for both
774 organic and aqueous liquid phases, corresponding to previously
775 computed low interfacial thermal conductance. In comparison, the
776 butanethiol covered Au(111) surface appeared to be sticky to the
777 organic liquid layers in our simulations. We have reported conductance
778 enhancement effect for this surface capping agent,\cite{kuang:AuThl}
779 and these observations have a qualitative agreement with the thermal
780 conductance results. This agreement also supports discussions on the
781 relationship between surface wetting and slip effect and thermal
782 conductance of the
783 interface.\cite{PhysRevLett.82.4671,doi:10.1080/0026897031000068578,garde:PhysRevLett2009}
784
785 \begin{table*}
786 \begin{minipage}{\linewidth}
787 \begin{center}
788
789 \caption{Computed interfacial friction coefficient values for
790 interfaces with various components for liquid and solid
791 phase. Error estimates are indicated in parentheses.}
792
793 \begin{tabular}{llcccccc}
794 \hline\hline
795 Solid & Liquid & $T$ & $j_z(p_x)$ & $\eta_{liquid}$ & $\kappa$
796 & $\delta$ & $G$\footnote{References \cite{kuang:AuThl} and
797 \cite{kuang:164101}.} \\
798 surface & molecules & K & MPa & mPa$\cdot$s &
799 10$^4$Pa$\cdot$s/m & nm & MW/m$^2$/K \\
800 \hline
801 Au(111) & hexane & 200 & 1.08 & 0.197(0.009) & 5.30(0.36) &
802 3.72 & 46.5 \\
803 & & & 2.15 & 0.141(0.002) & 5.31(0.26) &
804 2.76 & \\
805 Au-SC$_4$H$_9$ & hexane & 200 & 2.16 & 0.286(0.019) & $\infty$
806 & 0 & 131 \\
807 & & & 5.39 & 0.320(0.006) & $\infty$
808 & 0 & \\
809 \hline
810 Au(111) & toluene & 200 & 1.08 & 0.722(0.035) & 15.7(0.7) &
811 4.60 & 70.1 \\
812 & & & 2.16 & 0.544(0.030) & 11.2(0.5) &
813 4.86 & \\
814 Au-SC$_4$H$_9$ & toluene & 200 & 5.39 & 0.980(0.057) &
815 $\infty$ & 0 & 187 \\
816 & & & 10.8 & 0.995(0.005) &
817 $\infty$ & 0 & \\
818 \hline
819 Au(111) & water & 300 & 1.08 & 0.399(0.050) & 1.928(0.022) &
820 20.7 & 1.65 \\
821 & & & 2.16 & 0.794(0.255) & 1.895(0.003) &
822 41.9 & \\
823 \hline
824 ice(basal) & water & 225 & 19.4 & 15.8(0.2) & $\infty$ & 0 & \\
825 \hline\hline
826 \end{tabular}
827 \label{etaKappaDelta}
828 \end{center}
829 \end{minipage}
830 \end{table*}
831
832 An interesting effect alongside the surface friction change is
833 observed on the shear viscosity of liquids in the regions close to the
834 solid surface. In our results, $\eta$ measured near a ``slip'' surface
835 tends to be smaller than that near a ``stick'' surface. This may
836 suggest the influence from an interface on the dynamic properties of
837 liquid within its neighbor regions. It is known that diffusions of
838 solid particles in liquid phase is affected by their surface
839 conditions (stick or slip boundary).\cite{10.1063/1.1610442} Our
840 observations could provide a support to this phenomenon.
841
842 In addition to these previously studied interfaces, we attempt to
843 construct ice-water interfaces and the basal plane of ice lattice was
844 studied here. In contrast to the Au(111)/water interface, where the
845 friction coefficient is substantially small and large slip effect
846 presents, the ice/liquid water interface demonstrates strong
847 solid-liquid interactions and appears to be sticky. The supercooled
848 liquid phase is an order of magnitude more viscous than measurements
849 in previous section. It would be of interst to investigate the effect
850 of different ice lattice planes (such as prism and other surfaces) on
851 interfacial friction and the corresponding liquid viscosity.
852
853 \section{Conclusions}
854 Our simulations demonstrate the validity of our method in RNEMD
855 computations of thermal conductivity and shear viscosity in atomic and
856 molecular liquids. Our method maintains thermal velocity distributions
857 and avoids thermal anisotropy in previous NIVS shear stress
858 simulations, as well as retains attractive features of previous RNEMD
859 methods. There is no {\it a priori} restrictions to the method to be
860 applied in various ensembles, so prospective applications to
861 extended-system methods are possible.
862
863 Our method is capable of effectively imposing thermal and/or momentum
864 flux accross an interface. This facilitates studies that relates
865 dynamic property measurements to the chemical details of an
866 interface. Therefore, investigations can be carried out to
867 characterize interfacial interactions using the method.
868
869 Another attractive feature of our method is the ability of
870 simultaneously introducing thermal and momentum gradients in a
871 system. This facilitates us to effectively map out the shear viscosity
872 with respect to a range of temperature in single trajectory of
873 simulation with satisafactory accuracy. Complex systems that involve
874 thermal and momentum gradients might potentially benefit from
875 this. For example, the Soret effects under a velocity gradient might
876 be models of interest to purification and separation researches.
877
878 \section{Acknowledgments}
879 Support for this project was provided by the National Science
880 Foundation under grant CHE-0848243. Computational time was provided by
881 the Center for Research Computing (CRC) at the University of Notre
882 Dame.
883
884 \newpage
885
886 \bibliography{stokes}
887
888 \end{doublespace}
889 \end{document}