1: \documentclass[aps,pre,preprint,showpacs,showkeys,groupedaddress]{revtex4}
2: %\documentclass[aps,pre,twocolumn,groupedaddress,showpacs,showkeys]{revtex4}
3: \newcommand{\W}{\columnwidth}
4: \usepackage{graphicx}
5: \usepackage{amsmath}
6: \usepackage{color}
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %%%% SHOW LABELS
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: %\usepackage[notref,notcite]{showkeys}
11: %\oddsidemargin=2cm
12: %\textwidth=13cm
13: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
14: \bibliographystyle{apsrev}
15: \begin{document}
16: \title{Lattice-Boltzmann Method for Non-Newtonian Fluid Flows}
17: \author{Susana Gabbanelli}
18: \affiliation{Departamento de Matem\'atica y Grupo de Medios Porosos,
19: Facultad de Ingenier\'{\i}a, Universidad de Buenos Aires}
20: \author{German Drazer}
21: \email{drazer@mailaps.org}
22: \author{Joel Koplik}
23: \email{koplik@sci.ccny.cuny.edu}
24: \affiliation{Benjamin Levich Institute and Department of Physics,
25: City College of the City University of New York, New York, NY 10031}
26: \date{\today}
27: \begin{abstract}
28: We study an {\it ad hoc} extension of the Lattice-Boltzmann method
29: that allows the simulation of non-Newtonian fluids described by
30: generalized Newtonian models. We extensively test the accuracy of the
31: method for the case of shear-thinning and shear-thickening truncated
32: power-law fluids in the parallel plate geometry, and show that the
33: relative error compared to analytical solutions decays approximately
34: linear with the lattice resolution. Finally, we also tested the method
35: in the reentrant-flow geometry, in which the shear-rate is no-longer a
36: scalar and the presence of two singular points requires high accuracy
37: in order to obtain satisfactory resolution in the local stress near
38: these points. In this geometry, we also found excellent agreement with
39: the solutions obtained by standard finite-element methods,
40: and the agreement improves with higher lattice resolution.
41: \end{abstract}
42: \pacs{47.11+j, 47.50.+d, 47.10.+g, 02.70.Rr}
43: \keywords{Lattice-Boltzmann, non-Newtonian, power-law}
44: \maketitle
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47: %% SECTION Introduction
48: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: \section{Introduction}
51: \label{intro}
52:
53: Since its origin, more than 15 years ago, the Lattice Boltzmann Method (LBM) has proved to be
54: a powerful numerical technique for the simulation of single and multi-phase fluid flows in complex
55: geometries. In fact, the LBM has been successfully applied to different problems in fluid dynamics
56: and the interest on the method has grown rapidly in recent years. The LBM
57: is particularly suited for complex geometries and interfacial dynamics, and its initial
58: applications included transport in porous media, multiphase and multicomponent fluid flows \cite{ChenD98}.
59: It was then adapted by Ladd and others to simulate particle-fluid suspensions \cite{LaddV01}.
60: It has also been applied to high Reynolds number incompressible flows and turbulence, and the
61: implementation of thermal and compressible schemes is being actively pursued \cite{YuMLW03}.
62: One advantage of the LBM is that data communications between nodes is always local, which makes the
63: method extremely efficient for large-scale, massively-parallel computations (see Ref. \cite{NourgalievDTJ03}
64: for an interesting discussion on the LBM capabilities compared to the existing continuum-based computational
65: fluid dynamics methods). Another property of the LBM that has lately attracted considerable attention is the
66: microscopic origin of its mesoscopic kinetic equations, which could therefore
67: readily incorporate molecular level interactions. This makes the LBM very compelling for microscale fluid
68: dynamics in microfluidic devices \cite{Raabe04} which typically present non-continuum
69: and surface-dominated effects (e.g. high Knudsen number conditions, electrokinetic and wetting phenomena).
70: This microscopic based
71: approach also makes the LBM a good candidate for hybrid, multi-scale simulations of fluid flows.
72:
73: In contrast, the extension of the LBM to non-Newtonian fluids has received limited attention so far,
74: in spite of the fact that a reliable extension to the LBM to simulate non-Newtonian flows would be
75: very valuable; for instance, in studies of transport in geological porous media, an area in which
76: the LBM has been extensively applied \cite{DrazerK00,DrazerK01,DrazerK02}
77: due to its simple implementation in complex geometries,
78: In addition to geological systems,
79: the flow of non-Newtonian fluids is commonly found in many areas of science and technology.
80:
81: In this work, we study an {\it ad hoc} modification of the LBM, first introduced
82: by Aharonov and Rothman \cite{AharonovR93}, in which the local value of the viscosity depends on the
83: strain-rate tensor. We show that this modification to the LBM accurately describes the flow of truncated
84: power-law fluids, both shear-thinning and shear-thickening, not only in unidirectional flows (parallel
85: plates geometry) but also in two-dimensional flows with simultaneous shear components in more than one
86: direction (reentrant corner geometry).
87:
88: \section{Lattice-Boltzmann Method}
89: \label{lbm}
90: The LBM can be viewed as an implementation of the
91: Boltzmann equation on a discrete lattice and for a
92: discrete set of velocity distribution functions \cite{RothmanZ},
93: \begin{equation}
94: f_i(\mathbf{x+e}_i\Delta x,t+\Delta t)=f_i(\mathbf x,t) + \Omega_i(f(\mathbf x,t)),
95: \end{equation}
96: where $f_i$ is the particle velocity distribution function along the $i$th direction,
97: $\Omega_i(f(\mathbf x,t))$ is the {\it collision operator} which takes into account
98: the rate of change in the distribution function due to collisions, $\Delta x$ and
99: $\Delta t$ are the space and time steps discretization, respectively \cite{ChenD98}.
100: Then, the density $\rho$ and momentum density $\rho \mathbf u$ are given by the first
101: two moments of the distribution functions,
102: \begin{equation}
103: \rho = \sum_i f_i, \qquad
104: \rho \mathbf u = \sum_i f_i \mathbf e_i,
105: \end{equation}
106: where we assumed that the discretization is consistent with the Boltzmann equation,
107: $\mathbf{x+e}_i$ corresponding to the nearest neighbors of the point $\mathbf x$.
108: Note that in the previous equation, and in the reminder of the article, all quantities
109: are rendered dimensionless using $\Delta x$ and $\Delta t$ as the characteristic space
110: and time scales, respectively. Also note that, as we are concerned with incompressible
111: flows, we do not need to introduce a dimension of mass.
112:
113: \subsection{BGK Approximation}
114: Assuming that the system is close to equilibrium the collision operator is
115: typically linearized about a local equilibrium distribution function,
116: $f_i^{eq}$, and assuming further that the local particle distribution
117: relaxes to equilibrium with a single characteristic (relaxation) time $\tau$,
118: we arrive at the Bhatnagar, Gross and Krook (BGK) approximation of the LBM \cite{ChenD98},
119: \begin{equation}
120: f_i(\mathbf{x+e}_i\Delta x,t+\Delta t)=f_i(\mathbf x,t) +
121: \frac{f_i(\mathbf x,t)-f_i^{eq}(\mathbf x,t)}{\tau},
122: \end{equation}
123: where the relaxation time $\tau$ is directly related to the kinematic
124: viscosity of the fluid, $\nu=(2\tau-1)/6$.
125:
126: \subsection{Non-Newtonian flows}
127: \label{nnf}
128:
129: The {\it ad hoc} extension of the LBM proposed by Aharonov and Rothman \cite{AharonovR93} to simulate
130: non-Newtonian fluids consists of determining the value of the relaxation time $\tau$ locally,
131: in such a way that the desired local value of the viscosity is recovered.
132: The viscosity is related to the local rate-of-strain through the constitutive equation for the
133: stress tensor. A commonly used model of non-Newtonian fluids is the {\it generalized Newtonian}
134: model, in which the relation between the stress tensor, $\sigma_{ij}$, and the rate-of-strain
135: tensor, $D_{ij}$, is similar to that for Newtonian fluids, $\sigma_{ij}=2\mu D_{ij}$, but with $\mu$ a
136: function of the invariants of the local rate-of-strain tensor, $\mu=\mu(D_{ij})$.
137: In particular, we are interested in a widely used model: the power-law expression \cite{BirdSL},
138: $\mu=m \dot\gamma^{n-1}$, where $n>0$ is a constant characterizing the fluid. The case
139: $n<1$ correspond to shear-thinning (pseudoplastic) fluids, whereas $n>1$ correspond to
140: shear-thickening (dilatant) fluids, and $n=1$ recovers the Newtonian behavior. The magnitude of the
141: local shear-rate $\dot\gamma$ is related to the second invariant of the rate-of-strain tensor,
142: $\dot\gamma = \sqrt{D_{ij} D_{ij}}$, where the components of the rate-of-strain tensor, $D_{ij}$,
143: are computed locally from the velocity field. In particular, after obtaining the instantaneous velocity
144: field from the LBM we then compute $D_{ij}$ from a first-order finite-difference approximation to the
145: local derivatives of the velocity.
146:
147: However, there is an obvious obstacle to a direct implementation of the power-law fluid in the LBM,
148: in that the effective viscosity diverges for zero shear rates ($\dot\gamma=0$) in a
149: shear-thinning fluid ( $n<1$). Analogously, the viscosity becomes zero for a shear-thickening
150: fluid at zero shear rates. In previous studies it is not clear how this problem was avoided.
151:
152: Clearly, both limits are unphysical and, in fact, it is known that many non-Newtonian fluids
153: exhibit a power-law behavior only in some range of shear-rates, and a constant viscosity is
154: observed outside that range \cite{BirdSL}. Here, we used the simplest model of such fluids:
155: the {\it truncated power-law} model,
156: \begin{equation}
157: \label{visc}
158: \nu(\dot\gamma)=\mu(\dot\gamma)/\rho = \left\{
159: \begin{array}{ccl}
160: m \dot\gamma_0^{(n-1)} &\hspace{0.5cm}& \dot\gamma< \dot\gamma_0 \\
161: m \dot\gamma^{(n-1)} && \dot\gamma_0 < \dot\gamma < \dot\gamma_{\infty} \\
162: m \dot\gamma_{\infty}^{(n-1)} && \dot\gamma>\dot\gamma_{\infty}
163: \end{array} \right.
164: \end{equation}
165:
166:
167:
168: Using the truncated power-law model has an additional advantage in the LBM. It is well known that
169: the LBM can accurately simulate viscous flows only in a limited range of kinematic viscosities.
170: The method becomes unstable for relaxation times close to $\tau \gtrsim 1/2$ \cite{NiuSCW04}
171: (small kinematic viscosities, $\nu \lesssim 0.001$) and its accuracy is very poor
172: for $\tau \gtrsim 1$ \cite{BehrendHW94} (relatively large kinematic viscosities, $\nu \gtrsim 1/6$).
173: Therefore, we set the lower and upper saturation values
174: of the kinematic viscosity in Eq.~(\ref{visc}) to $\nu_{min}= 0.001$ and
175: $\nu_{max}= 0.1$.
176: It is clear that the maximum value of the viscosity corresponds to the value at zero shear
177: rate for shear-thinning fluids ($n<1$), whereas the opposite is true for shear-thickening
178: materials ($n>1$).
179: Note that setting the value of the maximum model viscosity to $\nu_{max}=0.1$ for a given
180: maximum fluid viscosity $\nu^{\star}_{max}$ and a spatial resolution $\Delta x$ simply
181: corresponds to choosing a particular value of the time
182: step in order to satisfy, $\nu^{\star}_{max}=\nu_{max} (\Delta x^2/\Delta t)$ \cite{NourgalievDTJ03}.
183: Since the kinematic viscosity scales with $\Delta x^2/\Delta t$, to keep the dimensionless viscosity
184: constant we shall rescale $\Delta t$ according to the previous relationship
185: when we increase the number of lattice nodes $N$, that is, since $\Delta x \propto 1/N$ then $\Delta t \propto 1/N^2$.
186: In what follow we use the three-dimensional (face-centered-hypercubic) FCHC-projection model of the LBM
187: with 19 velocities (D3Q19 following the notation in Ref.~\cite{QianDL92}).
188:
189: \section{Flow between parallel plates}
190: \label{heleshaw}
191:
192: We first test the proposed LBM for non-Newtonian flows in a simple unidirectional flow,
193: the flow between two parallel plates separated a distance $b$ in the $z$-direction
194: (Hele-Shaw cell) in the presence of a pressure gradient in the $x$-direction.
195: We use periodic boundary conditions in both $x$ and $y$ directions. The resulting flow field is
196: unidirectional, with $v_x(z)$ the only non-zero velocity component, the rate-of-strain is
197: a scalar function of the local velocity, $\dot\gamma=|dv_x/dz|$, and the
198: Navier-Stokes equations are greatly simplified \cite{Leal}.
199:
200: \begin{figure}[htb]
201: \includegraphics*[width=\W]{n050_intermediate.eps}
202: \caption{\label{n050} Comparison between a Lattice-Boltzmann simulation and the analytical solution for the
203: flow between two parallel plates separated a distance $b=10$. The power-law exponent of the fluid is $n=0.50$
204: (shear-thinning). The pressure gradient is $\nabla P=6\times10^{-6}$, $\rho=1$, $\nu_0=0.1$,
205: $\nu_{\infty}=0.001$, $m=10^{-3}$, and $N=400$. The circles correspond to the Lattice-Boltzmann simulations.
206: The solid line corresponds to the analytical solution given by Eq.~(\ref{sol}). Also shown, in dashed lines,
207: are the continuation of the Newtonian and Power-Law solutions outside their regions of applicability.
208: The vertical, dashed lines correspond to the transition points, $z_l$ and $\tilde z_l=b-z_l$,
209: between the low shear-rate region $L$, and
210: the region of intermediate shear rates, $I$. }
211: \end{figure}
212:
213: In order to compute the exact solution to the Navier-Stokes equations for a pressure driven flow of a
214: truncated power-law fluid in the Hele-Shaw geometry we split the system into (in principle)
215: three different regions. We shall describe the regions between $z=0$ and $z=b/2$ and the analogous
216: regions for $z>b/2$ follow by symmetry. The first region we consider is the high-shear rate region close
217: to the walls, {\it Region H} for $z<z_h$, in which the shear rate exceeds $\dot\gamma_{\infty}$, and the
218: fluid is Newtonian with effective kinematic viscosity $\nu_{\infty}=m \dot\gamma_{\infty}^{(n-1)}$;
219: the second one is the intermediate region in which the fluid behaves as a power law
220: according to Eq.~(\ref{visc}), {\it Region I} for $z_h<z<z_l$; and the last one is the low-shear rate region
221: close to the center of the channel, {\it Region L} for $z_l<z<b/2$,
222: in which the shear rate is lower than $\dot\gamma_0$ and the fluid is again
223: Newtonian, but with kinematic viscosity $\nu_0=m\dot\gamma_0^{(n-1)}$.
224: Matching then the solution obtained in each region with the conditions of continuity in the velocity and
225: the stress, we obtain the general solution to the problem, in terms of the pressure gradient $G \rho =-\nabla P$,
226: \begin{equation}
227: \label{sol}
228: v_x(z) = \left\{
229: \begin{array}{lcl}
230: \left( \frac{G}{2 \nu_{\infty}} \right) z (b-z)&\hspace{0.5cm} & 0 \leq z \leq z_h \\
231: \frac{n}{n+1} \left( \frac{G}{m} \right)^{\frac{1}{n}}
232: \left[\left( \frac{b}{2} \right)^{\frac{n+1}{n}}-\left(\frac{b}{2}-z\right)^{\frac{n+1}{n}}\right]+\alpha_1
233: &\hspace{0.5cm}& z_h \leq z \leq z_l \\
234: \left(\frac{G}{2 \nu_0}\right) z (b-z)+\alpha_2 &\hspace{0.5cm}& z_l \leq z \leq b/2
235: \end{array} \right.,
236: \end{equation}
237: with the constants $\alpha_1$ and $\alpha_2$ given by,
238: \begin{eqnarray}
239: \alpha_1 &=& \left( \frac{G}{2 \nu_{\infty}}\right) z_h (b-z_h)-
240: \frac{n}{n+1} \left( \frac{G}{m} \right)^{\frac{1}{n}}
241: \left[\left(\frac{b}{2}\right)^{\frac{n+1}{n}}-\left(\frac{b}{2}-z_h\right)^{\frac{n+1}{n}}\right], \\ \nonumber
242: \alpha_2&=&\frac{n}{n+1} \left( \frac{G}{m} \right)^{\frac{1}{n}}
243: \left[ \left( \frac{b}{2}\right)^{\frac{n+1}{n}}-\left(\frac{b}{2}-z_l\right)^ {\frac{n+1}{n}} \right]
244: -\left( \frac{G}{2 \nu_0} \right) z_l (b-z_l)+\alpha_1,
245: \end{eqnarray}
246: and the transition points, $z_h$ and $z_l$,
247: \begin{eqnarray}
248: \label{z}
249: z_h&=&\frac{b}{2}-\left(\frac{\nu_{\infty}}{m^{1/n}} \right)^{\frac{n}{n-1}} \frac{1}{G}
250: =\frac{b}{2}-\frac{m\dot\gamma_{\infty}^n}{G}, \\ \nonumber
251: z_l&=&\frac{b}{2}-\left(\frac{\nu_{0}}{m^{1/n}} \right)^{\frac{n}{n-1}} \frac{1}{G}
252: =\frac{b}{2}-\frac{m\dot\gamma_0^n}{G}.
253: \end{eqnarray}
254: Clearly, the number of regions that coexist will depend on the magnitude of the imposed pressure
255: gradient $G$. For very small pressure gradients, $G\ll1$, both $z_h$ and $z_l$ become negative
256: (see the previous equation), which means that shear rates are smaller than $\dot\gamma_0$ across
257: the entire gap and only {\it Region L} exists. As $G$ increases,
258: there is a range of pressure gradients, $m\dot\gamma_0^n < (b/2) G < m \dot\gamma_{\infty}^n$,
259: for which $z_l>0$ but $z_h<0$, and therefore regions {\it L} and {\it I} coexist. Finally,
260: for $G>(2/b)m\dot\gamma_{\infty}^n$ we obtain $z_h>0$, and all three
261: regions are present in the flow. Note that for large values of $G$ both transition
262: points converge to the center of the cell, $z_l, z_h \to b/2$.
263: Thus, Eq.~(\ref{z}) allows us to choose the appropriate value of $G$ in order to investigate the different
264: regimes.
265:
266: \begin{figure}[htb]
267: \includegraphics*[width=\W]{n200_powerlaw.eps}
268: \caption{\label{n200} Comparison between a Lattice-Boltzmann simulation and the analytical solution for the
269: flow between two parallel plates separated a distance $b=10$. The power-law exponent of the fluid is $n=2.00$
270: (shear-thickening). The pressure gradient is $\nabla P=5\times10^{-6}$, $\rho=1$, $\nu_0=0.001$,
271: $\nu_{\infty}=0.1$, $m=10^{-3}$, and $N=400$. The circles correspond to the Lattice-Boltzmann simulations and
272: the solid line corresponds to the analytical solution given by Eq.~(\ref{sol}).
273: The vertical dashed lines correspond to the transition points, $z_l$ and $\tilde z_l=b-z_l$,
274: between the low shear-rate region $L$ and
275: the intermediate shear rates region $I$. }
276: \end{figure}
277:
278: We performed a large number of simulations for different values of the power-law exponent $n$. Specifically,
279: we consider two shear-thinning fluids, $n=0.50$ and $n=0.75$, and two shear-thickening fluids,
280: $n=1.25$ and $n=2.00$. In all cases we performed simulations for two different magnitudes of the
281: external forcing: one for which the region of low shear rates {\it L} is important, that is relatively small
282: pressure gradients for which $z_l \sim b/4$; and a second one in which the fluid behaves as a power-law fluid
283: almost in the entire gap, that is $z_l \sim b/2$. In both cases the shear-rate does not exceeds $\dot\gamma_{\infty}$.
284: In Fig.~\ref{n050} we present a comparison between the Lattice-Boltzmann results and the analytical solution given
285: in Eq.~(\ref{sol}) for a shear-thinning fluid with power-law exponent $n=0.50$. The simulation corresponds
286: to a relatively small pressure gradient for which the region of small shear-rates is large,
287: $z_l\sim b/4$. Both regions, {\it Region L} in which the fluid behaves as a Newtonian one,
288: and {\it Region I} in which the effective viscosity is a power-law, are shown.
289: The agreement with the analytical solution is excellent, with relative error close to $0.1\%$.
290: In Fig.~\ref{n200} we present a similar comparison between the LBM and the analytical solution, but for a
291: shear-thickening fluid ($n=2.00$) which behaves as a power-law fluid across almost the entire channel. Again
292: the agreement is excellent with relative error smaller than $0.1\%$.
293:
294: \begin{figure}[htb]
295: \includegraphics*[width=\W]{error.eps}
296: \caption{\label{error}Relative error of the LBM compared to the analytical solution for the
297: flow between parallel plates, as a function of the number of lattice points used in the simulations.
298: The points correspond to simulations with the LBM for four different fluids, two shear-thinning fluids
299: ($n=0.50$ and $n=0.75$) and two shear-thickening fluids ($n=1.25$ and $n=2.00$). For all fluids we also present
300: results corresponding to two different regimes:
301: one at high pressure gradients, in which the low-shear-rates region, {\it region L},
302: is small (Power-Law) and the other one at intermediate pressure gradients in which both regions {\it L} and {\it I}
303: are comparable (note that with the exception of $n=0.75$ both regimes give almost exactly the same relative error and,
304: in fact, the corresponding points overlap almost entirely).
305: In all cases, we increased the lattice resolution until the relative error was on the order of
306: $0.1\%$. The solid line shows the general trend of the data, $1/N$.}
307: \end{figure}
308:
309: Finally, for each of these cases we run a series of simulations in which the number of lattice nodes, $N$, in the direction of
310: the gap was increased from $10$ to $400$, and computed the relative error of the LBM results compared to the
311: analytical solutions, defined as $\sum_{i=1}^{N} (1-v_i^{LBM}/v_i^{Anal.})^2$. In order to obtain the accuracy of the LBM as a function
312: of the number of nodes, we simulated the same physical problem but changed $\Delta x$ from $1$ to $0.025$.
313: In addition, since the accuracy of the LBM depends on the model viscosity, we also changed $\Delta t$ according to
314: $\Delta t=\Delta x^2$, so that the model viscosity remains the same, independent of the number of nodes.
315: Then, in order to compare the velocity field always at the same physical time since startup, the number of time steps was increased
316: inversely proportional to $\Delta t$ (reaching $\sim 10^8$ time steps for $N=400$). In Fig.~\ref{error} we present the results
317: obtained for the different fluids and different pressure gradients. It is clear that, in all cases, the relative error
318: decreases, approximately as $1/N$, as the number of nodes is increased, and eventually becomes of the order of $0.1\%$
319: (an arbitrary target accuracy that we set for our simulations). The error was found to be independent of the
320: pressure gradient, or the size of the non-Newtonian region $I$, but strongly depends on the power-law exponent. In particular,
321: the relative error seems to increase as the magnitude of $(1-n)/n$ increases, with the error in the Newtonian case ($n=1$) decaying
322: faster than $1/N$. This is probably related to the first-order finite-difference approximation used to compute the spatial
323: derivatives of the fluid velocity which determine the local viscosity through Eq.~(\ref{visc}). It would then be possible to improve
324: the accuracy of the method by implementing a higher order approximation of the local shear-rates.
325:
326: \section{Reentrant corner flow}
327: \label{reentrant}
328:
329: \begin{figure}[htb]
330: \includegraphics[width=\W]{streamlinesF.eps}
331: \caption{\label{setup} Streamlines in the reentrant flow geometry. The flow direction is shown at the center of the channel.
332: Note the asymmetry due to inertia effects. The dashed lines show the lines in which we compare the solutions of the LBM method
333: with the solutions obtained by finite-element calculations.}
334: \end{figure}
335:
336: \begin{figure}[htb]
337: \includegraphics[width=\W]{shearF.eps}
338: \caption{\label{stress} Contour-plot of the local magnitude of the shear-rate, as computed with the LBM.
339: The lowest level of the plot corresponds to $\dot\gamma_0$, that is the fluid behaves as Newtonian
340: in those regions. High shear-rates, and accordingly high shear-stresses are localized at the {\it entrant} and
341: {\it reentrant} corners.}
342: \end{figure}
343:
344: In the previous section we tested the LBM for non-Newtonian fluids in a Hele-Shaw geometry and
345: found excellent agreement with the analytical solutions as the number of nodes was increased.
346: In that case, the flow is unidirectional and therefore the shear-rate is a scalar, which
347: is a rather simple type of flow. In contrast, we shall now test the LBM in a more demanding geometry,
348: that is the {\it reentrant corner} geometry sketched in Fig.~\ref{setup}. In this case,
349: the shear-rate is no longer a scalar as in the Hele-Shaw geometry and, in addition, the presence of
350: two singular points, located at the {\it entrant} and {\it reentrant} corners (see Fig.~\ref{setup}),
351: requires high accuracy in order to obtain satisfactory stress resolution near these points
352: (although no analog to the Moffatt's analysis near the corner is available for non-Newtonian fluids, it is
353: believed that not-integrable stress singularities develop in this case, and numerical techniques do not
354: always converge \cite{KoplikB97}).
355: Motivated by these issues we simulated the flow in the reentrant corner geometry using the LBM
356: for a shear-thinning fluid with power-law exponent $n=0.50$. In Fig.~\ref{setup} we present the
357: streamlines corresponding to the computed velocity field, obtained for a pressure
358: gradient $\nabla P=10^{-5}$, where the recirculation region inside the cavity can be observed
359: (note that the separation between streamlines was chosen for visualization purposes only and it is not
360: related to the local flow rate, since the magnitude of the fluid velocity sharply decays inside the
361: cavity). The corresponding Reynolds number is $Re=4$, computed with the maximum
362: viscosity, $\nu_0$, and the measured mean velocity. In fact, the streamlines shown in Fig.~\ref{setup} are
363: fore-aft asymmetric, due to inertia effects, which are absent in low-Reynolds-number flows.
364: We also computed the local magnitude of the shear-rate, related to the local stress field
365: through the constitutive relation given by Eq.~(\ref{visc}).
366: In Fig.~\ref{stress} we present a contour plot of the magnitude of the shear-rate in the reentrant corner
367: geometry with the lowest level in the contour plot corresponding to $\dot\gamma_0$.
368: It can be seen that the fluid is Newtonian in small regions
369: at the center of the channel and inside the recirculation region. It is also clear that, as discussed before,
370: both the {\it entrant} and {\it reentrant} corners are singular points where the shear-rate increases
371: to its highest values in the system.
372:
373: \begin{figure}[htb]
374: \includegraphics*[width=\W]{ux.eps}
375: \caption{\label{ux} Velocity component along the channel, $u_x$, plotted in a line perpendicular to the flow
376: (dashed line $X=20$ in Fig.~\ref{setup}). We compare the results of the finite-element calculations (solid line)
377: with the results of the LBM (points) for different lattice resolutions. In the inset we plot the velocity profile
378: inside the recirculation region.}
379:
380: \end{figure}
381: \begin{figure}[htb]
382: \includegraphics*[width=\W]{uy.eps}
383: \caption{\label{uy} Velocity component perpendicular to the flow direction, $u_y$, plotted in a line parallel
384: to the flow and close to the top wall of the channel (dashed line $Y=16$ in Fig.~\ref{setup}). We compare the
385: results of the finite-element calculations (solid line) with the results of the LBM (points) for different
386: lattice resolutions.}
387: \end{figure}
388:
389: Finally, in order to perform a more quantitative test of the LBM, we solved the problem
390: numerically using the finite-element commercial software FIDAP (Fluent Inc.).
391: In Figs.~\ref{ux} and \ref{uy} we compare the solutions obtained with the LBM for different
392: resolutions, ranging from $N=40 \times 40$ to $N=320 \times 320$, with the finite-element results
393: obtained with FIDAP. The comparison is made along two lines,
394: one oriented along the flow direction (dashed line at $Y=16$ in Fig.~\ref{setup}) and a second one oriented in
395: the perpendicular direction (dashed line at $X=20$ in Fig.~\ref{setup}).
396: In Fig.~\ref{ux} we compare the velocity along the channel, $u_x$, in the line perpendicular to $x$
397: that is located at the center of the system ($X=20$, see Fig.~\ref{setup}).
398: A velocity profile similar to that in a Hele-Shaw cell is observed for $0 < y \lesssim 20$, as well
399: as the recirculation flow inside the cavity (see the inset).
400: In Fig.~\ref{uy} we plot the velocity in the vertical direction, $u_y$, along a horizontal line
401: close to the upper wall of the channel ($Y=16$, see Fig.~\ref{setup}).
402: It is clear that there is some fluid penetration into the cavity ({\it entrant flow})
403: in the first half of the channel and some {\it reentrant flow} in the second half
404: (note that, as mentioned before, the flow is not symmetric about $x=20$ due to inertia
405: effects).
406: In both cases we found an excellent agreement between the two methods, and the
407: agreement clearly improved as the number of lattice nodes was increased in the LBM (the number of elements
408: in the finite-element computations was fixed to $100 \times 100$).
409:
410: \section{Conclusions}
411: We have extensively tested an {\it ad hoc} modification of the Lattice-Boltzmann Method that extends
412: its use to Generalized Newtonian fluids, in which the non-Newtonian character of the
413: fluids is modelled as an effective viscosity. Specifically, we calculated the accuracy of the method
414: for Truncated Power-Law fluids and showed that the relative error decays (linearly) as the
415: resolution of the lattice (number of lattice points) is increased. The error was computed
416: directly from the analytical solutions of the problem. The same trend was observed for both
417: shear-thinning ($n<1$) and shear-thickening ($n>1$) fluids, as well as for intermediate and
418: high shear-rates. In all cases the relative error was of the order of $0.1\%$
419: for the highest resolution employed.
420: Finally, we also tested the method in the reentrant flow
421: geometry and showed that it is in excellent agreement with the solution obtained by means of finite-element
422: calculations. Again, the accuracy of the method was shown to increase with the number of lattice points.
423:
424: %\begin{figure}[htb]
425: %\includegraphics[angle=-90,width=\W]{frac3D_surface.eps}
426: %\caption[surface]{\label{surface} Example of a numerically generated}
427: %\end{figure}
428:
429:
430: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
431: %% ACKNOWLEDGMENTS
432: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
433: \section*{Acknowledgments}
434:
435: This work is part of a collaboration supported by the Office of International Science and Engineering
436: of the National Science Foundation under Grant No. INT-0304781.
437: This research was supported by the Geosciences Research Program, Office of
438: Basic Energy Sciences, U.S. Department of Energy, and computational facilities
439: were provided by the National Energy Resources Scientific
440: Computer Center.
441:
442: \bibliography{articles,book,mybooks,mios}
443: \bibliographystyle{apsrev}
444:
445: \end{document}
446: