1: \documentclass[11pt]{article}
2: \usepackage{amsbsy}
3: \usepackage{fullpage}
4: \usepackage[dvips]{epsfig}
5: \usepackage{latexsym}
6: \usepackage{psfig}
7: \usepackage{epsfig}
8: \usepackage{rotate}
9: %\usepackage{labelfig}
10: \usepackage{amsmath}
11: \newcommand{\del}{\delta}
12: \newcommand{\ep}{\epsilon}
13: \newcommand{\vep}{\varepsilon}
14: \newcommand{\bb}{\bibitem}
15: \newcommand{\ue}{\bf{e}}
16: \newcommand{\ug}{{\bf g}}
17: \newcommand{\un}{{\bf n}}
18: \newcommand{\up}{\bf{p}}
19: \newcommand{\uq}{\bf{q}}
20: \newcommand{\uu}{{\bf u}}
21: \newcommand{\ux}{{\bf x}}
22: \newcommand{\TI}{{\bf T}^i}
23: \newcommand{\TJ}{{\bf T}^j}
24: \newcommand{\DI}{D^i}
25: \newcommand{\DDJ}{D^j}
26: \newcommand{\te}{{\mathcal E}}
27: \newcommand{\ts}{\mbox{\boldmath $S$}}
28: \newcommand{\uv}{\bf v}
29: \newcommand{\ee}{\mbox{\boldmath $e$}}
30: \newcommand{\bound}{\partial\Omega}
31: \newcommand{\uom}{\hbox{\boldmath{$\omega$}}}
32: \newcommand{\om}{\omega}
33:
34: \title
35: {Effective shear and extensional viscosities of
36: concentrated disordered suspensions of rigid particles.}
37: \author{Leonid Berlyand\\
38: Department of Mathematics and Materials Research Institute\\ Penn State
39: University\\
40: University Park, PA 16802, USA\\
41: ({\tt berlyand@math.psu.edu})\\
42: and\\
43: Alexander Panchenko\\
44: Department of
45: Mathematics\\
46: Washington State University\\
47: Pullman WA 99164, USA \\
48: ({\tt
49: panchenko@math.wsu.edu})}
50:
51:
52:
53: \begin{document}
54: \maketitle
55: %%%%%%%%%
56: \begin{abstract}
57: We study effective shear viscosity $\mu^\star$ and effective extensional
58: viscosity $\lambda^\star$ of concentrated non-colloidal suspensions of rigid spherical particles. The focus is on the spatially disordered arrays, but periodic arrays are considered as well.
59: We use recently developed discrete network approximation techniques to obtain asymptotic formulas for $\mu^\star$ and $\lambda^\star$ as the typical inter-particle distance $\delta$ tends to zero, assuming that the fluid flow is governed
60: by Stokes equations. For disordered arrays, the volume fraction alone does not determine the effective viscosity. Use of the network approximation allows us to
61: study the dependence of $\mu^\star$ and $\lambda^\star$ on variable distances between neighboring particles in such arrays.
62:
63: Our analysis, carried out for a two-dimensional
64: model, can be characterized as global because it goes beyond the local analysis of flow between two particles and takes into account hydrodynamical interactions in the entire particle array.
65: Previously, asymptotic
66: formulas for $\mu^\star$ and $\lambda^\star$
67: were obtained via asymptotic analysis of lubrication effects
68: in a single thin gap between two closely spaced particles. The principal conclusion in the paper is that, in general, asymptotic formulas for $\mu^\star$ and $\lambda^\star$ obtained by global analysis are different from the formulas obtained
69: from local analysis. In particular, we show that the leading term in the asymptotics of $\mu^\star$ is of lower order than suggested by the local analysis (weak blow up), while the order of the leading term in the asymptotics of $\lambda^\star$ depends on the geometry of the particle array (either weak or strong blow up).
70: We obtain geometric conditions on a random particle array under which
71: the asymptotic order of $\lambda^\star$ coincides with the order of the local
72: dissipation in a gap between two neighboring particles, and show that these conditions are generic. We also provide an example of a uniformly closely packed particle array
73: for which the leading term in the asymptotics of $\lambda^\star$ degenerates (weak
74: blow up).
75:
76: %shows that taking into account
77: %long-range hydrodynamic interactions (global analysis vs. local analysis)
78: %provides the correct asymptotic formulas for $\mu^\star$ and $\lambda^\star$.
79:
80:
81:
82:
83: \end{abstract}
84: \makeatletter \@addtoreset{equation}{section}
85: \def\theequation{\thesection.\arabic{equation}}
86: \makeatother
87: \makeatletter %\@addtoreset{figure}{section}
88: %\def\thefigure{\thesection.\arabic{figure}}
89: \makeatother
90: \newtheorem{theorem}{Theorem}[section]
91: \newtheorem{algorithm}{Algorithm}[section]
92: \newtheorem{problem}{Problem}[section]
93: \newtheorem{lemma}{Lemma}[section]
94: \newtheorem{assumption}{Assumption}[section]
95: \newtheorem{property}{Property}[section]
96: \newtheorem{example}{Example}[section]
97: \newtheorem{corollary}{Corollary}[section]
98: \newtheorem{proposition}{Proposition}[section]
99: \newtheorem{definition}{Definition}[section]
100: \newtheorem{remark}{Remark}[section]
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
103: \section{Introduction}
104: Concentrated suspensions are important in many industrial applications such as
105: drilling, water-coal slurries transport, food processing, cosmetics and ceramics manufacture.
106: In nature, flows
107: of concentrated suspensions appear as mud slides, lava flows and soils
108: liquefied by the earthquake-induced vibrations (\cite{Carr}, \cite{Cous}, \cite{Shook}).
109:
110:
111: An asymptotic formula for the effective viscosity of a
112: suspension of non-colloidal particles in a Newtonian fluid,
113: derived in \cite{FA}, is based on the local lubrication analysis of the
114: energy dissipation rate in the
115: narrow gap between a pair of nearly touching particles. The
116: distance between two neighboring particles in a periodic array is the small
117: parameter in the problem. For periodic arrays, this
118: {\it inter-particle distance} is uniquely determined by the volume fraction
119: of particles, so that the asymptotics of the
120: effective viscosity is obtained as a function of the volume fraction $\phi$
121: that is close to the
122: maximal packing volume fraction $\phi_{rcp}$. the asymptotics of the effective
123: viscosity obtained in \cite{FA} has the form
124: \begin{equation}
125: \label{00}
126: A\epsilon^{-1}+O(\ln\epsilon),
127: \end{equation}
128: as $\epsilon\to 0$, where $\epsilon=1-(\phi/\phi_{rcp})^{1/3}$.
129: The formulas for effective viscosity of
130: periodic suspensions in the whole space ${\bf R}^3$ (without boundary) subject
131: to a prescribed linear flow,
132: obtained in \cite{Keller}, also rely on the local lubrication analysis.
133: Asymptotic representations for the components of the effective viscosity tensor
134: calculated by \cite{Keller} are of the form
135: \begin{equation}
136: \label{001}
137: A\epsilon^{-1}+B\ln\epsilon +O(1),
138: \end{equation}
139: Recently, concentrated random suspensions were investigated numerically
140: by \cite{Brady} using accelerated Stokesian dynamics. It was observed
141: that the behaviour of the effective high frequency dynamic shear viscosity
142: of disordered suspensions can be accurately
143: described by the asymptotic $B\ln \epsilon$, indicating degeneration of the leading term in the asymptotic expansions (\ref{001}) (weak blow up). The authors of \cite{Brady} also show that
144: their results are in good agreement with available
145: experimental data (\cite{ShP}, \cite{VW}).
146: This suggests that for generic random suspensions, the
147: asymptotics of the effective viscosity defined by the (properly normalized) global dissipation rate cannot be identified with the local
148: dissipation rate in a single gap.
149:
150: In this paper, we use the discrete network approximation proposed in \cite{BBP}
151: to study the asymptotics of the shear effective viscosity $\mu^\star$ and
152: the extensional effective viscosity $\lambda^\star$ corresponding
153: to general disordered particle arrays. For such arrays, the volume fraction
154: alone is not sufficient for determining the effective viscosity.
155: Therefore, instead of $\epsilon$, we use the
156: inter-particle distance parameter $\delta$ that controls the distances
157: $\delta_{ij}$ between neighboring particles.
158:
159:
160: The mathematical construction of \cite{BBP} accounts for the
161: long range hydrodynamical interactions between the particles, and
162: provides an algorithm for calculation of the effective viscosity, which
163: takes into account variable distances between neighboring particles in non-periodic arrays.
164: Furthermore, in \cite{BBP} it was observed that the leading term of the asymptotics may degenerate
165: due to the external boundary conditions and geometry of the particle array, while
166: in the scalar case (\cite{BK}) the order of the leading term is the same for all
167: particle arrays that form a connected network. This paper is devoted to a detailed
168: study of this degeneration phenomenon. In particular, we clarify the issue
169: of weak versus strong blow up in the asymptotics of the effective viscosity.
170:
171: The definition of the effective viscosity employed in this paper differs from the
172: definition in \cite{Keller}. Their definition is closely related
173: to the definition of the effective material properties of periodic media in homogenization theory
174: (see e.g. \cite{BLP}, \cite{JKO}), where these properties are defined
175: via analysis on a periodicity
176: cell. Then these properties
177: are determined by the properties of constituents and geometry of the periodic array,
178: independently of the external boundary conditions and applied forces.
179: Such properties are usually defined in the limit when particle size tends to
180: zero while the number of particles approaches infinity.
181:
182:
183: Our definition is more directly linked to the viscometric measurements
184: that necessarily involve boundary conditions.
185: We assume that particles of a fixed size are placed in a bounded domain
186: (apparatus), and the typical distance between the
187: neighboring particles approaches zero. Thus, the total volume fraction of particles approaches the maximal packing fraction. In this case the effective viscosity
188: will be influenced not just by the inter-particle interactions but also by the forces or velocities prescribed at the boundary of the apparatus. Since we work
189: with linear models, $\mu^\star$ ($\lambda^\star$) are independent of the applied
190: shear (extension) rate. However, we also show that the asymptotic order of the effective viscosities depends on the relation between
191: the orientation of the velocity prescribed at the boundary, shape of the boundary, and orientations of the line segments connecting pairs of
192: neighboring particles.
193:
194:
195: We present sufficient conditions for degeneration
196: (weak blow up) and non-degeneration (strong blow up) of the leading term.
197: We show that $\mu^\star$ always exhibits weak blow up, while local
198: analysis alone predicts strong blow up. We also show that the asymptotic order
199: of $\lambda^\star$ depends on the geometry
200: of the particle array. In the paper, we define a broad class of arrays of particles
201: for which the leading order term of $\lambda^\star$ does not degenerate.
202: This situation is typical in the sense that for a
203: generic array $\mu^\star$ and $\lambda^\star$ have vastly different values, and
204: their ratio depends on the inter-particle distance,
205: which indicates possible non-Newtonian behaviour of the effective fluid
206: under the imposed boundary conditions.
207: We also give an example of
208: a closely packed particle array for which the leading term in
209: the extensional viscosity degenerates.
210: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
211: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
212: \section{Mathematical model}
213: We consider a concentrated suspension of rigid,
214: non-Brownian, neutrally buoyant
215: particles in a viscous
216: incompressible Newtonian fluid. In the two-dimensional model,
217: the suspension occupies a
218: domain $\Omega$ which is a square of side length two centered at the origin.
219: The boundary of $\Omega$ is denoted by
220: $\partial
221: \Omega$.
222: The upper and lower sides of $\partial\Omega$ are denoted by
223: $\partial\Omega^+=\{{\ux}: x_2=1\}$ and $\partial\Omega^-=\{{\ux}:x_2=-1\}$, respectively. We
224: also let ${\bf e}_1, {\bf e_2}$ denote the Cartesian basis
225: vectors parallel to the sides of $\partial\Omega$.
226: The particles
227: $D^j$, $j=1,2, \ldots, N$ are modelled as
228: rigid disks with centers ${\bf x}^j$, placed in $\Omega$. For simplicity, we consider
229: the monodisperse case, so that all particles have radius $a$. The part of $\Omega$
230: which is not occupied by particles is the fluid domain, denoted
231: by $\Omega_F$.
232:
233:
234:
235:
236: The fluid at low Reynolds number is
237: governed by Stokes equations
238: \begin{equation}
239: \label{eq1}
240: \mu\Delta {\uv}-\nabla P=0,\;\;\;\;{\rm div}~{\uv}=0,\;\;\;{\rm in}\;\Omega_F.
241: \end{equation}
242: where $\mu$ is the fluid viscosity, $\uv$ the velocity field, and $P$ is
243: the pressure.
244:
245:
246:
247:
248:
249: In this paper, we consider the external
250: boundary conditions of the shear and extensional types.
251: The shear type boundary conditions are given by
252: \begin{equation}
253: \label{shear-fbc}
254: {\bf v}=
255: \left\{
256: \begin{array}{cc}
257: \gamma{\bf e}_1\;\;\;&{\rm on}~\partial\Omega^+ ,\\
258: -\gamma{\bf e}_1\;\;\;&{\rm on}~\partial\Omega^-,\\
259: \end{array}
260: \right.
261: \end{equation}
262: where $\gamma$ is a constant shear rate.
263:
264:
265: In the case of extensional boundary conditions the velocity is prescribed as
266: \begin{equation}
267: \label{ext-fbc}
268: {\bf v}=
269: \left\{
270: \begin{array}{cc}
271: \vep(-{\bf e}_2+x_1{\bf e}_1)\;\;\;&{\rm on}~\partial\Omega^+ ,\\
272: \vep({\bf e}_2+x_1{\bf e}_1)\;\;\;&{\rm on}~\partial\Omega^-,\\
273: \end{array}
274: \right.
275: \end{equation}
276: where $\vep$ is a constant extension rate.
277: The two remaining (vertical) sides of $\partial\Omega$ are free surfaces, where
278: the zero traction condition is prescribed.
279:
280:
281: To define particle velocities, we first recall that
282: a rigid body moving in the plane defined
283: by ${\bf e}_1, {\bf e}_2$ has a velocity vector of the form
284: \begin{equation}
285: \label{p-vel}
286: {\bf v}^j({\bf x})={\bf T}^j+\omega^j{\bf e}_3\times ({\bf x}-{\bf x}^j),\;\;\;{\bf x}\in D^j,
287: \end{equation}
288: where ${\bf e}_3$ is the unit vector perpendicular to the plane of motion. Therefore, if
289: ${\bf r}=a{\bf e}_1+b{\bf e}_2$, then
290: ${\bf e}_3\times {\bf r}=-b{\bf e}_1+a{\bf e}_2$.
291: Equation (\ref{p-vel}) shows that the
292: velocity of a particle $D^j$, $j=1,2, \ldots, N$ is completely determined by two parameters:
293: a constant translational velocity vector ${\bf T}^j$ and a scalar
294: angular velocity $\omega^j$.
295: Both ${\bf T}^j$ and $\omega^j$ are unknown and must be
296: determined in the course of solving the problem.
297:
298:
299:
300:
301: Since each rigid disk is in equilibrium, the total force and torque
302: exerted on $D^j$ by the fluid must be zero, which provides the
303: boundary conditions on the particle boundaries $\partial D^j$:
304: \begin{equation}
305: \int_{\partial D^j} \ts \un^{j} \, ds = {\bf 0} \;\;\hbox{and}\;\;
306: \int_{\partial D^j} \un^{j} \times \ts \un^{(j)}\, ds = {\bf 0},
307: \;\hbox{for}\; j = 1,2 \ldots N,
308: \label{p-bc},
309: \end{equation}
310: where ${\bf n}^j$ is the exterior unit normal to $\partial D^j$, and
311: \begin{equation}
312: \label{def-stress}
313: \ts=2\mu\mbox{\boldmath $e$}({\bf v})-P\boldsymbol I.
314: \end{equation}
315: In (\ref{def-stress}) and throughout the paper, $\mbox{\boldmath $e$}({\bf v})$ denotes the strain rate tensor defined by
316: \begin{equation}
317: \label{strain}
318: \mbox{\boldmath $e$}({\bf v})=
319: \frac 12(\nabla
320: {\uv}+
321: \nabla {{\uv}}^T),
322: \end{equation}
323: the superscript $T$ stands for the transposed tensor, and $\boldsymbol I$ denotes the unit tensor.
324:
325:
326:
327: Solving equation (\ref{eq1}) with the boundary
328: conditions (\ref{shear-fbc}) (or (\ref{ext-fbc})) and (\ref{p-bc})
329: is equivalent to minimizing the functional
330: \begin{equation}
331: W_{\Omega_F}(\uu) = \frac{\mu}{4} {\displaystyle
332: \sum_{i,j = 1}^n} {\displaystyle \int_{\Omega_F}} \left(\frac{\partial
333: u_i(\ux)}{\partial x_j} + \frac{\partial u_j(\ux)}{\partial x_i}
334: \right)^2 d\ux,
335: \label{primal}
336: \end{equation}
337: over the function space ${\cal U}$ of admissible velocity fields $\uu$.
338: This space is a space of vector functions in $\Omega$ satisfying
339: either (\ref{shear-fbc}) or (\ref{ext-fbc}), and such that
340: \begin{equation}
341: \uu = {\displaystyle \sum_{j=1}^n} u_j \, {\bf e}_j,
342: ~u_j \in H^1(\Omega_F),~ j = 1 \ldots n,\; \hbox{div}\, \uu = 0, \;
343: (\ref{p-vel})\,{\rm holds}.
344: \label{sp_U}
345: \end{equation}
346: Since the fluid velocity ${\bf v}$ is the minimizer of the variational principle
347: (\ref{primal})-(\ref{sp_U}), the energy dissipation rate $E$ in the fluid (defined
348: in equation (\ref{sus-dis} below)
349: can be written as
350: \begin{equation}
351: \label{fl-dis-var}
352: E=W_{\Omega_F}({\bf v})=\min_{{\bf u}\in{\cal U}} W_{\Omega_F}({\bf u}).
353: \end{equation}
354:
355: In next section we will see that calculation of the effective viscosities
356: essentially amounts to calculation of $E$.
357:
358:
359:
360:
361:
362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363: \section{Effective shear and extensional viscosities.}
364: \subsection{Effective dissipation rates}
365: We suppose that the suspension can be modelled on a macroscale by a single phase viscous
366: fluid, called an {\it effective fluid}. The velocity
367: field of the effective fluid is denoted by ${\uv}^0$. The effective fluid
368: is subject to the same external boundary conditions as the flow of the
369: suspension.
370:
371: We assume that the effective
372: stress tensor ${\ts}^0$ satisfies the constitutive equation
373: of the form
374: \begin{equation}
375: \label{const}
376: {\ts}^0=F({\ee}({\bf v}^0)),
377: \end{equation}
378: where $F$ is a symmetric tensor function that does not depend explicitly on ${\bf x}$. In this paper,
379: we do not derive or postulate the
380: precise from of the constitutive law for the
381: effective fluid. Instead, we use the fundamental principle (going back to \cite{Einst}),
382: that
383: viscous energy dissipation rate of the suspension must be equal to the dissipation
384: rate of the effective homogeneous fluid. The dissipation rates are
385: defined by
386: \begin{equation}
387: \label{sus-dis}
388: E=\int_{\Omega_F} {\ts}\cdot {\ee}({\uv}) d{\ux}=
389: 2\mu\int_{\Omega_F} {\ee}({\uv})\cdot {\ee}({\uv}) d{\ux}
390: \end{equation}
391: in the suspension,
392: and
393: \begin{equation}
394: \label{eff-dis}
395: E^0=\int_\Omega {\ts}^0 \cdot {\ee}({\uv}^0) d\ux,
396: \end{equation}
397: in the effective fluid. In the equations (\ref{sus-dis}), (\ref{eff-dis}),
398: ${\ts}\cdot {\ee}={\ts}_{ij}{\ee}_{ij}$ is the inner product of tensors.
399:
400: For small particle volume fractions, (\cite{Einst}, \cite{Batch}), this principle was
401: further combined
402: with the assumption that the effective fluid is Newtonian with a constant
403: effective viscosity.
404: %For highly concentrated
405: %suspensions, the same assumption was used in \cite{FA}.
406: However, for concentrated suspensions this assumption is not
407: validated by rigorous mathematical derivation or experimental data and at present
408: the problem of finding the effective constitutive law for such suspensions is still
409: under investigation. Some experimental studies (e.g. \cite{ShP}, \cite{VW}) suggest
410: that the effective fluid may be non-Newtonian.
411: Our calculations of the effective viscosity suggest that
412: non-Newtonian behavior is possible for irregular (non-periodic or random) suspensions.
413:
414:
415:
416: We use the rheological definitions of shear
417: and extensional viscosities as ratios of the corresponding components of the stress
418: and strain rate tensors. To calculate the asymptotics
419: of the two viscosities, we employ the network approximation introduced in (\cite{BBP}).
420: We analyze the network functional (discrete dissipation rate) introduced in \cite{BBP}
421: and show that the standard relation between two viscosities
422: which holds for Newtonian fluids (see \cite{Sh} ch. 9 for the 3D-case and Appendix A for 2D-case )
423: does not hold.
424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
425: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
426: \subsection{Shear viscosity}
427: Suppose that a homogeneous effective fluid undergoes a steady shear flow
428: with the shear rate $\gamma$. The velocity field
429: ${\uv}_{\it sh}$ satisfies the shear type boundary conditions (\ref{shear-fbc}).
430: The effective shear viscosity is defined by (see (\ref{eo-shear}))
431: \begin{equation}
432: \label{shev1}
433: \mu^\star=\frac{S^0_{12}}{\gamma}=2 \frac{E^0}{\gamma^2 |\Omega|},
434: \end{equation}
435: where $S^0_{12}$ is the corresponding component of the effective stress tensor
436: and $|\Omega|=\int_\Omega d\ux$.
437: Note that our definition of $F^0$ in (\ref{const}) implies that
438: $\ts^0$ is constant when $\ee({\bf v}^0)$ is constant.
439: Since $E=E^0$, the equivalent definition is
440: \begin{equation}
441: \label{shev2}
442: \mu^\star=2 E \gamma^{-2} |\Omega|^{-1}=
443: 4\mu \gamma^{-2} |\Omega|^{-1}
444: \int_{\Omega_F} {\ee}({\uv}_{\it sh})\cdot\ee(\uv_{\it sh})d \ux
445: \end{equation}
446: Thus the calculation of $\mu^\star$ amounts to evaluation of the total dissipation rate integral
447: \begin{equation}
448: \label{dirint}
449: E_{sh}=2\mu\int_{\Omega_F} {\ee}({\uv}_{\it sh})\cdot\ee(\uv_{\it sh}){\it d}\ux,
450: \end{equation}
451: where $\uv_{\it sh}$ solves (\ref{eq1})-(\ref{p-bc}).
452:
453:
454:
455: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
456: \subsection{Extensional viscosity}
457: A steady extensional flow of the effective fluid is characterized by
458: a constant extension rate $\vep$. The velocity ${\uv}^0_{\it ext}$ satisfies
459: the extensional boundary conditions (\ref{ext-fbc}).
460: The extensional viscosity (see, e.g. \cite{Sh}. ch.9) may be defined by
461: \begin{equation}
462: \label{ev}
463: \lambda^\star=\frac{S_{11}^0-S_{22}^0}{\vep},
464: \end{equation}
465: where $S_{11}^0, S_{22}^0$ are components of the effective stress tensor.
466: Since
467: $E^0=\int_\Omega {\ts}^0\cdot\ee({\uv})^0_{\it ext}d\ux= (S_{11}^0-S_{22}^0)\vep|\Omega|$,
468: the effective
469: extensional viscosity can be defined
470: in terms of the suspension dissipation rate $E$, as follows.
471: \begin{equation}
472: \label{eev1}
473: \lambda^\star=\frac{E}{\vep^2 |\Omega|}=
474: 2\mu\vep^{-2}|\Omega|^{-1}\int_{\Omega_F} {\ee} ({\uv}_{\it ext}) \cdot {\ee}
475: ({\uv}_{\it ext}) d\ux,
476: \end{equation}
477: and calculation of $\lambda^\star$ again reduces to evaluation of the total dissipation
478: rate (\ref{dirint}) with ${\uv}_{\it sh}$ replaced by ${\uv}_{\it ext}$.
479: In the remaining part of the paper we derive asymptotic formulas for the total
480: dissipation rate under boundary conditions (\ref{shear-fbc}) and (\ref{ext-fbc}).
481:
482:
483:
484:
485: %%%%%%%%%%%%%%%%%%%% VORONOI TESS. AND NETWORK GRAPH%%%%%%%%%%%%%%%%%
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487: \section{The Network}
488: Let us consider an arbitrary distribution of circular particles (disks) $D^i$, whose centers
489: are points ${\bf x}^i$ in $\Omega$, for $i = 1,2, \ldots, N$. We suppose that $N$
490: is close to $N_{\hbox{max}}$, so that neighboring particles can be close to
491: touching one another.
492: The network consists of vertices ${\bf x}^i$ and edges. The edges connect only
493: vertices that correspond to
494: `` neighboring `` particles. Note that while for a periodic array
495: the notion of a neighboring vertex (particle) is obvious, for
496: non-periodic (e.g. random) arrays of particles it is not immediate. We introduce it
497: via a Voronoi tessellation, which is a partition of a plane (or a planar domain) into
498: the union of convex polygons $V_i$, called Voronoi cells,
499: corresponding to the set of vertices ${\bf x}^i$.
500: A Voronoi cell $V_i$ consists of all points in the plane
501: which are closer to ${\bf x}^i$
502: than to any other vertex ${\bf x}^j, j \ne i$.
503:
504:
505: The edges of $V_i$ can lie either on $\partial \Omega$ or in
506: the interior of $\Omega$. On each face of $V_i$, that lies inside
507: $\Omega$,
508: $
509: \mid \ux - {\bf x}^i \mid = \mid \ux - {\bf x}^j \mid, \;\hbox{for
510: some}\; i \ne j.
511: $
512:
513: \begin{definition}
514: \label{def:neighbors}
515: For each $i =
516: 1,2 \ldots N$, define the index set ${\cal N}_i$ by
517: \[
518: {\cal N}_i = \left\{ j \in \{1, 2 \ldots, N\}, \; j \ne
519: i, \; \hbox{such that}\; V_i \; \hbox{and}\; V_j \; \hbox{have a
520: common edge} \right\}.
521: \]
522: We call a particle $D^j$ a {\bf neighbor} of $D^i$ if $j$ belongs to
523: ${\cal N}_i$ that is the vertices ${\bf x}^i$ and ${\bf x}^j$ have a common edge in the
524: Voronoi tessellation.
525: \end{definition}
526: Note, that according to this definition, two particles are not neighbors if
527: their Voronoi cells have a common vertex but do not share an edge.
528:
529: The minimal distance between neighboring particles $\DI$ and $D^j$
530: is given by
531: \begin{equation}
532: \delta_{ij} = \mid {\bf x}^i - {\bf x}^j \mid -2a.
533: \label{gap}
534: \end{equation}
535: We call $\delta_{ij}$ {\it inter-particle distances}.
536: If a disk $\DI$ is close to the external boundary $\bound$, we define
537: the particle-boundary minimal distance $\delta_i$ by
538: \begin{equation}
539: \label{bgap}
540: \delta^i={\rm dist}({\bf x}^i, \bound)-a.
541: \end{equation}
542:
543:
544: To model the high concentration regime, we assume that $\delta_{ij}$ and $\delta_{i}$
545: satisfy
546: \begin{equation}
547: \label{ip-dist}
548: \delta_{ij}=\delta d_{ij}, \;\;\;\;\;\delta_i=\delta d_i,
549: \end{equation}
550: where $\delta$ is a small parameter in the problem, and the
551: {\it scaled inter-particle distances} $d_{ij}, d_i$ satisfy
552: \begin{equation}
553: \label{sc-dist}
554: c \leq d_{ij}\leq 1,\;\;\;\;c \leq d_{i}\leq 1,
555: \end{equation}
556: with a fixed positive $c$ independent of $i,j$.
557:
558: % Note that the condition $c>0$ makes collisions impossible.
559: \begin{definition}
560: \label{def:network}
561: The centers ${\bf x}^i$ of the particles $D^i$ are called the
562: {\bf interior vertices} of the network (graph) $\Gamma$, $i = 1, 2 \ldots N$.
563: The {\bf interior edges}
564: $b_{ij}$ of the network connect neighboring vertices ${\bf x}^i$ and ${\bf x}^j$. If a
565: Voronoi cell $V_i$ has edges belonging to $\partial \Omega^+ \bigcup
566: \partial \Omega^-$, the corresponding ${\bf x}^i$ is called a {\bf near-boundary
567: vertex}. Each near-boundary vertex ${\bf x}^i$ is connected with
568: $\bound^\pm$
569: by an {\bf exterior edge}, which is a segment $\tilde b_i$ perpendicular to $\bound^\pm$.
570: This segment intersects $\bound^\pm$ at
571: an {\bf exterior vertex} denoted by $\tilde{\ux}^{(i)}$.
572:
573:
574: Finally, the {\bf network (graph)} $\Gamma$ is the collection of all interior vertices, exterior vertices, and all the edges connecting these vertices.
575: \end{definition}
576:
577: The set of indices of the near-boundary vertices will be denoted by $I$, that is
578: $i\in I$ if the vertex ${\bf x}^i$ is connected to $\bound$. Also, we write $i\in I^+(I^-)$
579: if ${\bf x}^i$ is connected to $\bound^+(\bound^-)$.
580:
581:
582:
583:
584: Note that $\Gamma$ is essentially the Delaunay graph dual to the Voronoi
585: tessellation, and the above notions admit straightforward generalization to
586: three dimensions.
587:
588:
589:
590:
591: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
593: \section{Network approximation of the effective viscosity}
594: \subsection{Network equations}
595:
596:
597: To define the network approximation, we first assign
598: a translational velocity ${\bf T}^i$ and an angular velocity
599: $\omega^i$ of a particle $D^i$ to the corresponding interior vertex ${\bf x}^i$.
600: At the exterior vertices $\tilde {\bf x}^i$ we prescribe
601: the velocity vector ${\bf g}$ which represents the boundary conditions
602: (\ref{shear-fbc}) or (\ref{ext-fbc}).
603:
604:
605: For each pair of neighboring particles $\DI$ and $D^j$ we introduce a gap $\Pi_{ij}$
606: which represents a fluid region where lubrication effects are very strong as shown on
607: Fig.1. The width $R_{ij}$ of such gap is independent of $\delta$. For technical
608: reasons it is convenient to work with non-intersecting gaps. Since the maximal number
609: of gaps adjacent to each particle is equal to the maximal coordination number (number of neighbors),
610: K, we can choose $R_{ij}=KMa$, where $M$ is sufficiently small and fixed.
611: The precise value of $M$ is not essential for our
612: purposes.
613:
614:
615: %%%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%%%%%
616: \begin{figure}%[htbp]
617: \label{gap1}
618: \begin{center}
619: \input{gap1.pstex_t}
620: \caption{A gap $\Pi_{ij}$ between the neighboring particles $D^i$ and $D^j$.}
621: \end{center}
622: \end{figure}
623:
624:
625:
626:
627:
628: The orientation of each interior gap $\Pi_{ij}$ relative to
629: a disk $D^i$ is specified by a unit vector
630: \begin{equation}
631: \label{orient}
632: {\bf q}^{ij}=\frac{{\bf x}^i-{\bf x}^j}{|{\bf x}^i-{\bf x}^j|}.
633: \end{equation}
634: We also let ${\bf p}^{ij}$ be the unit vector obtained by rotating
635: ${\bf q}^{ij}$ clockwise by $\pi/2$ (see Fig.2).
636:
637:
638:
639: Boundary edges and corresponding gaps are oriented perpendicular to ${\bf e}_1$.
640: This
641: reflects the physical fact that the zone of the largest energy
642: dissipation is located near the shortest line connecting ${\bf x}^i$ with
643: the boundary. Therefore, ${\bf q}^i={\bf e_2}$, (respectively $-{\bf e}_2$),
644: when $\DI$ is adjacent to $\bound^+$ (respectively $\bound^-$).
645:
646: %%%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%%%%%
647: \begin{figure}%[htbp]
648: \label{gap2}
649: \begin{center}
650: \input{gap2.pstex_t}
651: \caption{Assignment of ${\bf T}^i, \omega^i$ and
652: orientation of the gap between two neighboring particles.}
653: \end{center}
654: \end{figure}
655:
656:
657:
658: Next, to each edge of the network we associate a dissipation rate
659: $W^{ij}$ ($W^i$), calculated in the corresponding gap $\Pi_{ij}$ ($\Pi_i$).
660: The calculation of the dissipation rates employs lubrication approximation in the gap.
661: The velocity in the gap is decomposed into three velocities, representing
662: the "elementary" motions
663: called {\it spring motion}, shear, and rotation
664: (see Fig. 3). The total velocity field in a gap is the sum of
665: these elementary velocities and a "residual" velocity field, whose contribution
666: to the gap dissipation rate is $O(1)$ as $\delta\to 0$. Lubrication approximations
667: for each of the elementary velocities and estimates for the residual
668: are obtained in \cite{BBP} where more details can be found.
669:
670:
671:
672: %%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%%%%
673: \begin{figure}%[htbp]
674: \label{gap2}
675: \begin{center}
676: \includegraphics{elvel2.eps}
677: \caption{Three elementary motions. Arrows represent the boundary conditions.}
678: \end{center}
679: \end{figure}
680:
681: Using approximations of the elementary velocities to calculate (up to the terms of order
682: $O(1)$ as $\delta\to 0$) the
683: dissipation rates in each gap we obtain
684: \begin{equation}
685: \label{loc1-dis}
686: \begin{array}{ccc}
687: W^{ij} & = & \hspace{-5.1cm}\delta^{-3/2} C^{ij}_{sp}\left[(\TI-\TJ) \cdot {\bf
688: q}^{ij}\right]^2+ \\
689: & & \delta^{-1/2} C^{ij}_{sh}\left[(\TI -\TJ)\cdot
690: {\bf p}^{ij} + a{\om}^i + a{\om}^{j} \right]^2 +
691: \delta^{-1/2}C^{ij}_{rot}a^2(\omega^i-\omega^j)^2,\\
692: \end{array}
693: \end{equation}
694: in the interior gaps $\Pi_{ij}$,
695: and
696: \begin{equation}
697: \label{loc1-dis}
698: W^{i}=\delta^{-3/2} C^{i}_{sp}\left[(\TI-{\bf
699: g})\cdot{\bf q}^{i} \right]^2 +\delta^{-1/2} C^{i}_{rot}a^2\left(2 {\om}^i\right)^2+
700: \delta^{-1/2} C_{sh}^i \left[(\TI -{\bf g})\cdot {\bf p}^{i} + a{\om}^i
701: \right]^2
702: \end{equation}
703: in the boundary gaps $\Pi_i$.
704: The expressions for factors $C^{ij}_{sp}, C^{ij}_{sh}$ and
705: $C^{ij}_{rot}$ and $C^i_{sp}, C^i_{sh}, C^i_{rot}$ are calculated explicitly
706: in \cite{BBP}:
707: \begin{equation}
708: \label{net-2D-4}
709: \begin{array}{ccccccc}
710: C^{ij}_{sp} &=& \frac 34 \pi\mu \big(\frac{a}{d_{ij}}\big)^{3/2}+
711: \frac{27}{10} \pi\mu \big(\frac{a}{d_{ij}}\big)^{1/2},& \hspace{0.5cm}&
712: C^{i}_{sp} &=& \frac 34 \pi\mu \big(\frac{a}{d_{i}}\big)^{3/2}+
713: \frac{27}{10} \pi\mu \big(\frac{a}{d_{i}}\big)^{1/2},
714: \\
715: C^{ij}_{sh} & =& \frac 12 \pi\mu \big(\frac{a}{d_{ij}}\big)^{1/2},\hspace{1.4cm}&
716: \hspace{0.5cm} &
717: C^{i}_{sh} & =& \frac 12 \pi\mu \big(\frac{a}{d_{i}}\big)^{1/2},\hspace{1.4cm}
718: \\
719: C^{ij}_{rot} & =& \frac{9}{16} \pi\mu \big(\frac{a}{d_{ij}}\big)^{1/2}.\hspace{1.4cm}&
720: \hspace{0.5cm}&
721: C^{i}_{rot} & =& \frac{9}{16} \pi\mu \big(\frac{a}{d_{i}}\big)^{1/2}.\hspace{1.4cm}
722: \\
723: \end{array}
724: \end{equation}
725: In the formulas (\ref{net-2D-4}), $d_{ij}, d_i$ are the scaled inter-particle distances
726: defined in (\ref{sc-dist}).
727:
728: The sum of the local dissipation rates $W^{ij}, W^i$
729: is a quadratic form
730: \begin{equation}
731: \begin{array}{lll}
732: Q & = & \sum_{\Pi_{ij}} W^{ij} +\sum_{\Pi_i} W^i\\
733: &= & {\displaystyle
734: \sum_{i=1}^N} {\displaystyle \sum_{{\tiny \begin{array}{c}j \in {\cal
735: N}_i\\ j < i
736: \end{array}}}}\left\{
737: \delta^{-3/2} C^{ij}_{sp}\left[(\TI-\TJ) \cdot {\bf
738: q}^{ij}\right]^2+\delta^{-1/2} C^{ij}_{sh}\left[(\TI -\TJ)\cdot
739: {\bf p}^{ij} + a{\om}^i + a{\om}^{j} \right]^2 \right. \\ &&
740: +\left. \delta^{-1/2} C^{ij}_{rot}a^2\left({\om}^i -
741: {\om}^j\right)^2 \right\} + {\displaystyle
742: \sum_{i \in I}} \left\{ \delta^{-3/2} C^{i}_{sp}\left[(\TI-{\bf
743: g})\cdot{\bf q}^{i} \right]^2 +\delta^{-1/2} C^{i}_{rot}a^2\left(2 {\om}^i\right)^2 \right.
744: \\
745: && + \left. \delta^{-1/2} C_{sh}^i \left[(\TI -{\bf g})\cdot {\bf p}^{i} +
746: a{\om}^i
747: \right]^2 \right\},
748: \end{array}
749: \label{funct-q}
750: \end{equation}
751: where $I$ denotes the set of indices of the near-boundary vertices defined in
752: Section 4.
753:
754:
755: The form $Q$ is the discrete network approximation of the functional $W_{\Omega_F}$ in
756: the variational principle (\ref{primal}).
757: The main idea of the network approximation is that the most of the energy is dissipated
758: in the gaps $\Pi_{ij}, \Pi_i$, so that
759: \begin{equation}
760: \label{E}
761: E=E_{net}+O(1),\;\;\;\;{\rm as}\;\delta\to 0.
762: \end{equation}
763: The discrete dissipation rate $E_{net}$ in (\ref{E}) is
764: defined by
765: \begin{equation}
766: \label{ddrate}
767: E_{net}=\min_{{\bf T}^i, \omega^i} Q=Q({\bf T}_{min}^1, {\bf T}^2_{min},...
768: {\bf T}^N_{min}, {\om}^1_{min},...{\om}^N_{min}),
769: \end{equation}
770: where the vectors $\TI_{min}, i=1,2,\ldots, N$ and scalars ${\om}^i, i=1, 2, \ldots, N$ minimize
771: $Q$ in (\ref{funct-q}). The minimum in (\ref{ddrate}) is taken over all possible
772: collections of ${\bf T}^i, \omega^i$.
773:
774: It is well known that solving the minimization problem for $Q$ is
775: equivalent to solving the linear system (Euler-Lagrange equations), which is obtained
776: by equating the gradient of $Q$ to zero. Setting to zero partial derivatives with respect to
777: $T^i_l, l=1,2$ we obtain
778: \begin{eqnarray}
779: \sum_{j \in {\cal N}_i} \left\{ \delta^{-3/2} C^{ij}_{sp}
780: \left[(\TI-\TJ)\cdot{\uq}^{ij}\right] {\uq}^{ij}\right\}+\hspace{3.65cm}
781: \nonumber\\
782: \sum_{j \in {\cal N}_i} \left\{\delta^{-1/2} C^{ij}_{sh}\left[(\TI-\TJ) \cdot{\up}^{ij}+ a{\om}^i +
783: a{\om}^j \right] {\up}^{ij} \right\} +{\bf B}^i = {\bf F}^i, ~ ~
784: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
785: \label{bal-force}
786: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
787: \end{eqnarray}
788: for each $i=1,2,..., N$,
789: where
790: \begin{eqnarray}
791: {\bf B}^i &=& \left\{ \begin{array}{l}
792: \delta^{-3/2} C^{i}_{sp}(\TI \cdot {\bf
793: q}^i) {\bf q}^i
794: + \delta^{-1/2} C^{i}_{sh} \left[
795: \TI\cdot {\bf p}^i + a {\om}^i\right] {\bf p}^i, ~
796: ~\mbox{if} ~i \in I \\
797: {\bf 0} ~ ~\mbox{otherwise},
798: \end{array} \right.
799: %%%%%%%
800: \label{bd-force} \\
801: \end{eqnarray}
802: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
803: \begin{eqnarray}
804: {\bf F}^i &=& \left\{ \begin{array}{l}
805: \delta^{-3/2} C^{i}_{sp}({\bf g} \cdot {\bf
806: q}^i) {\bf q}^i
807: + \delta^{-1/2} C^{i}_{sh} (
808: {\bf g}\cdot {\bf p}^i) {\bf p}^i, ~
809: ~\mbox{if} ~i \in I \\
810: {\bf 0} ~ ~\mbox{otherwise},
811: \end{array} \right.
812: %%%%%%%
813: \label{rhs-force} \\
814: \end{eqnarray}
815: Next, equating the partial derivatives $\frac{\partial Q}{\partial \omega^i}$ to zero we obtain
816: \begin{eqnarray}
817: \sum_{j\in
818: {\cal N}_i} \left\{ \delta^{-1/2} C^{ij}_{sh}\left[(\TI-\TJ)\cdot{\up}^{ij}+ a{\om}^i +
819: a{\om}^j\right]\right\} + \nonumber \\
820: \sum_{j\in
821: {\cal N}_i} \left\{ \delta^{-1/2} C^{ij}_{rot}\left({\om}^i -
822: {\om}^j\right)\right\} + {\cal B}^i = {\cal M}^i,\hspace{1.5cm}
823: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
824: \label{bal-torque}
825: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
826: \end{eqnarray}
827: for all $i = 1, \ldots, N$, where
828: %%%%%% RHS%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
830: \begin{eqnarray}
831: {\cal B}^i &=& \left\{ \begin{array}{l}
832: \delta^{-1/2} C_{sh}^i \left(\TI \cdot {\bf
833: p}^i +a {\om}^i\right) + 4 \delta^{-1/2} C^{i}_{rot} {\om}^i, ~
834: ~\mbox{if} ~i \in I, \\ 0 ~ ~\mbox{otherwise}.
835: \end{array}\right.
836: %%%%%%%%
837: \label{bd-torque}
838: %%%%%%%%
839: \end{eqnarray}
840: %%%%%%%%%%%%%%%%%%%%%%%%%%%
841: \begin{eqnarray}
842: {\cal M}^i &=& \left\{ \begin{array}{l}
843: \delta^{-1/2} C_{sh}^i \left({\bf g} \cdot {\bf
844: p}^i\right) ~
845: ~\mbox{if} ~i \in I, \\ 0 ~ ~\mbox{otherwise}.
846: \end{array}\right.\hspace{4cm}
847: %%%%%%%%
848: \label{torque-rhs}
849: %%%%%%%%
850: \end{eqnarray}
851:
852: Equations (\ref{bal-force}) and (\ref{bal-torque}) are, respectively,
853: the equations of force and torque balance of the particles,
854: and the minimization in (\ref{funct-q}) ensures that the rigid body
855: translational and angular velocities are chosen in such a way that
856: the suspension is in mechanical equilibrium. Note also that (\ref{bal-force})
857: is a system of $2N$ equations, and (\ref{bal-torque}) is a system of
858: $N$ equations. Together they form $3N$ equations for $3N$ unknowns
859: $({\bf T}^i, \omega^i)$. The coefficients and right hand side of
860: (\ref{bal-force}) are of order $\delta^{-3/2}$ and $\delta^{-1/2}$ while
861: all the coefficients in (\ref{bal-torque}) are of order $\delta^{-1/2}$.
862: When all ${\bf T}^i$ are zero (no translations), the remaining terms
863: are of order $\delta^{-1/2}$, but in the case $\omega^i=0, i=1,2,\ldots, N$
864: (no rotations), the remaining equations contain terms of order $\delta^{-3/2}$.
865: This reflects the well known fact that the contributions from local translational spring motions are stronger
866: than the contributions from rotational and other translational motions. Keeping only
867: spring translations we obtain
868: the truncated (leading order) discrete dissipation functional
869: \begin{equation}
870: \label{lead-funct}
871: \delta^{-3/2}\widehat Q=
872: \delta^{-3/2} \displaystyle{\left[\frac 12 \sum_{i=1}^N \sum_{j\in {\cal N}_i} C^{ij}_{sp}
873: (({\bf T}^i-{\bf T}^j)\cdot{\bf q}^{ij})^2+\sum_{i\in I} C_{sp}^{i}
874: ((\TI-\ug)\cdot {\bf q}^{i})^2\right].}
875: \end{equation}
876: Then $Q$ can be decomposed as follows.
877: \begin{equation}
878: \label{q-decomp}
879: Q= \delta^{-3/2}\widehat Q+ \delta^{-1/2} Q^\prime,
880: \end{equation}
881: where the coefficients of the forms $\widehat Q$ and $Q^\prime$ do not depend on
882: $\delta$.
883:
884:
885: We next show that the total discrete dissipation rate $E_{net}$ can be estimated by the truncated
886: dissipation rate obtained by minimizing $\widehat Q$.
887: Let $\widehat{\TI}, i=1,2,\ldots, N$
888: denote the translational velocities which minimize $\widehat Q$ ( they are clearly
889: independent
890: of $\delta$). Since $Q^\prime$ in (\ref{q-decomp}) is non-negative,
891: \begin{eqnarray}
892: \delta^{-3/2}\widehat Q(\widehat{\TI})\leq \delta^{-3/2}\widehat Q(\TI_{min},
893: {\om}^i_{min})\leq
894: E_{net}=\nonumber \hspace{1cm}\\
895: Q(\TI_{min}, \omega^i_{\min})\leq \delta^{-3/2}\widehat
896: Q(\widehat{\TI})+\delta^{-1/2}Q^\prime(\widehat{\TI},{\om}^i=0).
897: \label{var-in}
898: \end{eqnarray}
899: Since $Q^\prime(\widehat{\TI},{\om}^i=0)\leq C$ with $C$ independent of
900: $\delta$ as $\delta\to 0$,
901: (\ref{var-in}) implies
902: \begin{equation}
903: \label{lead-term}
904: E_{net}=\delta^{-3/2}\widehat E+O(\delta^{-1/2}),\;\;\;\;{\rm as}~\delta\to 0,
905: \end{equation}
906: where
907: \begin{equation}
908: \label{lead-dis}
909: \widehat E=\widehat Q(\widehat{\TI})={\displaystyle \min_{{\tiny
910: {\bf T}^1,\ldots {\bf T}^N}} \widehat Q ({\bf T}^1,\ldots {\bf T}^N}).
911: \end{equation}
912: This equation enables one to calculate the leading term in the asymptotics
913: of the effective viscosity by solving a simplified minimization problem
914: involving only the translational particle velocities. However, this algorithm
915: is useful only when
916: \begin{equation}
917: \label{crucial}
918: \widehat E>0,
919: \end{equation}
920: because in this case the leading term in the asymptotics of the dissipation rate is of order
921: $\delta^{-3/2}$.
922: If $\min\widehat Q=0$, the leading term degenerates, and
923: the rate of blow up in (\ref{lead-dis}) is at most $\delta^{-1/2}$.
924:
925:
926: Minimization of the truncated quadratic form $\widehat Q$ corresponds to solving
927: the truncated linear system of the Euler-Lagrange equations
928: \begin{equation}
929: \label{lead-system}
930: \sum_{j\in {\cal N}_i}\big[C^{ij}_{sp}
931: (\widehat{{\bf T}}^i-\widehat{{\bf T}}^j)\cdot{\uq}^{ij}{\uq}^{ij}\big]+
932: {\bf B}({\widehat{\TI}})={\bf R}^i,\;\;\;\;i=1,2, \ldots, N,
933: \end{equation}
934: where
935: \begin{equation}
936: \label{B}
937: {\bf B}(\widehat{\TI})=
938: \begin{cases}
939: C_{sp}^{i}
940: (\widehat{\TI}\cdot {\bf q}^{i}){\bf q}^{i},\;\;\;\;{\rm when}~i \in I&\\
941: 0 \;\;\;\;\;\;{\rm otherwise}.&\\
942: \end{cases}
943: \end{equation}
944: The right hand side vectors ${\bf R}^i$ represent external boundary conditions:
945: \begin{equation}
946: \label{lead-rhs}
947: {\bf R}^i=
948: \begin{cases}
949: C_{sp}^{i}
950: ({\bf g}\cdot {\bf q}^{i}){\bf q}^{i},\;\;\;\;{\rm when}~i \in I&\\
951: 0 \;\;\;\;\;\;{\rm otherwise}.&\\
952: \end{cases}
953: \end{equation}
954:
955:
956: To better see the structure of the functional $\widehat Q$ and the system (\ref{lead-system})
957: it is convenient to rewrite them
958: in a compact form as
959: \begin{equation}
960: \label{comp-hatq}
961: \hat Q({\bf z})=\frac 12 A{\bf z}\cdot{\bf z}-{\bf f}\cdot {\bf z}+r,
962: \end{equation}
963: and
964: \begin{equation}
965: \label{compls}
966: A{\bf z}={\bf f},
967: \end{equation}
968: where ${\bf z}$ is the particle velocity vector, and ${\bf f}$ is the vector of discretized
969: boundary conditions. The vectors
970: ${\bf z}, {\bf f}\in {\bf R}^{2N}$ are defined by
971: \begin{equation}
972: \label{zf}
973: {\bf z}=
974: \left(
975: \begin{array}{c}
976: T^1_1\\
977: T^1_2\\
978: T^2_1\\
979: T^2_2\\
980: ...\\
981: T^N_1\\
982: T^N_2\\
983: \end{array}
984: \right),
985: \;\;\;\;\;\;\;{\rm and}\;\;\;\;\;\;\;
986: {\bf f}=2
987: \left(
988: \begin{array}{c}
989: R^1_1\\
990: R^1_2\\
991: R^2_1\\
992: R^2_2\\
993: ...\\
994: R^N_1\\
995: R^N_2,\\
996: \end{array}
997: \right).
998: \end{equation}
999: The fixed scalar $r$ equals $\tilde {\bf f}\cdot {\bf f}$ where components of
1000: $\tilde{\bf f}$ are
1001: defined by the components of ${\bf f}$, as follows.
1002: $$
1003: \tilde f_{2k-1}=\frac{1}{4 C_{sp}^k}f_{2k-1},\;\;\;\;
1004: \tilde f_{2k}=\frac{1}{4 C_{sp}^k}f_{2k}\;\;\;\;\;\;k=1,2,\ldots, N.
1005: $$
1006: The matrix $A$ in (\ref{comp-hatq}) is symmetric and non-negative definite.
1007: Its entries are determined by the scaled inter-particle distances $d_{ij}$,
1008: particle radius $a$ and vectors
1009: ${\bf q}^{ij}, {\bf q}^i$ defined by (\ref{orient}).
1010: The vector ${\bf f}$ and the scalar $r$ in (\ref{comp-hatq}) are determined by the
1011: boundary conditions on $\bound^+,\bound^-$ (extensional or shear). The linear term
1012: ${\bf f}\cdot{\bf z}$ depends on the translational velocities of the particles located
1013: near $\bound^+$ or $\bound^-$.
1014:
1015: To the leading order in $\delta$, the effective viscosity is determined by the
1016: discrete dissipation rate $\delta^{-3/2}\widehat E$ where $\widehat E$
1017: is the minimum of
1018: $\widehat Q$. Thus the qualitative behaviour of the effective viscosity is
1019: determined
1020: by solving (\ref{compls}). We now sketch the issues which arise in computing the leading
1021: term of the effective viscosity. First, in the case of shear viscosity,
1022: ${\bf f}=0$, which means that the shear boundary conditions do not contribute
1023: to the strong blow up. This results in the weak blow up of $\mu^\star$ (see section 6).
1024: In case of extensional conditions, ${\bf f}\ne 0$, which means that the right hand side
1025: of the full network equations (\ref{bal-force}) contains terms of order $\delta^{-3/2}$.
1026: In this case, calculation of the leading term is impeded by the fact that the matrix
1027: $A$ is not invertible. Indeed,
1028: from (\ref{lead-funct}) it is clear that the value of $\widehat Q$ does not
1029: change when all vectors ${\bf T}^i, i=1,2, \ldots, N$ are replaced by
1030: $\TI+t{\bf e}_1$ (horizontal translation by
1031: $t{\bf e}_1$, $t$ arbitrary real). This is not surprising,
1032: because the functional $\widehat Q$
1033: is invariant under horizontal translation of ${\bf T}^i$. The invariance is due
1034: to the vertical orientation of the boundary gaps
1035: explained in section 5.1.
1036: Note that this is not equivalent to translation of a coordinate system, since ${\bf g}$
1037: is not changed. Translational invariance of $\widehat Q$ implies that
1038: any ${\bf z}^*$ that solves the non-homogeneous system (\ref{compls}) produces other solutions
1039: of the form
1040: $
1041: {\bf z}^*+t{\bf w}_0,
1042: $
1043: where $t$ is arbitrary real, and
1044: \begin{equation}
1045: \label{w-not}
1046: {\bf w}_0=(1,0, 1, 0, \ldots, 1,0)^T.
1047: \end{equation}
1048: This means that vectors of the form $t{\bf w}_0$ solve the homogeneous system
1049: \begin{equation}
1050: \label{comphom}
1051: A{\bf z}={\bf 0}.
1052: \end{equation}
1053: Because of the above mentioned invariance of
1054: the functional $\widehat Q$, the leading term in the asymptotics of effective
1055: viscosity will be uniquely defined, unless (\ref{comphom}) has
1056: some other nontrivial solutions. Therefore, it makes sense to look
1057: for conditions on the network which would guarantee that every solution of the
1058: homogeneous system is of the form $t{\bf w}_0$. Then the
1059: non-homogeneous system (\ref{compls}) would be uniquely solvable
1060: up to horizontal translation.
1061: In Section 7.3 we show that for a typical random distribution of particles
1062: this is indeed the case.
1063:
1064: The last issue concerns the validity of the estimate (\ref{crucial}). The functional
1065: $\widehat Q$ is non-negative, but it may be zero. When (\ref{crucial})
1066: holds, local lubrication analysis provides the correct order of the leading
1067: term in the asymptotics of the extensional effective viscosity
1068: ($\delta^{-3/2}$ in dimension two and $\delta^{-1}$ in dimension three).
1069: If (\ref{crucial})
1070: does not hold, that is,
1071: \begin{equation}
1072: \label{anticrucial}
1073: \min \widehat Q=0,
1074: \end{equation}
1075: then the leading term in dimension two is of order $\delta^{-1/2}$, ($\ln(1/\delta)$ in
1076: dimension three).
1077:
1078: Whether or not
1079: the estimate (\ref{crucial}) holds, depends on the geometry of the
1080: particle array as well as
1081: the boundary conditions on $\bound^+$ and $\bound^-$.
1082: In section 7.4 we show that in the case of extensional boundary conditions,
1083: (\ref{crucial}) holds for generic arrays, and thus the leading
1084: term in the asymptotic of $E_{net}$ ( and extensional effective viscosity
1085: $\lambda^*$, see (\ref{eev1})) is of the order $\delta^{-3/2}$.
1086: However, there exist special arrays for which the extensional effective viscosity
1087: is of order $\delta^{-1/2}$. An example of such an array is presented in section 7.5.
1088:
1089:
1090: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1091: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1092: \section{Effective shear viscosity}
1093: In this Section we show that in dimension two, the asymptotic order of the shear
1094: effective viscosity $\mu^\star$ is $\delta^{-1/2}$, while
1095: the local lubrication analysis predicts
1096: the rate $\delta^{-3/2}$. In three
1097: dimensions, $\delta^{-1/2}$ and $\delta^{-3/2}$ should be replaced by,
1098: respectively,
1099: $\ln\delta$ and $\delta^{-1}$ (see \cite{BBP}).
1100: The local analysis in three dimensions predicts that the asymptotics of the
1101: shear effective viscosity $\mu^\star$ (see (\ref{shev2}))
1102: should be of order $\delta^{-1}$, but
1103: numerical simulations in \cite{Brady} and experimental results
1104: in \cite{ShP} and \cite{VW} show that random suspensions in shear flow have effective viscosity
1105: of order $\ln \delta$. Our estimate $\mu^\star=O(\delta^{-1/2})$ is therefore
1106: in agreement with the three-dimensional results in \cite{Brady},
1107: up to the difference in the dimension of the space.
1108:
1109: The decrease in the asymptotic order of $\mu^\star$ is a global
1110: phenomenon, which shows that
1111: the local analysis could be misleading, and analysis of the entire
1112: particle array leads to qualitatively different results.
1113: This fact can be explained as follows.
1114: The "strong" blow up rate ($\delta^{-1}$ in three dimensions and $\delta^{-3/2}$ in two
1115: dimensions) is obtained using classical lubrication techniques applied to two
1116: particles $D^i, D^j$ whose translational and angular velocities $\TI, \TJ, \omega^i, \omega^j$
1117: are {\it independent of $\delta$}. However,
1118: the network analysis of the ensemble of particles,
1119: interacting with each other and with the external boundary, shows that the above velocities
1120: may depend on $\delta$. In the case of shear boundary conditions, the right hand side of the
1121: network equations (\ref{bal-force}), (\ref{bal-torque}) is
1122: of order $\delta^{-1/2}$, whereas the matrix of the network
1123: equations is of order $\delta^{-3/2}$. The solution of the network equations
1124: is thus small (of order $\delta$) which makes local and global dissipation rates small.
1125:
1126: When the boundary conditions are given by (\ref{shear-fbc}), the vectors ${\bf R}^i$ in
1127: (\ref{lead-rhs}) are zero (since ${\bf q}^i\cdot{\bf e}_1=0$ for all $i=1,2, \ldots, N$).
1128: Consequently,
1129: the right hand side ${\bf f}$ in (\ref{compls}) is zero. The functional $\widehat Q$ reduces to
1130: $A{\bf z}\cdot {\bf z}$ which is clearly zero for every solution of the homogeneous system
1131: $A{\bf z}=0$. Since ${\bf z}=0$ is an admissible trial vector for $\widehat Q$,
1132: $\widehat E=\min \widehat Q=0$, and from (\ref{lead-term}) we obtain
1133: \begin{equation}
1134: \label{shear-diss}
1135: \mu^\star= C E_{net}=O(\delta^{-1/2}),
1136: \end{equation}
1137: where $C=2\gamma^{-2}|\Omega|^{-1}$ (see (\ref{shev2})).
1138: The same conclusion can be obtained by directly estimating the minimum of
1139: the form
1140: $Q$. Consider first the shear boundary conditions (\ref{shear-fbc}) with $\gamma=1$
1141: and a simple example of a two-disk network on Fig. 4.
1142: %%%% FIGURE%%%%%%%
1143: \begin{figure}%[htbp]
1144: \label{chain}
1145: \begin{center}
1146: \includegraphics{3path1.eps}
1147: \caption{A two-disk network with shear boundary conditions.}
1148: \end{center}
1149: \end{figure}
1150: %%%%FIGURE%%%%%%%%%%%%%%%%%%%%%%%%%%%
1151: The functional $Q$ for this example has the form
1152: \begin{equation}
1153: \label{shearex1}
1154: \begin{array}{ccc}
1155: Q &= & \hspace{-0.3cm}\delta^{-3/2}\left\{
1156: C_{sp}^{12}\left[({\bf T}^1-{\bf T}^2)\cdot {\bf q}^{12}\right]^2+
1157: C_{sp}^1\left[({\bf T}^1-{\bf e}_1)\cdot {\bf e}_2\right]^2+
1158: C_{sp}^2\left[({\bf T}^2+{\bf e}_1)\cdot {\bf e}_2\right]^2\right\}\\
1159: &&\hspace{-0.65cm} +\delta^{-1/2}\left\{C_{sh}^{12}\left[({\bf T}^1-{\bf T}^2)\cdot {\bf p}^{12}+a\omega_1+
1160: a\omega_2\right]^2+C_{rot}^{12}a^2(\omega_1-\omega_2)^2\right\}\hspace{2cm}\\
1161: &&\hspace{-2.3cm} +\delta^{-1/2}\left\{C_{sh}^1\left[({\bf T}^1-{\bf e}_1)\cdot {\bf e}_1+a\omega_1\right]^2+
1162: C_{sh}^2\left[({\bf T}^2+{\bf e}_1)\cdot {\bf e}_1+a\omega_2\right]^2\right\}\\
1163: & &\hspace{-7.25cm}+\delta^{-1/2}\left\{4C_{rot}^1a^2{\omega_1}^2+
1164: 4C_{rot}^2a^2{\omega_2}^2\right\}.
1165: \end{array}
1166: \end{equation}
1167: The dissipation rate $E$ is the minimum of $Q$. Hence, for any collection
1168: ${\bf T}^1, {\bf T}^2, \omega_1, \omega_2$, we have $E\leq Q({\bf T}^1,
1169: {\bf T}^2, \omega_1, \omega_2)$.
1170: In particular, choosing ${\bf T}^1={\bf T}^2={\bf 0}, \omega_1=\omega_2=0$ in
1171: (\ref{shearex1}) we obtain
1172: $$
1173: E\leq \delta^{-1/2}(C^1_{sh}+C^2_{sh})
1174: $$
1175: Since $C_{sh}^1, C_{sh}^2$ are independent of $\delta$, the blow up rate of $E$ is
1176: at most $\delta^{-1/2}$.
1177:
1178: Next, consider the general case.
1179: Since for shear boundary condition $\ug \bot {\bf q}^i$ for all $i\in I$, we obtain
1180: \begin{equation}
1181: \label{gen-ex}
1182: \begin{array}{lll}
1183: Q &= & {\displaystyle
1184: \sum_{i=1}^N} {\displaystyle \sum_{{\tiny \begin{array}{c}j \in {\cal
1185: N}_i\\ j < i
1186: \end{array}}}}\left\{
1187: \delta^{-3/2} C^{ij}_{sp}\left[(\TI-\TJ) \cdot {\bf
1188: q}^{ij}\right]^2\right\}+\\
1189: & &
1190: {\displaystyle
1191: \sum_{i=1}^N} {\displaystyle \sum_{{\tiny \begin{array}{c}j \in {\cal
1192: N}_i\\ j < i
1193: \end{array}}}}\left\{\delta^{-1/2} C^{ij}_{sh}\left[(\TI -\TJ)\cdot
1194: {\bf p}^{ij} + a{\om}^i + a{\om}^{j} \right]^2 \right.
1195: +\left. \delta^{-1/2} C^{ij}_{rot}a^2\left({\om}^i -
1196: {\om}^j\right)^2 \right\} + \\
1197: & & {\displaystyle
1198: \sum_{i \in I}} \left\{ \delta^{-3/2} C^{i}_{sp}\left[\TI\cdot{\bf q}^{i}
1199: \right]^2 +\delta^{-1/2} C^{i}_{rot}a^2\left(2 {\om}^i\right)^2 \right\} + \\
1200: & &
1201: {\displaystyle
1202: \sum_{i \in I}}\left\{\delta^{-1/2}
1203: C_{sh}^i \left[(\TI\cdot {\bf p}^{i}-{\bf g}\cdot{\bf p}^i +
1204: a{\om}^i
1205: \right]^2 \right\},
1206: \end{array}
1207: \end{equation}
1208:
1209: Choosing the trial vectors $\TI$ as $\TI=0, i=1,2, \ldots, N$ and $\omega^i=0$, $i=1,2, \dots, N$
1210: we obtain
1211: \begin{equation}
1212: \label{sp-en-shear}
1213: E_{net}=\min Q\leq Q({\bf T}^i={\bf 0}, \omega^i=0)=\delta^{-1/2} \sum_{i\in I}
1214: C_{sh}^{i}(\ug\cdot {\bf p}^{i})^2 =\gamma^2 \delta^{-1/2}\sum_{i \in I}C_{sh}^i.
1215: \end{equation}
1216: where we used the shear boundary conditions (\ref{shear-fbc}).
1217: The sum in the right hand side of (\ref{sp-en-shear})
1218: is independent of $\delta$ and the shear rate $\gamma$.
1219: This estimate shows that under the boundary conditions
1220: (\ref{shear-fbc}),
1221: the effective viscosity blows up as $\delta^{-1/2}$ at most.
1222: More precisely, we have the following\newline
1223: \noindent
1224: {\bf Conclusion.} {\it Effective shear
1225: viscosity of a concentrated suspension
1226: has the blow up rate $\delta^{-1/2}$ in two dimensions, that is
1227: $$
1228: \mu^\star\leq C \delta^{-1/2}\;\;\;{\rm as}~\delta\to 0,
1229: $$
1230: with $C$ independent of $\delta, \gamma$.}
1231:
1232:
1233:
1234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1235: \section{Extensional effective viscosity}
1236: \subsection{Simplification of boundary conditions and outline of the method}
1237: In this Section we show that for the extensional boundary conditions, the leading term
1238: in the asymptotics of the dissipation rate $E$ from (\ref{lead-term}) may or may not be zero
1239: depending on the geometry of
1240: a particle array, that is, the leading term in the asymptotics of the effective
1241: extensional viscosity is either of order $\delta^{-3/2}$ (strong blow up) or $\delta^{-1/2}$
1242: (weak blow up). We provide two geometric conditions on the network graph which ensure
1243: strong blow up.
1244:
1245: In a planar steady extensional flow of the effective fluid, the rate of strain
1246: tensor is
1247: \begin{equation}
1248: \label{eff-ext-strain}
1249: e({\bf v}^0)=
1250: \left(
1251: \begin{array}{cc}
1252: \epsilon & 0\\
1253: 0 & -\epsilon\\
1254: \end{array}
1255: \right),
1256: \end{equation}
1257: where $\epsilon$ denotes a constant extension rate. The corresponding
1258: velocity field is of the form
1259: \begin{equation}
1260: \label{ext-eff-vel}
1261: {\bf v}^0=(\epsilon x_1, -\epsilon x_2)^T,
1262: \end{equation}
1263: which gives the boundary conditions
1264: \begin{equation}
1265: \label{ext-eff-bc}
1266: {\bf v}^0=
1267: \begin{cases}
1268: (\epsilon x_1, -\epsilon)^T, \;\;{\rm when}\;x_2=1 \;\;({\rm on}\;\partial\Omega^+),&\\
1269: (\epsilon x_1, \epsilon)^T, \;\;{\rm when}\;x_2=-1 \;\;({\rm on}\;\partial\Omega^-).&\\
1270: \end{cases}
1271: \end{equation}
1272: We decompose ${\bf v}^0$ as
1273: \begin{equation}
1274: \label{ext-vsplit}
1275: {\bf v}^0={\bf v}^0_{vc}+{\bf v}^0_{ge},
1276: \end{equation}
1277: where
1278: ${\bf v}^0_{vc}$ is a vertical contraction velocity
1279: satisfying
1280: \begin{equation}
1281: \label{contr-bc}
1282: {\bf v}^0_{vc}={\bf g}_{vc}=
1283: \begin{cases}
1284: -\epsilon{\bf e}_2~~~{\rm on}~\bound^+, &\\
1285: \;\;\;\epsilon {\bf e}_2~~~{\rm on}~\bound^-, &\\
1286: \end{cases}
1287: \end{equation}
1288: and ${\bf v}^0_{ge}$ is the horizontal extension velocity field
1289: with the boundary conditions given by
1290: \begin{equation}
1291: \label{ext-bc}
1292: {\bf v}^0_{ge}={\bf g}_{ge}
1293: \begin{cases}
1294: \epsilon x_1{\bf e}_1~~~{\rm on}~\bound^+, &\\
1295: \epsilon x_1 {\bf e}_1~~{\rm on}~\bound^-. &\\
1296: \end{cases}
1297: \end{equation}
1298: Since
1299: \begin{equation}
1300: \label{ext-gsplit}
1301: {\bf g}={\bf g}_{vc}+{\bf g}_{ge},
1302: \end{equation}
1303: and ${\bf g}_{ge}\bot {\bf q}^i, i\in I$, the value of $\widehat Q$ in
1304: (\ref{lead-funct}) does not change when ${\bf g}$ in (\ref{lead-funct}) is replaced by
1305: ${\bf g}_{vc}$. Hence,
1306: \begin{equation}
1307: \label{q-vext}
1308: Q({\bf T}^i, \omega^i, {\bf g})=\delta^{-3/2}\widehat Q({\bf T}^i,{\bf g}_{vc})+
1309: \delta^{-1/2}Q^\prime({\bf T}^i, \omega^i, {\bf g}_{vc}+{\bf g}_{ge}),
1310: \end{equation}
1311: To determine the rate of blow up of the dissipation rate, we need
1312: to analyze the minimizers of $\widehat Q({\bf T}^i,{\bf g}_{vc})$. The estimate (\ref{lead-term})
1313: implies that the second term in the right hand side of (\ref{q-vext}) is of order $\delta^{-1/2}$
1314: (at most). Since $\widehat Q$ is independent of $\delta$, its minimizing vectors
1315: $\widehat \TI$ are also $\delta$-independent. Consequently, the blow up rate of the
1316: dissipation
1317: depends on whether the minimum of $\widehat Q(\TI, {\bf g}_{vc})$ is positive. If it is,
1318: the extensional effective
1319: viscosity $\lambda^\star$ is of order $\delta^{-3/2}$, otherwise $\lambda^\star$ grows no
1320: faster
1321: than $\delta^{-1/2}$. Which type of behavior occurs, depends on the validity of the estimate
1322: (\ref{crucial}). As mentioned in section 5, (\ref{crucial}) may fail for certain particle
1323: arrays. In this section we provide geometric conditions which insure positivity of
1324: $\min \widehat Q$,
1325: and give examples of networks for which this minimum is zero. The principal conclusion
1326: here is that extensional viscosity of suspensions with a comparable volume fraction of particles
1327: may vary by an order of magnitude in the inter-particle distance, depending
1328: on the geometry of a particle array.
1329:
1330:
1331: Our method of analysis is based on the following simple observation. The form $\widehat Q(\TI, {\bf g}_{vc})$
1332: is a sum of non-negative terms, namely
1333: \begin{equation}
1334: \label{whq-simpl}
1335: \widehat Q(\TI, {\bf g}_{vc})=
1336: \displaystyle{\frac 12 \sum_{i=1}^N \sum_{j\in {\cal N}_i} C^{ij}_{sp}
1337: (({\bf T}^i-{\bf T}^j)\cdot{\bf q}^{ij})^2+\sum_{i\in I} C_{sp}^{i}
1338: ((\TI-{\bf g}_{vc})\cdot {\bf q}^{i})^2.}
1339: \end{equation}
1340: This shows that $\min \widehat Q(\TI, {\bf g}_{vc})=0$ if and only if the minimizing vectors
1341: ${\TI}, i=1,2, \dots, N$ satisfy the system of equations
1342: \begin{equation}
1343: \label{ns}
1344: \begin{array}{c}
1345: ({\bf T}^i-{\bf T}^j)\cdot{\bf q}^{ij}=0,\;\;\;\;\;i=1,2..N, j\in {\cal N}_i,\\
1346: \hspace{-2.2cm}({\TI}-{\bf g}_{vc})\cdot {\bf q}^{i}=0, \;\;\;\;\;i\in I.
1347: \end{array}
1348: \end{equation}
1349: Hence, if (\ref{ns}) does not have solutions, (\ref{crucial}) must hold.
1350: Also, it should be noted that if the estimate (\ref{crucial}) holds for
1351: some subgraph of the network, then it also holds for the whole network.
1352: This observation can be used to reduce the network to
1353: a simpler graph, for which it is easier to prove (\ref{crucial}).
1354:
1355:
1356: %%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1357: \subsection{Simple examples}
1358: Suppose that the boundary conditions are given by (\ref{contr-bc}) with $\epsilon=-1$.
1359: In this section we present three simple examples of networks with small number of vertices.
1360: One of these networks
1361: has $\widehat E=0$, and for the other two examples $\widehat E>0$.
1362:
1363: \noindent
1364: {\bf Example 1.} This an example of the network for which $\widehat E=0$.
1365: To demonstrate this, we show that
1366: there is a nontrivial particle velocity vector ${\bf z}$ such that $\widehat Q({\bf z})=0$.
1367: Consider the network on Fig. 5.
1368: %%%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%
1369: \begin{figure}%[htbp]
1370: \label{ex1}
1371: \begin{center}
1372: \includegraphics{ex1.eps}
1373: \caption{A network of four vertices with $\widehat E=0$.}
1374: \end{center}
1375: \end{figure}
1376: The vectors ${\bf q}^{ij}$ are defined as follows.
1377: \begin{equation}
1378: \label{qi}
1379: \begin{array}{cc}
1380: \hspace{-3cm}{\bf q}^1={\bf e}_2, & \hspace{-4cm}{\bf q}^4=-{\bf e}_2, \\
1381: {\bf q}^{12}={\bf q}^{34}=\frac{1}{\sqrt{2}}({\bf e}_1-{\bf e}_2),& \\
1382: {\bf q}^{13}={\bf q}^{24}=-\frac{1}{\sqrt{2}}({\bf e}_1+{\bf e}_2) & \\
1383: \end{array}
1384: \end{equation}
1385: Next, define ${\bf T}^i$ as follows.
1386: \begin{equation}
1387: \label{ti}
1388: {\bf T}^1={\bf e}_2,\;\;{\bf T}^2=-{\bf e}_1,\;\;\;
1389: {\bf T}^3={\bf e}_1,\;\;\;
1390: {\bf T}^4=-{\bf e}_2.
1391: \end{equation}
1392: The functional $\widehat Q$ corresponding to this network is
1393: \begin{equation}
1394: \label{ex1q}
1395: \begin{array}{ccc}
1396: \widehat Q & = &C^1_{sp}\left[({\bf T}^1-{\bf e}_2)\cdot{\bf q}^1\right]^2+
1397: C^{12}_{sp}\left[({\bf T}^1-{\bf T}^2)\cdot{\bf q}^{12}\right]^2+\\
1398: & & C^{13}_{sp}\left[({\bf T}^1-{\bf T}^3)\cdot{\bf q}^{13}\right]^2+
1399: C^{24}_{sp}\left[({\bf T}^2-{\bf T}^{4})\cdot{\bf q}^{24}\right]^2+\\
1400: & & C^{34}_{sp}\left[({\bf T}^3-{\bf T}^4)\cdot{\bf q}^{34}\right]^2+
1401: C^4_{sp}\left[({\bf T}^4+{\bf e}_2)\cdot{\bf q}^4\right]^2.
1402: \end{array}
1403: \end{equation}
1404: When ${\bf T}^i$ are defined by (\ref{ti}), all the scalar product in brackets in (\ref{ex1q})
1405: are zero, and therefore $\min\widehat Q=0$.\\
1406: \noindent
1407: {\bf Example 2.} For the network of three vertices in Fig. 6, $\min \widehat Q>0$. To show this,
1408: %%%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%
1409: \begin{figure}%[htbp]
1410: \label{ex1}
1411: \begin{center}
1412: \includegraphics{ex2.eps}
1413: \caption{A network of three vertices with $\widehat E>0$.}
1414: \end{center}
1415: \end{figure}
1416: consider the system corresponding to the general system (\ref{ns}).
1417: \begin{equation}
1418: \label{ex2.1}
1419: \begin{array}{c}
1420: ({\bf T}^1+{\bf e}_2)\cdot {\bf e}_2=0,\;\;\;({\bf T}^1-{\bf T}^2)\cdot {\bf q}^{12}=0,\;\;\;\;
1421: ({\bf T}^1-{\bf T}^3)\cdot {\bf q}^{13}=0,\\
1422: ({\bf T}^2-{\bf T}^3)\cdot {\bf q}^{23}=0,
1423: \;\;\;\;({\bf T}^2-{\bf e}_2)\cdot {\bf e}_2=0,\;\;\;\;\;
1424: ({\bf T}^3-{\bf e}_2)\cdot {\bf e}_2=0.\\
1425: \end{array}
1426: \end{equation}
1427: We prove that the system (\ref{ex2.1}) has no solutions. Indeed, the last two equations
1428: imply ${\bf T}^2=t_2{\bf e}_1+{\bf e}_2$, ${\bf T}^3=t_3{\bf e}_1+{\bf e}_2$, for some scalars
1429: $t_2, t_3$. Next, the first equation in the second row of (\ref{ex2.1}) yields
1430: $(t_2-t_3){\bf e}_1\cdot{\bf q}^{23}=0$, and thus $t_2=t_3=t$. Substituting
1431: ${\bf T}^2={\bf T}^3=t{\bf e}_1+{\bf e}_2$ into the second
1432: and third equations in the first row of (\ref{ex2.1}) we obtain
1433: \begin{equation}
1434: \label{ex2.2}
1435: \begin{array}{c}
1436: ({\bf T}^1-t{\bf e}_1-{\bf e}_2)\cdot{\bf q}^{12}=0,\\
1437: ({\bf T}^1-t{\bf e}_1-{\bf e}_2)\cdot{\bf q}^{13}=0.\\
1438: \end{array}
1439: \end{equation}
1440: Since ${\bf q}^{12}, {\bf q}^{13}$ are non-collinear, (\ref{ex2.2}) yields
1441: ${\bf T}^1=t{\bf e}_1+{\bf e}_2$, which contradicts the first equation in the first row of
1442: (\ref{ex2.1}).\\
1443: \noindent
1444: {\bf Example 3.} Next, consider a rectangular network of four vertices in Fig. 7.
1445: %%%%%%%%%%%%%%%FIGURE%%%%%%%%%%%%%%
1446: \begin{figure}%[htbp]
1447: \label{ex1}
1448: \begin{center}
1449: \includegraphics{ex3.eps}
1450: \caption{A network of four vertices with $\widehat E>0$.}
1451: \end{center}
1452: \end{figure}
1453: The system (\ref{ns}) for this example becomes
1454: \begin{equation}
1455: \label{ex3.1}
1456: \begin{array}{c}
1457: ({\bf T}^1-{\bf T}^2)\cdot {\bf e}_1=0,\;\;\;\;
1458: ({\bf T}^1-{\bf T}^4)\cdot {\bf e}_2=0,\;\;\;\;
1459: ({\bf T}^2-{\bf T}^3)\cdot {\bf e}_2=0,\;\;\;\;
1460: ({\bf T}^3-{\bf T}^4)\cdot {\bf e}_1=0,\;\;\;\;\\
1461: \hspace{-0.6cm}({\bf T}^1+{\bf e}_2)\cdot {\bf e}_2=0,\;\;\;\;\;\;
1462: ({\bf T}^2+{\bf e}_2)\cdot {\bf e}_2=0,\;\;\;\;
1463: ({\bf T}^3-{\bf e}_2)\cdot {\bf e}_2=0,\;\;\;\;
1464: ({\bf T}^4-{\bf e}_2)\cdot {\bf e}_2=0.\\
1465: \end{array}
1466: \end{equation}
1467: The last two equations in the second row of (\ref{ex3.1}) yield ${\bf T}^3=t_3{\bf e}_1+{\bf e}_2$,
1468: ${\bf T}^4=t_4{\bf e}_1+{\bf e}_2$, with some scalars $t_3, t_4$. Next, the second and third
1469: equations in the first row of
1470: (\ref{ex3.1}) produce ${\bf T}^1=t_1{\bf e}_1+{\bf e}_2$, ${\bf T}^2=t_2{\bf e}_1+{\bf e}_2$,
1471: which contradict, respectively, the first and second equations in the second row.
1472:
1473: The three examples above seem to indicate
1474: that two basic building blocks for networks with $\widehat E>0$ (strong blow up) are
1475: triangles (Fig. 6) and rectangles aligned with the edges of $\Omega$ (Fig. 7).
1476: Misaligned rectangular structures such as shown in Fig. 8 would produce $\widehat E=0$ (weak blow up).
1477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1478: \subsection{Quasi-triangulated graphs}
1479: \subsubsection{Definition of quasi-triangulated graphs. Solvability of the system (\ref{compls})}
1480: The network $\Gamma$ partitions $\Omega$ into a a
1481: disjoint union of convex polygons, which are called
1482: {\it Delaunay cells}. When points ${\bf x}^j$ are distributed randomly in $\Omega$, the interior
1483: Delaunay cells are
1484: typically triangles. This simple but important fact can be explained as follows.
1485: The edges of Voronoi tessellation are
1486: perpendicular bisectors of the edges of Delaunay cells.
1487: If an interior Delaunay cell is, for instance, a quadrilateral, then
1488: any two vertices lying on a diagonal cannot be neighbors, and therefore
1489: the point of intersection of four edges of the Voronoi tessellation must
1490: be equidistant from the four vertices. This means that a convex quadrilateral may be a Delaunay
1491: cell only if all four vertices lie on a circle. When the vertices of the network
1492: are randomly placed, the likelihood of four (or more) points lying on the same
1493: circle is small. It is natural to call such cells the defect cells.
1494: Therefore, most of the interior Delaunay cells of a random
1495: network are triangles. Other polygonal cells (quadrilateral, pentagonal etc.)
1496: are typically small in number, isolated and are likely to be unstable in the actual flow.
1497:
1498:
1499: Since the exterior edges of the network are vertical, the cells adjacent to the boundary
1500: are typically quadrilateral. An example of a generic network is shown on Fig. 8.
1501:
1502:
1503: %%%% FIGURE%%%%%%%
1504: \begin{figure}%[htbp]
1505: \label{del1}
1506: \begin{center}
1507: \includegraphics{delaunay2.eps}
1508: \caption{An example of the network. The
1509: shaded quadrilateral represents a defect cell.}
1510: \end{center}
1511: \end{figure}
1512:
1513:
1514:
1515:
1516: Next, we define a broad class of graphs containing
1517: generic Delaunay-type networks. The graphs in this class are
1518: called {\it quasi-triangulated}, because of the presence of triangular cells.
1519: The number of triangular cells may be relatively small, as indicated by the examples
1520: below. We show that for quasi-triangulated graphs the leading term in the asymptotics
1521: can be effectively computed by minimizing the functional $\widehat Q$, that is the
1522: minimum is unique and can be found by solving the linear system (\ref{compls}).
1523: We also show that the crucial estimate (\ref{crucial}) holds even for more general
1524: graphs which contain a "spanning" quasi-triangulated subgraph.
1525:
1526:
1527: Given an arbitrary network graph $\Gamma$, we define its
1528: maximal quasi-triangulated
1529: subgraph $\Gamma_M$ by the following
1530: iterative procedure.\newline
1531: \noindent
1532: {\it Step 1.} Consider interior vertices which are connected to
1533: $\bound^-$ and call these vertices {\it generation one} vertices. All interior edges
1534: connecting these vertices are {\it generation one} edges. Add all generation one
1535: edges and vertices to the subgraph.\newline
1536: \noindent
1537: {\it Step 2.} Consider all remaining vertices which are connected to the vertices
1538: of the subgraph
1539: by at least two non-collinear edges. These vertices and edges belong to generation
1540: two. Add them to the subgraph. Note that the non-collinearity condition
1541: leads to formation of "supportive triangles". \newline
1542: \noindent
1543: {\it Step 3.} Repeat step 2 until no more vertices can be added.
1544:
1545:
1546: If the maximal quasi-triangulated subgraph $\Gamma_M$
1547: contains all interior vertices of $\Gamma$, we call the graph $\Gamma$
1548: {\it quasi-triangulated}. It turns out that for quasi-triangulated graphs
1549: the system (\ref{compls}) is "almost uniquely" solvable, that is
1550: two solutions differ by a vector of the form $t{\bf w}_0$, where ${\bf w}_0$ is defined by (\ref{w-not}), and
1551: the value of the form $Q$ for these solutions is the same.
1552:
1553:
1554:
1555: \begin{proposition}
1556: \label{quasi}
1557: Suppose that the network graph $\Gamma$ is quasi-triangulated. Then there is a
1558: unique solution of the system (\ref{compls}), up to a horizontal translation.
1559: \end{proposition}
1560:
1561: This proposition is proved in Appendix B.
1562:
1563:
1564:
1565: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1566: \subsubsection{Examples of quasi-triangulated graphs}
1567: First we observe that a restriction to $\Omega$ of a periodic rectangular lattice is not
1568: quasi-triangulated. By contrast, a periodic triangular lattice restricted to $\Omega$ is
1569: quasi-triangulated.
1570: (see Fig. 9).
1571:
1572:
1573:
1574:
1575:
1576: %%%% FIGURE%%%%%%%
1577: \begin{figure}%[htbp]
1578: \label{del1}
1579: \begin{center}
1580: \includegraphics{del1.eps}
1581: \caption{A rectangular graph on the left is not quasi-triangulated. Clearly, in this case
1582: $\Gamma_F=\Gamma^-$.
1583: The triangular graph on the right
1584: is quasi-triangulated.}
1585: \end{center}
1586: \end{figure}
1587:
1588:
1589: In fact, if a network $\Gamma$ is not periodic, but all of its interior Delaunay cells
1590: are triangles,
1591: then $\Gamma$ is quasi-triangulated.
1592: The converse is false (see an example in Fig. 11). This means that the
1593: quasi-triangulation property is
1594: more general than triangulation property. In \cite{BBP} we considered a class of graphs
1595: containing a triangulated path (see Fig. 10). In Fig 11, we present two examples
1596: showing that a quasi-triangulated graph need not contain a triangulated path. On
1597: the other hand,
1598: a graph containing a triangulated path is not necessarily quasi-triangulated.
1599: However, when a triangulated path exists, it must be contained in the maximal
1600: quasi-triangulated subgraph. Therefore, in this case the maximal
1601: quasi-triangulated subgraph is spanning ("extends from top to bottom"). Below in section
1602: 7.4.1 we show that this condition is sufficient for positivity of the minimum
1603: of $\widehat Q$.
1604:
1605: %%%% FIGURE%%%%%%%
1606: \begin{figure}%[htbp]
1607: \label{del3}
1608: \begin{center}
1609: \includegraphics{del3.eps}
1610: \caption{A triangulated path.}
1611: \end{center}
1612: \end{figure}
1613:
1614: %%%% FIGURE%%%%%%%
1615: \begin{figure}%[htbp]
1616: \label{del2}
1617: \begin{center}
1618: \includegraphics{del2.eps}
1619: \caption{The graph on the left is quasi-triangulated but does not contain a
1620: triangulated path. The graph
1621: on the right contains a triangulated path, but is not quasi-triangulated.}
1622: \end{center}
1623: \end{figure}
1624:
1625:
1626:
1627:
1628: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1629: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1630: \subsection{Strong blow up of $\lambda^\star$. Percolating rigidity networks}
1631: \subsubsection{Quasi-triangulated subgraphs}
1632:
1633: For the rest of this section, we consider the steady flow of the suspension
1634: corresponding to
1635: the boundary conditions (\ref{contr-bc}) with the extension
1636: rate $\ep=1$.
1637:
1638:
1639: We shall say that a network is a {\it percolating rigidity graph} when
1640: the crucial estimate (\ref{crucial}) holds.
1641: In this subsection we show that quasi-triangulated graphs are percolating rigidity graphs.
1642: Moreover, a graph has percolating rigidity even when it is not
1643: quasi-triangulated but contains a spanning quasi-triangulated subgraph.
1644:
1645:
1646: \begin{proposition}
1647: \label{tricon}
1648: Suppose that the boundary conditions are given by (\ref{contr-bc}) and the network
1649: graph $\Gamma$ contains a spanning quasi-triangulated subgraph. Then $\Gamma$ is a percolating rigidity graph. Consequently, the
1650: extensional effective viscosity $\lambda^\star$ of suspensions corresponding to such networks
1651: is $O(\delta^{-3/2})$.
1652: \end{proposition}
1653:
1654: The proof of this proposition is given in Appendix C.
1655:
1656:
1657: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1658: \subsubsection{Networks containing a vertical path. Periodic square networks}
1659: We now present another class of percolating rigidity graphs. Namely, these
1660: are graphs that contain a path connecting $\bound^+$ and $\bound^-$, such that
1661: all edges in this path are oriented along ${\bf e}_2$-direction
1662: (vertical). The simplest representative of this class of graphs is a periodic square
1663: lattice.
1664: A periodic square graph does not contain a spanning quasi-triangulated
1665: subgraph (in this case $\Gamma_M$ is just a path
1666: which consists of the vertices adjacent to $\bound^-$, see Fig. 9). However, the
1667: estimate (\ref{crucial}) holds for a periodic square graph such as the one shown on Fig. 9.
1668: This follows from the more general criterion
1669:
1670: \begin{proposition}
1671: \label{cubic}
1672: Suppose that a network graph $\Gamma$ contains a path $\Gamma^\pm$ such that
1673:
1674: \noindent
1675: i) it connects $\bound^+$ and $\bound^-$,
1676:
1677:
1678: \noindent
1679: ii) all edges of $\Gamma^\pm$ are vertical.
1680:
1681:
1682:
1683: Then $\Gamma$ is a percolating rigidity graph.
1684: \end{proposition}
1685:
1686: As noted above, if a path $\Gamma^\pm$ has percolating rigidity, then
1687: the "larger" graph $\Gamma$ is also a percolating rigidity graph. Therefore, we only need
1688: to show that a path of vertical edges has percolating rigidity, that is, the quadratic form
1689: $\widehat Q$ corresponding to such path is positive-definite.
1690:
1691: For simplicity, consider a
1692: path containing three
1693: vertices. The argument can be directly generalized to an arbitrary number of vertices.
1694: As before, (\ref{crucial}) will hold if the corresponding system (\ref{ns}) has
1695: no solutions. The latter now has the form
1696: \begin{equation}
1697: \label{vpath1}
1698: \begin{array}{c}
1699: ({\bf T}^1-{\bf e}_2)\cdot{\bf e}_2=0,\\
1700: ({\bf T}^1-{\bf T}^2)\cdot{\bf e}_2=0,\\
1701: ({\bf T}^2-{\bf T}^3)\cdot{\bf e}_2=0,\\
1702: ({\bf T}^3+{\bf e}_2)\cdot{\bf e}_2=0,\\
1703: \end{array}
1704: \end{equation}
1705: Introduce new unknown vectors $\widehat {\bf T}^i={\bf t}^i+{\bf e}_2$ $i=1,2,3$.
1706: The last equation of (\ref{vpath1}) yields $\widehat {\bf T}^3=t_3 {\bf e}_1$,
1707: where $t_3$ is a scalar. Then from the third equation we obtain $\widehat {\bf T}^2=
1708: t_2{\bf e}_1$, and then the second equation yields $\widehat{\bf T}^1=t_1{\bf e}_1$, which
1709: contradicts the first equation. The contradiction shows that the system (\ref{vpath1})
1710: has no solutions. The argument can be easily modified to show that if at least one
1711: of the edges of a path is non-vertical, then this path does not have percolating rigidity.
1712:
1713:
1714: Proposition \ref{cubic} shows that for rectangular lattices oriented
1715: parallel to the edges of $\bound$, asymptotics of extensional effective viscosity
1716: is of order $\delta^{-3/2}$ which corresponds to strong blow up.
1717: Asymptotic formulas in \cite{Keller} also predict strong blow up
1718: of the extensional viscosity for cubic lattices. Thus our results are consistent
1719: with the results in \cite{Keller}. We refer to section 8 for a more detailed
1720: comparison.
1721:
1722:
1723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1724: \subsection{Weak blow up}
1725: In this section we present an example of a particle array for which
1726: the leading term in the asymptotics of $\lambda^\star$ is zero.
1727: Roughly speaking, the array in question is a rectangular lattice
1728: rotated so that none of its interior edges are vertical.
1729: Let ${\bf k}$ denote a unit vector non-collinear to either ${\bf e}_1$ or ${\bf e}_2$. The interior
1730: edges of the rectangular network in Fig. 12 are either parallel or perpendicular
1731: to ${\bf k}$, while the prescribed boundary velocities are parallel to
1732: ${\bf e}_2$.
1733: This misalignment will lead to weak blow up.
1734: To show that
1735: the system (\ref{ns}) is solvable, we
1736: first consider a single path connecting $\bound^+$ and $\bound^-$.
1737: This path may be any of the three such paths in Fig. 12.
1738: The argument we use admits a straightforward generalization to
1739: a network with an arbitrary number of vertices.
1740: %%%% FIGURE%%%%%%%
1741: \begin{figure}%[htbp]
1742: \label{chain}
1743: \begin{center}
1744: \includegraphics{del4.eps}
1745: \caption{A rotated rectangular lattice of 12 vertices.}
1746: \end{center}
1747: \end{figure}
1748: %%%%FIGURE%%%%%%%%%%%%%%%%%%%%%%%%%%%
1749:
1750: The system (\ref{ns}) written for the path has the form
1751: \begin{equation}
1752: \label{npath1}
1753: \begin{array}{c}
1754: ({\bf T}^1-{\bf e}_2)\cdot{\bf e}_2=0,\\
1755: ({\bf T}^1-{\bf T}^2)\cdot{\bf k}=0,\\
1756: ({\bf T}^2-{\bf T}^3)\cdot{\bf k}=0,\\
1757: ({\bf T}^3-{\bf T}^4)\cdot{\bf k}=0,\\
1758: ({\bf T}^4+{\bf e}_2)\cdot{\bf e}_2=0,\\
1759: \end{array}
1760: \end{equation}
1761: For technical reasons it is convenient to introduce new unknowns ${\bf u}_1={\bf T}^1-{\bf e}_2$,
1762: ${\bf u}_{12}={\bf T}^1-{\bf T}^2$, ${\bf u}_{23}={\bf T}^2-{\bf T}^3$,
1763: ${\bf u}_{34}={\bf T}^3-{\bf T}^4$. The relations between ${\bf u}_{ij}$ and
1764: ${\bf T}^i$ are
1765: \begin{eqnarray}
1766: \label{relate}
1767: {\bf T}^1={\bf u}_1+2{\bf e}_2,\nonumber\\
1768: {\bf T}^2={\bf u}_1+2{\bf e}_2-{\bf u}_{12},\nonumber\\
1769: {\bf T}^3={\bf u}_1+2{\bf e}_2-{\bf u}_{12}-{\bf u}_{23},\nonumber\\
1770: {\bf T}^4={\bf u}_1+2{\bf e}_2-{\bf u}_{12}-{\bf u}_{23}-{\bf u}_{34}\label{last}.
1771: \end{eqnarray}
1772: From the first four equations of (\ref{npath1}) we see that ${\bf u}_1=t_1{\bf e}_1$, and
1773: ${\bf u}_{i, i+1}=t_{i,i+1}{\bf k}^\bot$, $i=1,2,3$, where
1774: ${\bf k}^\bot$ denotes a unit vector orthogonal to ${\bf k}$, and $t_1, t_{i,i+1}$ are
1775: scalars. The problems of solving
1776: (\ref{npath1}) is now reduced to finding $t_1, t_{i,i+1}$ such that the fifth equation
1777: of (\ref{npath1}) is satisfied. The fifth equation shows that
1778: \begin{equation}
1779: \label{t4}
1780: {\bf T}^4=t_4{\bf e}_1-{\bf e}_2,
1781: \end{equation}
1782: where $t_4$ is a scalar. Equating (\ref{t4}) and (\ref{last}) we obtain the equation
1783: for $t_1, t_{12}, t_{23}, t_{34}$ and $t_4$:
1784: \begin{equation}
1785: \label{ts}
1786: t_4{\bf e}_1=t_1{\bf e}_1+{\bf e}_2-(t_{12}+t_{23}+t_{34}){\bf k}^\bot.
1787: \end{equation}
1788: This yields two scalar equations
1789: \begin{equation}
1790: \label{first}
1791: 0=1-(t_{12}+t_{23}+t_{34}){\bf k}^\bot\cdot{\bf e}_2,
1792: \end{equation}
1793: and
1794: \begin{equation}
1795: \label{second}
1796: t_1-t_4=(t_{12}+t_{23}+t_{34}){\bf k}^\bot\cdot{\bf e}_1.
1797: \end{equation}
1798: The system of two equations (\ref{first}),(\ref{second}) for five unknowns has infinitely many
1799: nontrivial solutions as long as ${\bf k}^\bot\cdot {\bf e}_2\ne 0$ , that is, the interior
1800: edges of the path are non-vertical.
1801:
1802:
1803: At the next step of construction, we consider the full lattice on Fig. 6 containing 12 vertices.
1804: The interior edges of the graph are oriented either by the unit vector ${\bf k}$ as above
1805: (longitudinal edges), or
1806: by ${\bf k}^\bot$ (latitudinal edges).
1807: We view the graph as the union of three paths of longitudinal edges
1808: extending from $\bound^-$ to $\bound^+$,
1809: with latitudinal edges connecting
1810: these paths. To obtain the desired example, we choose the vectors ${\bf T}^i$ for one
1811: of the paths, as explained above. Then we prescribe the same vectors to the corresponding
1812: vertices of two remaining paths. Now, if two neighbors ${\bf x}^i$ , ${\bf x}^j$ belong
1813: to different paths, then ${\bf T}^i={\bf T}^j$, so the corresponding equation
1814: $({\bf T}^i-{\bf T}^j)\cdot{\bf k}^\bot=0$ of the system (\ref{ns}) is satisfied. When the neighbors
1815: ${\bf x}^i, {\bf x}^j$ belong to the same path, the corresponding equation
1816: $({\bf T}^i-{\bf T}^j)\cdot{\bf k}=0$ is satisfied by the choice of ${\bf T}^i, {\bf T}^j$.
1817: Therefore, the whole system (\ref{ns}) in this case has infinitely many nontrivial solutions.
1818: Each of these solutions makes $\widehat Q$, and thus the leading term in the asymptotics of
1819: $\lambda^\star$, zero.
1820:
1821: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1822: \section{Comparison with some results for periodic cubic arrays in dimension three}
1823: The main objective of network approximation is to define effective
1824: properties for non-periodic deterministic or random arrays. While techniques of
1825: periodic homogenization are well developed (\cite{BLP}, \cite{SP}, \cite{JKO} and references therein), non-periodic geometries are much less understood.
1826: In the end of this section, we compare our results applied in particular case
1827: of a periodic square array (in dimension two) with the results of \cite{Keller}
1828: obtained for cubic arrays in dimension three.
1829:
1830:
1831: The effective viscosity of an infinite periodic suspension obtained in
1832: \cite{Keller}
1833: is the fourth order tensor ${\mbox{\boldmath $\mu$}}^\star$ (in this section we use the
1834: notation from from \cite{Keller}).
1835: In an effective flow
1836: with the constant strain rate $\gamma$, the effective stress is
1837: \begin{equation}
1838: \label{k1}
1839: S_{ij}^0=2\mu^\star_{ijkl}\gamma_{kl}-P\delta_{ij},
1840: \end{equation}
1841: where $P$ is an effective pressure. The authors of \cite{Keller} obtained the following formula
1842: for the components of
1843: ${\mbox{\boldmath $\mu$}}^\star$:
1844: \begin{equation}
1845: \label{visc-tensor}
1846: \mu^\star_{ijkl}=\frac 12 \mu (1+\beta)(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}-
1847: \frac{2}{3}\delta_{ij}\delta_{kl})+\mu(\alpha-\beta)(\delta_{ijkl}-
1848: \frac 12 \delta_{ij}\delta_{kl}),
1849: \end{equation}
1850: where $\delta_{ijkl}=1$ if all indices are equal, otherwise $\delta_{ijkl}=0$.
1851: $\mu$ is the fluid viscosity, and $\alpha, \beta$ are functions of the small
1852: parameter $\epsilon$, related to the inter-particle distance $\delta$ as follows:
1853: \begin{equation}
1854: \label{del-ep}
1855: \epsilon=\frac{\delta}{2a+\delta}\approx \frac{\delta}{2a},\;\;\;{\rm as}\;\delta\to 0.
1856: \end{equation}
1857: When the effective flow is incompressible, $\gamma_{ii}=0$, so the formula (\ref{visc-tensor})
1858: simplifies to
1859: \begin{equation}
1860: \label{vtensor}
1861: \mu^\star_{ijkl}=\frac 12 \mu (1+\beta)(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk})+
1862: \mu(\alpha-\beta)\delta_{ijkl}.
1863: \end{equation}
1864:
1865: For simple cubic lattices,
1866: up to the terms of order $O(1)$ in $\epsilon$, (\cite{Keller})
1867: \begin{equation}
1868: \label{alpha}
1869: \alpha=\frac{3}{16}\pi \epsilon^{-1}+\frac{27}{80}\pi\ln \epsilon^{-1},
1870: \end{equation}
1871: and
1872: \begin{equation}
1873: \label{beta}
1874: \beta=\frac 14 \pi\ln \epsilon^{-1}.
1875: \end{equation}
1876:
1877: Suppose that the imposed effective flow is a steady shear with the velocity
1878: ${\bf v}=(\kappa x_3, 0, 0)$ (three-dimensional analogue of (\ref{shear-v})),
1879: where $\kappa>0$ is a constant shear rate.
1880: The components of the corresponding strain rate tensor ${\mbox{\boldmath $\gamma$}}$ are
1881: $\gamma_{13}=\gamma_{31}=\kappa>0$ and $\gamma_{ij}=0$ for other values of $(i,j)$.
1882: Then, using (\ref{vtensor}), we obtain from (\ref{k1}) that the only nonzero
1883: components of the effective deviatoric stress
1884: $2{\mbox{\boldmath $\mu$}}^\star{\mbox{\boldmath $\gamma$}}$ are
1885: \begin{equation}
1886: \label{k2}
1887: (2\mu^\star\gamma)_{13}=(2\mu^\star\gamma)_{31}=2\mu (1+\beta)\kappa.
1888: \end{equation}
1889: Since $\alpha$ is not present in (\ref{k2}), the nonzero components of the deviatoric
1890: effective stress are of order $\ln\epsilon^{-1}$ and thus the shear effective viscosity
1891: calculated by the three-dimensional analogue of
1892: definition (\ref{shev2}) is of order $\ln\epsilon^{-1}$ (weak blow up in dimension three). The weak blow up was not identified in \cite{Keller}, but
1893: it can be easily deduced from the formulas derived there.
1894:
1895: In the case of an extensional flow, velocity vector
1896: ${\bf v}=(\kappa x_1, \kappa x_2, -2\kappa x_3)$, (compare with (\ref{ef-v-ext})),
1897: where $\kappa>0$ is a constant extension
1898: rate. The strain rate tensor is
1899: \begin{equation}
1900: \label{ext-st-3d}
1901: {\mbox{\boldmath $\gamma$}}=
1902: \left(
1903: \begin{array}{ccc}
1904: \kappa & 0 & 0\\
1905: 0 & \kappa & 0\\
1906: 0 & 0 & -2\kappa \\
1907: \end{array}
1908: \right)
1909: \end{equation}
1910: Then, using (\ref{k1}) and (\ref{vtensor}) we obtain the deviatoric stress
1911: \begin{equation}
1912: \label{ext-stress}
1913: 2{\mbox{\boldmath $\mu$}}^\star{\mbox{\boldmath $\gamma$}}=
1914: \left(
1915: \begin{array}{ccc}
1916: L & 0 & 0\\
1917: 0 & L & 0\\
1918: 0 & 0 & -2L, \\
1919: \end{array}
1920: \right)
1921: \end{equation}
1922: where
1923: \begin{equation}
1924: \label{L}
1925: L=2\kappa\mu(1+\beta)+\kappa\mu(\alpha-\beta).
1926: \end{equation}
1927: Components of the deviatoric stress contain $\alpha$ and are therefore of order $\ep^{-1}$. Consequently, the
1928: extensional effective viscosity is of order $\ep^{-1}$
1929: (strong blow up in dimension three).
1930:
1931:
1932: Although Noonan and Keller did not address the issue of weak versus strong blow up
1933: for the effective viscosity in \cite{Keller}, the formulas (\ref{alpha})- (\ref{L})
1934: are consistent with the results for square lattices in Section 6
1935: of this paper in the following sense. If a periodicity cell corresponding to a simple
1936: cubic lattice is subjected to a uniform shear (extensional) flow,
1937: then the straightforward calculation presented above shows that the
1938: shear (extensional) effective
1939: viscosity exhibits weak (strong) blow up. However, for other lattice types such
1940: as BCC and FCC, formulas from \cite{Keller} imply strong blow up of both
1941: viscosities, while our approach leads to the weak blow up of the shear viscosity
1942: for all two-dimensional lattices. Perhaps, this can be attributed to the fact
1943: that in \cite{Keller} effective viscosity is defined in a different way, as
1944: explained in the introduction. In particular, our definition takes into account
1945: boundary effects which are known to be essential in rheological measurements
1946: (\cite{VW}, \cite{ShP}).
1947: Considerations based on infinite periodic lattices cannot capture these boundary
1948: effects. Futhermore, if our approach is applied to a rectangular periodic lattice
1949: in a rectangular domain chosen so that lattice periodicity is preserved, then
1950: our results are consistent with \cite{Keller}.
1951: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1952: \section{Conclusions}
1953: We have studied properties of the network approximation
1954: of the shear effective viscosity
1955: $\mu^\star$ and extensional effective viscosity $\lambda^\star$ of two-dimensional
1956: flows of concentrated suspensions
1957: with complex geometry. The flow domain was chosen to be a square, upper
1958: and lower sides of which represented the physical flow boundary, while
1959: the other two sides were assumed to be free surfaces.
1960:
1961: A small inter-particle
1962: distance parameter $\delta$ was used to describe the high concentration regime
1963: for particle arrays which are not necessarily periodic (i. g. random).
1964: In the recent paper \cite{Brady} by Sierou and Brady,
1965: the high frequency dynamic viscosity of concentrated
1966: suspensions
1967: (see \cite{VW}) was calculated as a function of volume fraction $\phi$ by means of accelerated Stokesian dynamics
1968: simulations. Numerical results indicate
1969: a singular behaviour of the effective viscosity as $\phi$ approaches
1970: the maximal close packing fraction $\phi_{rcp}$. We quote here from (\cite{Brady}):
1971: ''... The exact form of this singular behaviour
1972: is not known. Results from lubrication theory for cubic lattices would suggest
1973: that the singular form should consist both of $1/\epsilon$ and $\ln\epsilon$, where
1974: $\ep=1-(\phi/\phi_{rcp})^{1/3}$, but the relative amount of each term is unknown...
1975: As far as we are able to tell at this point, the $\ln\epsilon$ behaviour accurately describes
1976: the numerical data.''
1977:
1978: When the volume fraction is (approximately) constant over the flow region, $\ep\approx \delta$.
1979: One of the objectives of our investigation
1980: was to address the issue of the unexpectedly weak blow up in
1981: \cite{Brady}, and determine the asymptotic order of the effective
1982: viscosity coefficients as $\delta\to 0$.
1983: Our analysis of the shear viscosity $\mu^\star$, based on the discrete network approximation,
1984: showed that $\mu^\star=O(\delta^{-1/2})$ as $\delta\to 0$, while
1985: the formal estimate based of the local lubrication analysis would give
1986: a higher rate $O(\delta^{-3/2})$.
1987: In dimension three, the corresponding rates, given by the network approximation, are, respectively, $\ln\delta$ and $\delta^{-1}$ (see \cite{BBP}).
1988: Our analysis offers an explanation of the weak blow up of the shear viscosity. This analysis is also consistent with
1989: the numerical simulations in \cite{Brady} up to the difference in
1990: the space dimension.
1991:
1992:
1993: We also show that the asymptotic order of the extensional viscosity $\lambda^\star$ depends on the geometry
1994: of the particle array. For generic disordered arrays the network partitions the domain into
1995: polygons (Delaunay cells), most of which are triangles. We have showed that for these generic
1996: arrays $\lambda^\star=O(\delta^{-3/2})$. The same asymptotic rate is
1997: obtained for a larger class of networks called quasi-triangulated.
1998: In a quasi-triangulated network, the percentage of triangular cells may be relatively small, but
1999: the subnetwork containing triangular cells must be spanning. Another class of networks for which
2000: $\lambda^\star=O(\delta^{-3/2})$ consists of rectangular periodic arrays aligned with the boundary of the flow. More
2001: generally, the same rate is obtained for arrays containing a single spanning chain
2002: of neighboring particles, perpendicular to the part of the boundary where the velocity is prescribed.
2003:
2004:
2005:
2006: We show that in case of the strong blow up,
2007: the leading term in the
2008: asymptotics of $\lambda^\star$ can be uniquely determined by solving a simplified linear system of the
2009: network equations, provided the array is quasi-triangulated. The simplified system, obtained by neglecting rotational and pairwise shear motions of particles,
2010: provides an efficient computational tool for evaluating the dependence of the
2011: effective viscosity on the geometry of the particle array and external boundary
2012: conditions.
2013: When a network contains neither a spanning triangulated subgraph, nor a spanning vertical
2014: path, the extensional effective viscosity can be of order
2015: $\delta^{-1/2}$. This weak blow up
2016: is obtained, for instance, in the case of a periodic rectangular lattice, with interior edges misaligned with the orientation of the prescribed boundary data.
2017:
2018:
2019: Our results imply that the ratio of $\lambda^\star$
2020: to $\mu^\star$ for generic disordered particle arrays is $O(\delta^{-1})$. By contrast,
2021: in a Newtonian fluid this ratio depends only on the space dimension.
2022: Therefore, our results indicate possible non-Newtonian behaviour of the effective fluid.
2023: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2024: \section{Acknowledgments}
2025: The authors wish to thank Professor John F. Brady for bringing the work \cite{Brady} to their attention. Work of Leonid Berlyand was supported in part by NSF grant DMS-0204637. Work of Alexander Panchenko was supported in part by the ONR grant N00014-001-0853.
2026:
2027:
2028:
2029:
2030:
2031:
2032:
2033: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2034:
2035:
2036:
2037: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2038: \appendix
2039: \section{Shear and extensional flows. Ratio of the viscosities in a
2040: Newtonian fluid}
2041: \label{Appendix A}
2042: The following types of flows are relevant to our investigation.
2043:
2044: \noindent
2045: {\it Shear flow.} Consider the steady shear flow a homogeneous fluid
2046: characterized by the constant shear rate $\gamma$. The velocity is given by
2047: \begin{equation}
2048: \label{shear-v}
2049: {\uv}^0_{sh}=
2050: \left(
2051: \begin{array}{c}
2052: \gamma x_2\\
2053: 0
2054: \end{array}
2055: \right).
2056: \end{equation}
2057: and the strain rate tensor is
2058: \begin{equation}
2059: \label{sh-strain}
2060: {\ee}_{sh}^0=\frac{1}{2}
2061: \left(
2062: \begin{array}{cc}
2063: 0 & \gamma \\
2064: \gamma & 0
2065: \end{array}
2066: \right).
2067: \end{equation}
2068: Since the stress tensor is symmetric and independent of $\ux$,
2069: \begin{equation}
2070: \label{eo-shear}
2071: E^0=\int_\Omega \ts^0 \cdot {\ee}_{\it sh}^0d{\ux}=\frac 12 S^0_{12}\gamma |\Omega|,
2072: \end{equation}
2073: where $|\Omega|=\int_{\Omega} d\ux$.
2074:
2075: In the case of a homogeneous Newtonian fluid with viscosity $\mu$, $\ts^0=2\mu {\ee}_{sh}^0-PI$,
2076: so that $S^0_{12}=\mu\gamma$. Using the formula (\ref{shev1}) we obtain
2077: \begin{equation}
2078: \label{newtmu}
2079: \mu^\star=\mu,
2080: \end{equation}
2081: as expected.
2082:
2083:
2084:
2085: \noindent
2086: {\it Extensional flow.} In this case the fluid is being extended
2087: in the horizontal direction and simultaneously contracted in the vertical direction at
2088: the same constant rate $\vep$. The velocity is
2089: \begin{equation}
2090: \label{ef-v-ext}
2091: {\uv}^0_{\it ext}=
2092: \left(
2093: \begin{array}{c}
2094: \vep x_1\\
2095: -\vep x_2
2096: \end{array}
2097: \right),
2098: \end{equation}
2099: and
2100: \begin{equation}
2101: \label{ext-strain}
2102: {\ee}_{ext}^0=
2103: \left(
2104: \begin{array}{cc}
2105: \vep & 0 \\
2106: 0 & -\vep
2107: \end{array}
2108: \right).
2109: \end{equation}
2110: For a homogeneous Newtonian fluid, $\ts^0_{ext}=2\mu {\ee}^0_{ext}-PI$, and thus
2111: $S^0_{11}=2\mu\vep-P$, $S^0_{22}=-2\mu\vep-P$. Next using the definition (\ref{ev})
2112: we obtain
2113: \begin{equation}
2114: \label{newtev}
2115: \lambda^\star=4\mu.
2116: \end{equation}
2117: Therefore, for a Newtonian effective fluid the ratio $\lambda^\star/\mu^\star$ equals $4$
2118: (in two dimensions).
2119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2120: \section{Proof of Proposition \ref{quasi}}
2121: \label{Appendix B}
2122: The proposition \ref{quasi} will be proved if we prove the following
2123:
2124: \begin{proposition}
2125: \label{homsolve}
2126: Suppose the network graph $\Gamma$ is quasi-triangulated. Then every solution of the
2127: homogeneous
2128: system (\ref{comphom}) is of the form $t{\bf w}_0$ where $t$ is arbitrary
2129: real and ${\bf w}_0$
2130: is given by (\ref{w-not}).
2131: \end{proposition}
2132: \noindent
2133: {\it Proof.}
2134: First note that (\ref{comphom}) is Euler-Lagrange system for the functional
2135: \begin{equation}
2136: \label{q-hom}
2137: Q_{hom}=\frac 12 A{\bf z}\cdot{\bf z}=
2138: \displaystyle{\frac 12 \sum_{i=1}^N \sum_{j\in {\cal N}_i} C^{ij}_{sp}
2139: (({\bf T}^i-{\bf T}^j)\cdot{\bf q}^{ij})^2+\sum_{i\in I} C_1^{i}
2140: (\TI\cdot {\bf q}^{i})^2.}
2141: \end{equation}
2142: Clearly the minimum of $Q_{hom}$ is zero. Thus every solution of (\ref{comphom}) is a
2143: minimizer
2144: of $Q_{hom}$. On the other hand,
2145: $Q_{hom}({\bf T}^1,\ldots {\bf T}^N)=0$ if and only if
2146: the vectors $\TI$ satisfy the system of equations
2147: \begin{equation}
2148: \label{imp}
2149: \begin{array}{cc}
2150: ({\bf T}^i-{\bf T}^j)\cdot{\bf q}^{ij}=0,&~~~~i=1,2\ldots N,~~j\in {\cal N}_i\\
2151: \hspace{-11mm}\TI\cdot {\bf q}^{i}=0, & \hspace{-21mm}i\in I.
2152: \end{array}
2153: \end{equation}
2154: Therefore, a vector ${\bf z}=({\bf T}^1, \ldots {\bf T}^N)^T$ solves (\ref{comphom})
2155: if and only if $\TI, i=1,\ldots, N$ solve (\ref{imp}).
2156: The solvability of (\ref{imp}) will be directly linked to
2157: the geometric structure of the graph $\Gamma$.
2158: We begin by observing that ${\bf q}^i=\pm{\bf e}_2$. Thus the second set of equations
2159: in (\ref{imp}) yields
2160: \begin{equation}
2161: \label{e1}
2162: \TI=t^i {\bf e}_1, i\in I,
2163: \end{equation}
2164: that is, $\TI$ are horizontal
2165: for all boundary vertices. Next, consider boundary vertices ${\bf x}^i, i\in I^-$
2166: (these are vertices connected to $\bound^-$), and recall that they belong to a path
2167: $\Gamma^-$, edges of which are interior edges of $\Gamma$.
2168: Hence, if $i_1\in I^-$, then there is at least one
2169: $i_2\in I^-, i_2\ne i_1$ such that ${\ux}^{i_1}$ and ${\ux}^{i_2}$
2170: are connected by an {\it interior} edge $b^{i_1i_2}$.
2171: Using the first set of equations in (\ref{imp}) and (\ref{e1}) we obtain
2172: \begin{equation}
2173: \label{e2}
2174: ({\bf T}^{i_1}-{\bf T}^{i_2})\cdot
2175: {\bf q}^{i_1i_2}=(t^{i_1}-t^{i_2}){\bf e}_ 1\cdot{\bf q}^{i_1i_2}=0.
2176: \end{equation}
2177: Furthermore, $b^{i_1i_2}$ is non-vertical, that is,
2178: ${\bf q}^{i_1i_2}\cdot {\bf e}_1\ne 0$, which yields
2179: $t^{i_1}=t^{i_2}$. Since each ${\bf x}^i, i\in I^-$ is
2180: connected to at least one other, we obtain
2181: \begin{equation}
2182: \label{i-minus}
2183: \TI=t{\bf e}_1, ~~~~i\in I^-,
2184: \end{equation}
2185: with the same scalar $t$.
2186:
2187: Since the graph is quasi-triangulated, there
2188: exists an {\it interior}
2189: vertex ${\ux}^{l_1}\in \Gamma$ , ${\ux}^{l_1}\notin \Gamma^-$, connected
2190: to at least two
2191: vertices ${\ux}^{i_1}, {\ux}^{i_2},
2192: i_1, i_2 \in I^-$ by non-collinear edges. Then,
2193: using the first set of equations in (\ref{imp})
2194: we obtain
2195: \begin{equation}
2196: \label{e3}
2197: \begin{array}{c}
2198: ({\bf T}^{l_1}-t{\bf e}_1)\cdot{\bf q}^{l_1,i_1}=0,\\
2199: ({\bf T}^{l_1}-t{\bf e}_1)\cdot{\bf q}^{l_1,i_2}=0.
2200: \end{array}
2201: \end{equation}
2202: Since ${\bf q}^{l_1,i_1}$ and ${\bf q}^{l_1,i_2}$ are linearly independent,
2203: (\ref{e3}) yields
2204: ${\bf T}^{l_1}=t{\bf e}_1$. Next, let $G_1$ be the
2205: union of $\Gamma^-$,
2206: ${\ux}^{l_1}$ and all the edges which connect ${\ux}^{l_1}$ to $\Gamma^-$.
2207: Using the definition of the quasi-triangulated graph again, we find a vertex ${\ux}^{l_2}$, not
2208: contained in
2209: $G_1$, and connected to $G_1$ by two non-collinear edges.
2210: Repeating the argument following (\ref{e3}), we see that ${\bf T}^{l_2}=t{\bf e}_1$.
2211: Then we choose $G_2$ to be the union of $G_1$, ${\ux}^{l_2}$ and all
2212: edges of $\Gamma$ which connect
2213: them. Repeating the process we find the
2214: vertex ${\ux}_{l_3}$ and continue until we obtain
2215: \begin{equation}
2216: \label{e4}
2217: \TI =t{\bf e}_1, ~~~~~~~i=1, \dots N
2218: \end{equation}
2219: with the same scalar $t$. This means that every solution of (\ref{imp}) is of the form
2220: $t{\bf w}_0$. Since solution spaces of (\ref{imp}) and (\ref{comphom}) are the same,
2221: the proposition is proved.
2222:
2223:
2224: The following discrete Korn inequality follows immediately from the Proposition
2225: \ref{homsolve}.
2226:
2227: \begin{corollary}
2228: Suppose that $\Gamma$ is quasi-triangulated.
2229: Let $W\subset {\bf R}^{2N}$ be the one-dimensional subspace spanned by ${\bf w}_0$ from
2230: (\ref{w-not}), and let $W^\bot$ denote the orthogonal complement of $W$ in ${\bf R}^{2N}$.
2231: Also, let $Q_{hom}$ and $A$ be, respectively the quadratic form defined in (\ref{q-hom})
2232: and its matrix.
2233: Then there is a constant $C>0$ such that the Korn-type inequality
2234: \begin{equation}
2235: \label{d-korn}
2236: \frac 12 A{\bf z}\cdot{\bf z}=Q_{hom}({\bf z})\geq C {\bf z}\cdot{\bf z}
2237: \end{equation}
2238: holds for all ${\bf z}\in W^\bot$.
2239: \end{corollary}
2240: Another straightforward corollary is as follows.
2241: \begin{corollary}
2242: \label{u-solve}
2243: Suppose that $\Gamma$ is quasi-triangulated. Then the system
2244: $$
2245: A{\bf z}={\bf f}
2246: $$
2247: has a unique solution ${\bf z}\in W^\bot$ provided ${\bf f}\bot W$.
2248: \end{corollary}
2249:
2250: \noindent
2251: {\bf Remark}. The projection $P_W$ onto the subspace $W$ is defined by
2252: $$
2253: P_W {\bf z}=\frac{{\bf z}\cdot{\bf w}_0}{N}{\bf w}_0.
2254: $$
2255: In terms of vectors $\TI$,
2256: $$
2257: P_W( {\bf T}^1,{\bf T}^2,\ldots {\bf T}^{N})=\frac{\sum_{i=1}^n T^i_1}{N}{\bf w}_0
2258: $$
2259: Therefore, using the definition of ${\bf f}$ in terms of ${\bf R}^i$, we can write the
2260: condition ${\bf f}\bot W$ as
2261: \begin{equation}
2262: \label{orthog}
2263: \sum_{i=1}^n R^i_1=\sum_{i=1}^N{\bf R}^i\cdot {\bf e}_1=0
2264: \end{equation}
2265:
2266: The vectors ${\bf R}^i$ in (\ref{lead-rhs}) satisfy (\ref{orthog}), so that ${\bf f}$ in
2267: (\ref{zf}) with $R^i$ defined by (\ref{lead-rhs}) is orthogonal to $W$.
2268: This gives the unique solvability of the network
2269: equations (\ref{compls}).
2270:
2271:
2272: \begin{corollary}
2273: \label{al-solve}
2274: Suppose that $\Gamma$ is quasi-triangulated.
2275: Then there is a unique ${\bf z}^*\in W^\bot$ such that every solution of (\ref{compls})
2276: is of the form ${\bf z}^*+t{\bf w}_0$, where $t{\bf w}_0\in W$.
2277: \end{corollary}
2278: This means that solution of the network equations (\ref{compls}) is unique up to a
2279: horizontal translation.
2280:
2281: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2282: \section{Proof of proposition \ref{tricon}}
2283: \label{Appendix C}
2284: \noindent
2285: {\it Proof.}
2286: First we observe that the form $\widehat Q$ is a sum of non-negative terms, each of
2287: which corresponds to an edge of the network graph $\Gamma$. Removal of an edge from
2288: $\Gamma$ corresponds to deletion of one non-negative term in $\widehat Q$.
2289: This means that for each subgraph $\Gamma^\prime$ of $\Gamma$,
2290: $\widehat Q(\Gamma)\geq \widehat Q(\Gamma^\prime)$. Next we choose $\Gamma^\prime$ to be
2291: the maximal quasi-triangulated subgraph of $\Gamma$.
2292: We show that $\min \widehat Q(\Gamma^\prime)>0$. Indeed, $\min \widehat Q(\Gamma^\prime)=0$
2293: if and only if
2294: the corresponding system (\ref{ns}) has a solution.
2295: To show that this system
2296: has no solutions, introduce
2297: new unknowns $\hat{\bf T}^i={\bf T}^i-{\bf e}_2$. Then from
2298: (\ref{ns}) we obtain
2299: \begin{equation}
2300: \label{ns1}
2301: (\hat{\bf T}^i-\hat{\bf T}^j)\cdot {\bf q}^{ij}=0,
2302: \end{equation}
2303: \begin{equation}
2304: \label{ns2}
2305: \hat{\bf T}^i\cdot {\bf q}^i=
2306: \begin{cases}
2307: -2\;\;\;\;{\rm when}\;i\in I^+&\\
2308: 0\;\;\;\;{\rm when}\;i\in I^-.&\\
2309: \end{cases}
2310: \end{equation}
2311: Since the vectors ${\bf q}^i$ are vertical, (\ref{ns2}) yields
2312: \begin{equation}
2313: \label{bd}
2314: \hat{\bf T}^i=t^i {\bf e}_1,\;\;\;\;i \in I^-,
2315: \end{equation}
2316: where $t^i$ is a constant.
2317: Recall that $\Gamma$ contains a path $\Gamma^-$ which consists of all
2318: boundary vertices connected to $\bound^-$ and all interior edges
2319: connecting these vertices. Hence,
2320: each ${\bf x}^{i_1}, i_1\in I^-$ has a neighbor ${\bf x}^{i_2},
2321: i_2\in I^-$ and thus
2322: $(t^{i_1}-t^{i_2}){\bf e}_1\cdot{\bf q}^{i_1i_2}=0$ from (\ref{ns1}).
2323: Since two boundary vertices cannot be joined by a vertical edge,
2324: ${\bf e}_1\cdot{\bf q}^{i_1i_2}\ne 0$. This implies that all
2325: $t^i, i\in I^-$ are equal, that is
2326: \begin{equation}
2327: \label{t}
2328: \hat{\bf T}^i=t{\bf e}_1, i\in
2329: I^-,
2330: \end{equation}
2331: where $t$ is a constant.
2332: Next, consider the boundary path $\Gamma^-$. By definition of $\Gamma^\prime$
2333: there is a vertex ${\bf x}^{j_1}, j_1 \not\in I^-$ connected
2334: to the boundary vertices ${\bf x}^{i_1}, {\bf x}^{i_2}$,
2335: $ i_1, i_2 \in I^-$ by non-collinear edges of $\Gamma$.
2336: Then from (\ref{ns1}) and (\ref{t}) we have
2337: $$
2338: \begin{array}{c}
2339: (\hat{\bf T}^{j_1}-t{\bf e}_1)\cdot {\bf q}^{j_1,i_1}=0,\\
2340: (\hat{\bf T}^{j_1}-t{\bf e}_1)\cdot {\bf q}^{j_1,i_2}=0.\\
2341: \end{array}
2342: $$
2343: Since ${\bf q}^{j_1,i_1}$ and ${\bf q}^{j_1,i_2}$ are linearly
2344: independent, we obtain $\hat{\bf T}^{j_1}=t{\bf e}_1$. Now this
2345: argument can be used recursively. Next we choose $G_1$ to
2346: be the union of vertices ${\bf x}^i, i\in I^-, {\bf x}^{j_1}$ and
2347: the edges of $\Gamma^\prime$ which connect these vertices. Repeating the
2348: argument, we find a vertex ${\bf x}^{j_2}\not\in G_1$,
2349: connected to at least two vertices of $G_1$ by non-collinear
2350: edges, which yields $\hat{\bf T}^{j_2}=t{\bf e}_1$, and so on, until
2351: we obtain $\hat{\bf T}^i=t{\bf e}_1$ for all vertices ${\bf x}^i$ which belong to
2352: $\Gamma^\prime$. By assumption, $\Gamma^\prime$ contains at least one vertex ${\bf x}^+ \in I^+$.
2353: But then $\hat{\bf T}^+ \cdot {\bf q}^+=-t{\bf e}_1\cdot {\bf e}_2=0$ which contradicts
2354: (\ref{ns2}). This contradiction shows that the system (\ref{ns1}), (\ref{ns2})
2355: has no solutions.
2356:
2357:
2358:
2359: %%%%%%%%%%%%%% R E F E R E N C E S %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2360: \begin{thebibliography}{}
2361: \bibitem{Batch} \textsc{Batchelor, G. K. \& Green, J. T.} 1972
2362: The determination
2363: of the bulk stress in a suspension of spherical particles to order
2364: $c^2$. {\em J. Fluid Mech.} {\bf 56} Part 3,401--427.
2365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2366: \bibitem{Bach} \textsc{Bakhvalov, N. \& Panasenko, G.} 1989
2367: {\em Homogenization: averaging processes in periodic media}. Kluwer.
2368: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2369: \bibitem{BLP} \textsc{Bensoussan, A., Lions, J. L., \& Papanicolaou, G.} 1978 {\em Asymptotic analysis in periodic structures}.
2370: North-Holland.
2371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2372: \bibitem{BBP}
2373: \textsc{Berlyand, L., Borcea, L \& Panchenko, A.} 2003 Network approximation
2374: for effective viscosity of concentrated suspensions with complex geometries.
2375: {\em SIAM Journ. Math. Anal.}, to appear.
2376: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2377: \bibitem{BK}
2378: \textsc{Berlyand, L. \& Kolpakov, A.} 2001 Network approximation
2379: in the limit of small inter-particle distance of the effective
2380: properties of a high-contrast random dispersed composite.
2381: {\em Arch. Rat. Mech. Anal.} {\bf 159}, 179--227.
2382: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2383: \bibitem{Carr}
2384: \textsc{Carreau, P. J. \& Cotton, F.} 2002 Rheological properties
2385: of concentrated suspensions. In: {\em Transport processes in bubbles, drops and particles}.
2386: D. De Kee and R. P. Chhabra, eds., Taylor \& Fransis.
2387: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2388: \bibitem{Cous}\textsc{Coussot, P.} 2002 Flows of concentrated
2389: granular mixtures. In: {\em Transport processes in bubbles, drops and particles}.
2390: D. De Kee adn R. P. Chhabra, eds., Taylor \& Fransis.
2391: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2392: \bibitem{Einst}
2393: \textsc{Einstein, A.} 1906 Eine neue Bestimmung der Molek\"{u}ldimensionen,
2394: {\em Annln. Phys.} {\bf 19} p. 289 and {\bf 34} p. 591.
2395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2396: \bibitem{FA} \textsc{
2397: Frankel, N. A. \& Akrivos, A.} 1967 On the Viscosity of
2398: a Concentrated Suspension of Solid Spheres. {\em Chemical Engineering Science}
2399: {\bf 22}, 847--853.
2400: %%%%%%%%%%%%%%%%
2401: \bibitem{JKO}
2402: \textsc{Jikov, V., Kozlov, S. \& Oleinik, O.}
2403: 1994 {\em Homogenization of differential operators and integral functionals}.
2404: Springer.
2405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2406: %\bb{Noll} W. Noll, Arch. Rat. Mech. Anal., 1958, {\bf 2}, 197-226.
2407: %%%%%%%%%%%%%%%%%%%%%%
2408: \bibitem{Keller} \textsc{Nunan. K. C. \& Keller, J. B.} 1984
2409: Effective viscosity of periodic suspensions
2410: {\em Journ. Fluid Mech.} {\bf 142}, 269--287.
2411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2412: \bibitem{SP} \textsc{Sanchez-Palencia, E.} 1980
2413: {\em Non-homogeneous media and vibration theory}. Springer.
2414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2415: \bibitem{Sh} \textsc{Schowalter, W. R.} 1978
2416: {\em Mechanics of Non-Newtonian Fluids}. Pergamon Press.
2417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2418: \bibitem{Brady} \textsc{Sierou, A. \& Brady, J. F.} 2001
2419: Accelerated Stokesian
2420: Dynamic simulations. {\em J. Fluid Mech.} {\bf 448}, 115--146.
2421: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2422: \bibitem{ShP} \textsc{Shikata, T \& Pearson, D. S.} 1994
2423: Viscoelastic behavior of
2424: concentrated spherical suspensions. {\em Journ. Rheology} {\bf 38}, 601--616.
2425: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2426: \bibitem{Shook}
2427: \textsc{Shook, C. A. \& Rocko, M. C.} 1991 {\em Slurry flow. Principles and practice.}
2428: Butterworth-Heinemann.
2429: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2430: \bibitem{VW} \textsc{Van der Werff, J. C., de Kruif, C. G.,
2431: Blom, C, \& Mellema, J.} 1989
2432: Linear viscoelastic behavior of dense hard-sphere dispersions.
2433: {\em Phys. Rev. A}. {\bf 39}, 795--807.
2434: \end{thebibliography}
2435:
2436:
2437:
2438:
2439: \end{document}
2440:
2441:
2442: