1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: \documentclass[twocolumn,showpacs,amsmath,amssymb,bibnotes]{revtex4}
3: %\documentclass[preprint,showpacs,amsmath,amssymb,bibnotes]{revtex4}
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \pdfoutput=1
6: \usepackage{bm}
7: \usepackage{graphicx,epsfig,subfigure,afterpage,bm}
8: \usepackage{amsfonts}
9: %\usepackage{mcite}
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11:
12: \newcommand{\be}{\begin{equation}}
13: \newcommand{\ee}{\end{equation}}
14: \newcommand{\bea}{\begin{eqnarray}}
15: \newcommand{\eea}{\end{eqnarray}}
16: \newcommand{\gdot}{\dot{\gamma}}
17: \newcommand{\gdotbar}{\overline{\dot{\gamma}}}
18: \newcommand{\ie}{{\it i.e.\/}}
19: \newcommand{\eg}{{\it e.g.\/}}
20: \newcommand{\etc}{{\it etc.\/}}
21: \newcommand{\versus}{{\it vs.\/}}
22: \newcommand{\etal}{{\it et al.\/}}
23: \newcommand{\bw}{\begin{widetext}}
24: \newcommand{\ew}{\end{widetext}}
25: \newcommand{\lae}{\stackrel{<}{\scriptstyle\sim}}
26: \newcommand{\gae}{\stackrel{>}{\scriptstyle\sim}}
27:
28: \newcommand{\tmax}{t_{\rm max}}
29: \newcommand{\xhat}{\vecv{\hat{x}}}
30: \newcommand{\yhat}{\vecv{\hat{y}}}
31: \newcommand{\zhat}{\vecv{\hat{z}}}
32:
33: %\newcommand{\vecv}[1]{\mathbf{{#1}}}
34: \newcommand{\vecv}[1]{\bm{{#1}}}
35: %\newcommand{\tens}[1]{\mathbf{{#1}}}
36: \newcommand{\tens}[1]{\bm{{#1}}}
37: \newcommand{\nablu}{{\bf \nabla}}
38:
39: \newcommand{\vt}{\tilde{\vecv{v}}}
40: \newcommand{\st}{\tilde{\psi}}
41: \newcommand{\wt}{\tilde{\omega}}
42: \newcommand{\pt}{\tilde{\phi}}
43:
44: \newcommand{\ldv}{L_{\rm DV}}
45: \newcommand{\lvi}{L_{\rm VI}}
46:
47: \newcommand{\lke}{L_{\rm KE}}
48: \newcommand{\lpe}{L_{\rm PE}}
49: \newcommand{\lsm}{L_{\rm Sm}}
50: \newcommand{\lx}{L_{\rm x}}
51: \newcommand{\ly}{L_{\rm y}}
52: \newcommand{\lint}{L_{\rm INT}}
53:
54: \newcommand{\Dx}{\delta_x}
55: \newcommand{\Dy}{\delta_y}
56: \newcommand{\Dt}{\delta_t}
57:
58:
59: \newcommand{\tdv}{T_{\rm DV}}
60: \newcommand{\tvi}{T_{\rm VI}}
61:
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: \begin{document}
64:
65: \title{The need for inertia in nonequilibrium steady states of sheared binary fluids}
66: \author{Suzanne M. Fielding}
67: \email{suzanne.fielding@manchester.ac.uk}
68: \affiliation{School of Mathematics and Manchester Centre for Nonlinear Dynamics, University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom }
69:
70: \date{\today}
71: \begin{abstract}
72: We study numerically phase separation in a binary fluid subject to
73: an applied shear flow in two dimensions, with full hydrodynamics.
74: To do so, we introduce a mixed finite-differencing/spectral
75: simulation technique, with a transformation to render trivial the
76: implementation of Lees-Edwards sheared periodic boundary conditions.
77: For systems with inertia, we reproduce the nonequilibrium steady
78: states reported in a recent lattice Boltzmann study. The domain
79: coarsening that would occur in zero shear is arrested by the applied
80: shear flow, which restores a finite domain size set by the inverse
81: shear rate. For inertialess systems, in contrast, we find no
82: evidence of nonequilibrium steady states free of finite size
83: effects: coarsening persists indefinitely until the typical domain
84: size attains the system size, as in zero shear. We present an
85: analytical argument that supports this observation, and that
86: furthermore provides a possible explanation for a hitherto puzzling
87: property of the nonequilibrium steady states with inertia.
88: \end{abstract}
89: \pacs{ {64.75.+g},
90: {47.11.Qr}.
91: }
92: \maketitle
93:
94: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
95:
96: \section{Introduction}
97: \label{sec:intro}
98:
99: When an initially homogeneous mixture of two incompressible fluids (A
100: and B) undergoes a deep temperature quench into the spinodal regime,
101: it becomes unstable with respect to spatial fluctuations in the
102: composition field $\phi(\vecv{x},t)$. The mixture then phase separates
103: into well defined domains of A-rich and B-rich fluid, which, after a
104: rapid initial transient, attain local equilibrium within each domain.
105: The system remains globally out of equilibrium on long timescales,
106: however, due to the excess energy that resides in the interfaces
107: between the domains. Here we consider the maximally symmetric case of
108: a 50:50 mixture of two mutually phobic fluids with matched viscosities
109: and densities.
110:
111: Following the initial transient of their formation, the domains slowly
112: coarsen in time through the action of the surface tension in the
113: interfaces that separate them. (For reviews of phase ordering
114: kinetics, see
115: Refs.~\cite{hohenberg77,bray-aip-43-357-1994,cates-fd--1-1999}.) In
116: this way, the excess interfacial energy of the system progressively
117: relaxes towards its minimal equilibrium value. This coarsening process
118: proceeds through three distinct regimes that are successively
119: dominated by diffusive, viscous and inertial dynamics. In the limit
120: of an infinite system size $\Lambda\to\infty$ (taken first), the
121: typical domain size perpetually increases without bound: the system
122: never globally equilibrates, even in the limit of infinite time
123: $t\to\infty$ (taken second). For any finite system size, coarsening
124: is in practice eventually cutoff once the typical domain size $L(t)$
125: attains the system size $\Lambda$.
126:
127: Besides these systems that remain out of equilibrium as they slowly
128: relax towards a fully phase separated state, another class of systems
129: comprises those that are continuously driven out of equilibrium by the
130: external application of a steady shear flow. In this paper, we
131: consider systems that are {\em both} undergoing phase separation {\em
132: and} simultaneously subject to an applied shear flow. They thus
133: combine both of the nonequilbrium features just described. The main
134: question that we address is whether shear interrupts domain coarsening
135: and restores a nonequilibrium steady state with a typical domain size
136: $L(\gdot^{-1})$ set by the inverse of the applied shear rate $\gdot$;
137: or whether coarsening persists indefinitely, up to the system size, as
138: in zero shear.
139:
140: Despite previous
141: experimental~\cite{hashimoto-jcp-88-5874-1988,takahashi-jr-38-699-1994,chan-pra-43-1826-1991,krall-prl-69-1963-1992,hashimoto-prl-74-126-1995,beysens-pra-28-2491-1983,chan-prl-61-412-1988,matsuzaka-prl-80-5441-1998,hobbie-pre-54-R5909-1996,qiu-pre-58-R1230-1998,lauger-prl-75-3576-1995},
142: numerical~\cite{shou-pre-61-R2200-2000,zhang-jcp-113-8348-2000,yamamoto-pre-59-3223-1999,corberi-prl-81-3852-1998,corberi-prl-83-4057-1999,corberi-pre-61-6621-2000,corberi-pre-62-8064-2000,rothman-prl-65-3305-1990,rothman-el-14-337-1991,chan-el-11-13-1990,qiu-jcp-108-9529-1998,padilla-jcp-106-2342-1997,ohta-jcp-93-2664-1990,wagner-pre-59-4366-1999,lamura-pamia-294-295-2001,lamura-epjb-24-251-2001,berthier-pre-6305--2001}
143: and
144: theoretical~\cite{onuki-jpm-9-6119-1997,lamura-pamia-294-295-2001,corberi-prl-81-3852-1998,corberi-prl-83-4057-1999,corberi-pre-61-6621-2000,corberi-pre-62-8064-2000,bray-ptrslsapes-361-781-2003,cavagna-pre-62-4702-2000,bray-jpag-33-L305-2000,Onuk97,doi-jcp-95-1242-1991}
145: work, this question remained open until the recent simulation studies
146: of Stansell et al. in Refs~\cite{stansell-prl-96--2006,cates-3d-2007}.
147: Using Lattice Boltzmann techniques, they gave convincing evidence for
148: the formation of nonequilibrium steady states, with domains of a
149: finite size set by the inverse shear rate, independent of the system
150: size. The domain morphology inherits the anisotropy of the applied
151: shear flow, and is therefore characterised by two lengthscales
152: $L_{\|}$ and $L_\bot$, respectively describing the major and minor
153: domain axes. A remarkable achievement of
154: Refs.~\cite{stansell-prl-96--2006,cates-3d-2007} was the exploration,
155: by judicious parameter steering, of a range of inverse shear rates
156: spanning six decades on a scaling plot. For reasons discussed below
157: this range would, a priori, be expected to encompass two different
158: regimes, separately dominated by viscous and inertial dynamics.
159: Surprising, therefore, is the observation of an apparently single
160: scaling with shear rate for each of $L_{\|}$ and $L_\bot$, across all
161: six decades.
162:
163: \begin{figure*}[tbp]
164: \includegraphics[scale=0.35]{./domain.png}
165: \caption{Flow geometries in the (a) unsheared and (b) sheared cases.
166: The crossed connected by thin dotted lines show representative
167: pairs of points connected by periodic boundary conditions.}
168: \label{fig:domain}
169: \end{figure*}
170:
171: All the simulations reported in
172: Refs.~\cite{stansell-prl-96--2006,cates-3d-2007} have non zero fluid
173: inertia. Indeed, even when attempting to access the limit of zero
174: inertia, a small but non-zero inertia remains a practical requirement
175: of the Lattice Boltzmann technique. In this paper, we investigate
176: phase separation under shear in systems that are strictly inertialess.
177: To do so, we apply a different simulation technique to this problem,
178: comprising finite differencing combined with Fourier spectral methods.
179: We also use a convenient transformation to render trivial the
180: implementation of sheared periodic boundary
181: conditions~\cite{onuki-jpsj-66-1836-1997}.
182:
183: Our main numerical result will be that, in truly inertialess systems,
184: coarsening persists indefinitely, up to the system size, despite the
185: external shear flow. We will also present an analytical argument that
186: supports this observation. A simple extension to the same argument
187: provides a possible explanation for the existence of a single scaling
188: for each of $L_{\|}, L_\bot$ across all six decades of scaled shear
189: rate in the simulations of Ref.~\cite{stansell-prl-96--2006}, with
190: inertia.
191:
192: We start by introducing the model equations, flow geometry and choice
193: of units in Secs.~\ref{sec:equations},~\ref{sec:geometry}
194: and~\ref{sec:units} respectively. In Sec.~\ref{sec:numerics} we
195: briefly outline our numerical method, of which further details are
196: given in the Appendices. The length and timescales that characterise
197: demixing are presented in Sec.~\ref{sec:lengthscales}, leading to a
198: discussion in Sec.~\ref{sec:parameters} of the choice of parameter
199: values in our simulations. In Sec.~\ref{sec:results} we present our
200: numerical results, starting in Sec.~\ref{sec:zeroShear} with the case
201: of coarsening in zero shear, before considering an applied shear flow
202: with and without inertia in Secs.~\ref{sec:inertia}
203: and~\ref{sec:noInertia} respectively. Sec.~\ref{sec:link} contains a
204: linking discussion. The results in Sec.~\ref{sec:noInertia} are (we
205: believe) novel. In contrast, our aim in Secs.~\ref{sec:zeroShear}
206: and~\ref{sec:inertia} is simply to reproduce the behaviour seen in
207: previous lattice Boltzmann
208: studies~\cite{wagner-prl-80-1429-1998,kendon-jfm-440-147-2001,stansell-prl-96--2006},
209: thereby gaining confidence in our own simulation method. In
210: Sec.~\ref{sec:argument} we present an analytical argument that
211: supports our numerical observations, as well as those of
212: Ref.~\cite{stansell-prl-96--2006}. Sec.~\ref{sec:conclusions} contains
213: a summary and an outlook to future work.
214:
215: \section{Governing equations}
216: \label{sec:equations}
217:
218: The fluid velocity field $\vecv{v}(\vecv{x},t)$ and pressure
219: $p(\vecv{x},t)$ are governed by the Navier-Stokes equation
220: %
221: \be
222: \label{eqn:ns}
223: \rho\left(\partial_t\, \vecv{v} + \vecv{v}\cdot \nablu \, \vecv{v} \right)=\eta \nabla^2 \vecv{v} -\phi \nablu \mu - \nablu p
224: \ee
225: %
226: together with the incompressibility constraint
227: %
228: \be
229: \label{eqn:incomp}
230: \nablu\cdot \vecv{v}=0.
231: \ee
232: %
233: Here $\rho$ is the fluid density and $\eta$ the viscosity.
234: $p(\vecv{x},t)$ is the thermodynamic part of the pressure tensor and
235: $\mu(\vecv{x},t)$ is a chemical potential, defined below. In what
236: follows, we shall consider two dimensional (2D) flow in the $x-y$
237: plane with velocity components $v_x,v_y$.
238:
239: To eliminate the pressure, we take the curl of Eqn.~\ref{eqn:ns}. The
240: $z$ component of the resulting equation is
241: %
242: \be
243: \label{eqn:vorticity}
244: \rho\left(\partial_t\,\omega + \vecv{v}\cdot\nablu \, \omega\right)=\eta
245: \,\nabla^2 \omega - \left[\nablu \wedge \phi \nablu \mu\right]\cdot \zhat
246: \ee
247: %
248: in which the vorticity $\omega$ obeys
249: %
250: \be
251: \nabla\wedge \vecv{v}=\omega \zhat.
252: \ee
253: %
254: To ensure that the incompressibility constraint~(\ref{eqn:incomp}) is
255: automatically obeyed, we define a streamfunction $\psi$ via
256: %
257: \be
258: \label{eqn:stream}
259: \vecv{v}=\nablu\wedge \psi \zhat,
260: \ee
261: %
262: which is related to the vorticity as follows:
263: %
264: \be
265: \label{eqn:streamVort}
266: \omega=-\nabla^2\psi.
267: \ee
268: %
269:
270: The dynamics of the compositional order parameter $\phi(\vecv{x},t)$
271: are prescribed by an advection-diffusion equation of Cahn-Hilliard
272: type (see Refs.~\cite{bray-aip-43-357-1994,swift-pre-54-5041-1996})
273: %
274: \be
275: \label{eqn:phi1}
276: \partial_t \phi + \vecv{v}\cdot \nablu \phi = M\, \nabla^2 \mu.
277: \ee
278: %
279: Here $M$ is the mobility, assumed constant, controlling the rate of
280: intermolecular diffusion. The chemical potential
281: %
282: \be
283: \label{eqn:mu1}
284: \mu= G\,\phi(\phi^2-1)-\kappa \nabla^2\phi,
285: \ee
286: %
287: in which $G$ is a positive constant with the dimensions of stress.
288: $\kappa$ is also a positive constant, controlling the characteristic
289: width $l$ of the interface between domains:
290: %
291: \be
292: l=\left(\frac{\kappa}{G}\right)^\frac{1}{2}.
293: \ee
294: %
295: The surface tension of the interface is given by
296: %
297: \be
298: \sigma=\frac{2\sqrt{2}lG}{3}.
299: \ee
300: %
301:
302: \section{Flow geometry}
303: \label{sec:geometry}
304:
305: In the absence of shear we consider a square domain of size
306: $\Lambda_x=\Lambda_y$ in the $x-y$ plane (Fig.~\ref{fig:domain}a). We
307: adopt periodic boundary conditions in each direction:
308: %
309: \bea
310: \phi(x=0,y)&=&\phi(x=\Lambda_x,y)\;\;\;\forall\;\;\;y,\\
311: \phi(x,y=0)&=&\phi(x,y=\Lambda_y)\;\;\;\forall\;\;\;x,
312: %v_\alpha(x=0,y)&=&v_\alpha(x=\Lambda_x,y) \;\;\;\forall\;\;\;y,
313: \eea
314: %
315: and similarly for $v_x,v_y,\psi$ and $\omega$.
316:
317: To study shear, we consider a system that is rectangular with
318: $\Lambda_x=2 \Lambda_y$. Conceptually, a shear flow of rate $\gdot$ is
319: applied by moving the upper boundary to the right at a constant speed
320: $\gdot \Lambda_y$, as shown in Fig.~\ref{fig:domain}b (left), so that
321: the velocity field then has a constant affine part $\gdot y\xhat$ and
322: an additive fluctuating contribution $\vt(\vecv{x},t)$. In practice,
323: this is implemented via Lees-Edwards sliding periodic boundary
324: conditions (Fig.~\ref{fig:domain}b, right), such that the system has
325: no physical borders but instead comprises an infinitely repeating
326: array of images arranged in horizontally sliding layers. The boundary
327: conditions are then
328: %
329: \bea
330: \phi(x=0,y)&=&\phi(x=\Lambda_x,y)\;\;\;\forall\;\;\;y,\\
331: \phi(x,y=0)&=&\phi(x+\gdot
332: \Lambda_y t,y=\Lambda_y)\;\;\;\forall\;\;\;x. \eea
333: %
334: These also apply to $\tilde{v}_x,\tilde{v}_y$, as well as the
335: corresponding fluctuating contribution to the streamfunction
336: $\tilde{\psi}$ and vorticity $\tilde{\omega}$.
337:
338: \section{Choice of units}
339: \label{sec:units}
340:
341: The model equations contain five parameters: $\rho, \eta, M, G$ and
342: $l$. (We must specify two of $G,l,\kappa$ and $\sigma$, and here have
343: chosen $G$ and $l$.) In addition, we must also specify the system's
344: dimensions $\Lambda_x,\Lambda_y$ and the applied shear rate $\gdot$,
345: making eight parameters in all. Three of these can be eliminated by
346: choosing appropriate units for length, time and stress. Accordingly,
347: we measure lengths in units of the vertical system size
348: $\Lambda_y\equiv 1$; stresses in units of $G\equiv 1$; and times in
349: units of the characteristic ``microscopic'' time for diffusion across
350: an interface,
351: %
352: \be
353: \tau_0=\frac{(\sqrt{2}l)^3}{3M\sigma}\equiv 1.
354: \ee
355: %
356: In these units, the governing equations comprise
357: Eqns.~\ref{eqn:vorticity},~\ref{eqn:stream} and~\ref{eqn:streamVort}
358: (unchanged) together with
359: %
360: \be
361: \label{eqn:phi}
362: \partial_t\, \phi + \vecv{v}\cdot \nablu \phi = l^2 \, \nabla^2 \mu
363: \ee
364: %
365: and
366: %
367: \be
368: \label{eqn:mu}
369: \mu= \phi(\phi^2-1)- l^2 \nabla^2\phi.
370: \ee
371: %
372: This leaves the rescaled density $\rho$, viscosity $\eta$, interfacial
373: width $l$, aspect ratio $\Lambda_x$ and applied shear rate $\gdot$ as
374: the five control parameters. In our numerical work we fix $\Lambda_x$
375: and $l$, varying $\rho,\eta$ and $\gdot$ between runs.
376:
377:
378: \section{Numerical method}
379: \label{sec:numerics}
380:
381: In the laboratory frame, the numerical implementation of sheared
382: periodic boundary conditions is rather involved. We now discuss a
383: convenient transformation that renders it
384: trivial~\cite{onuki-jpsj-66-1836-1997}. We do so here in outline
385: only, referring the interested reader to appendix~\ref{sec:appendix1}
386: for details.
387:
388: The transformation comprises two steps. In the first, the velocity
389: field is expressed as the sum of a constant affine part and a
390: fluctuating contribution:
391: %
392: \be
393: \vecv{v}(\vecv{x},t)=\gdot y\xhat + \vt(\vecv{x},t).
394: \ee
395: %
396: In the second step, we transform to the cosheared frame
397: %
398: \be
399: (x,y,t)\to(x'=x-\gdot ty, y'=y,t'=t),
400: \ee
401: %
402: defining for convenience the cosheared gradient operator
403: %
404: \be
405: \label{eqn:coshearedGradient}
406: \nablu_c=\xhat\, \partial_{x'}+\yhat\,(-\gdot t\partial_{x'}+\partial_{y'}).
407: \ee
408: %
409: As shown in Appendix~\ref{sec:appendix1}, the final transformed
410: equation set, dropping the dashes for clarity, is then:
411: %
412: \be
413: \label{eqn:equation1}
414: {\omega}=-\nabla_c^2\,{\psi},
415: \ee
416: %
417: together with
418: %
419: \be
420: \label{eqn:equation2}
421: \rho\,\partial_{t}{\omega}+\rho\left(\partial_{y}{\psi}\partial_{x}{\omega}-\partial_{x}{\psi}\partial_{y}{\omega}\right)=\eta\,\nabla_c^2{\omega}-\left[\partial_{x}\phi\,\partial_{y} \mu-\partial_{y}\phi \,\partial_{x} \mu\right],
422: \ee
423: %
424: \be
425: \label{eqn:equation3}
426: \partial_{t}\phi+\left(\partial_{y}{\psi}\partial_{x}\phi-\partial_{x}{\psi}\partial_{y}\phi\right)=l^2\nabla_c^2\mu,
427: \ee
428: %
429: %
430: \be
431: \label{eqn:equation4}
432: \mu=\phi(\phi^2-1)-l^2\nabla_c^2\phi.
433: \ee
434: %
435: %
436: A priori, the bracketed expressions in Eqns.~\ref{eqn:equation2}
437: and~\ref{eqn:equation3} contain terms in the applied shear rate
438: $\gdot$. However in each bracket these are equal and opposite, and so
439: cancel. An equivalent statement is that the bracketed expressions are
440: invariant under shear.
441:
442: This transformation considerably simplifies the implementation of an
443: applied shear flow. Indeed, looking at Eqns.~\ref{eqn:equation1}
444: to~\ref{eqn:equation4} we see that the only effect of shear is to
445: renormalise the gradient operator $\nabla\to\nabla_c$. Furthermore,
446: the dynamical variables $\phi$, $\psi$ and $\omega$ are now subject to
447: {\em ordinary} periodic boundary conditions:
448: %
449: \bea
450: \phi(x=0,y)&=&\phi(x=\Lambda_x,y)\;\;\;\forall\;\;\;y,\nonumber\\
451: \phi(x,y=0)&=&\phi(x,y=\Lambda_y)\;\;\;\forall\;\;\;x,
452: \label{eqn:opbcs}
453: \eea
454: %
455: and similarly for $\psi$ and $\omega$. This will allow us to use
456: Fourier transforms in our numerical algorithm, as described below.
457:
458:
459: Under the transformation described so far, the relative shear of the
460: laboratory and cosheared frames becomes large at long times, diverging
461: at a constant rate $\gdot$. This is clearly expected to give rise to
462: numerical instabilities. To circumvent this problem, once the relative
463: shear $s(t)$ attains $\Lambda_x/2 \Lambda_y$ we perform by hand an
464: instantaneous shift $s(t)\to s(t)-\Lambda_x/\Lambda_y$. In this way,
465: the function $s(t)$ is bounded between $s=-\Lambda_x/2 \Lambda_y$ and
466: $s=\Lambda_x/2 \Lambda_y$, comprising a sawtooth with sections of
467: constant slope $\gdot$ connected by negative step discontinuities of
468: height $\Lambda_x/\Lambda_y$ at equal time intervals $\Delta
469: t=\Lambda_x/\gdot\Lambda_y$. Because this shift involves moving the
470: upper wall by exactly one multiple of the system length, the periodic
471: boundary conditions (\ref{eqn:opbcs}) are unaffected. For times
472: between these shifts, the sole effect of this modification appears in
473: the cosheared gradient operator, which, in place of
474: (\ref{eqn:coshearedGradient}), is now
475: %
476: \be
477: \label{eqn:coshearedNabluFinal}
478: \nablu_c=\xhat\, \partial_{x}+\yhat\,(-s(t)\partial_{x}+\partial_{y}).
479: \ee
480: %
481: The special case of zero shear is trivially achieved by setting
482: $s(t)=0\;\forall\;t$.
483:
484: In appendix~\ref{sec:appendix2}, we discuss the numerical algorithm
485: used to study the dynamical evolution of
486: $\phi(\vecv{x},t),\omega(\vecv{x},t)$ and $\psi(\vecv{x},t)$, as
487: specified by Eqns.~\ref{eqn:equation1} to~\ref{eqn:equation4}
488: and~\ref{eqn:coshearedNabluFinal}. Readers who are not interested in
489: these issues can skip straight to Sec.~\ref{sec:lengthscales} without
490: loss of thread.
491:
492: \section{Lengthscales and timescales}
493: \label{sec:lengthscales}
494:
495: Following a deep temperature quench of an unsheared system into the
496: two phase regime, well defined domains of each phase form, separated
497: by sharp interfaces. See Fig.~\ref{fig:state_u}, for example, in which
498: the white (resp. black) patches are domains of A (resp. B)-rich
499: fluid. After a non-universal transient associated with their initial
500: formation, these domains progressively grow in size through the action
501: of surface tension, passing through regimes that are successively
502: dominated by diffusive, viscous and inertial dynamics. During this
503: process of domain coarsening, the dynamical scaling
504: hypothesis~\cite{siggia,furukawa-aip-34-703-1985,bray-aip-43-357-1994} states that any structural
505: lengthscale $L(t)$ characterising the typical domain size ({\it e.g.},
506: as measured by some moment of the structure factor) should depend in
507: the same way on time $t$ as any other ({\it e.g.}, as measured by the
508: total amount of interface present in the system). Within each regime,
509: dimensional analysis can be used to construct the functional form of
510: $L(t)$ out of the model parameters that are relevant to that
511: regime~\cite{siggia,furukawa-aip-34-703-1985}.
512: %
513: \begin{itemize}
514: \item In the diffusive regime, the system is unaware of the density
515: $\rho$ and the viscosity $\eta$. Out of the remaining parameters
516: $M,\sigma$, and the time $t$ we can construct only one lengthscale:
517: %
518: \be
519: \label{eqn:LD}
520: L_{\rm D}(t)=\left(M\sigma t\right)^{1/3}.
521: \ee
522: %
523: In our units $L_{\rm D}= \alpha\, l \,t^{1/3}$ with $\alpha=
524: (2\sqrt{2}/3)^{1/3}$.
525:
526: \item In the viscous hydrodynamic regime, the system is unaware of $M$
527: and $\rho$. From the remaining parameters $\eta,\sigma$ and $t$ we
528: have
529: %
530: \be
531: \label{eqn:LV}
532: L_{\rm V}(t)=\frac{\sigma t}{\eta}.
533: \ee
534: %
535: In our units $L_{\rm V} = \alpha^3 l\, t/\eta$.
536:
537: \item In the inertial hydrodynamic regime, the system is unaware of
538: $M$ and $\eta$. From the remaining parameters $\rho,\sigma$ and
539: $t$ we have
540: %
541: \be
542: \label{eqn:LI}
543: L_{\rm I}(t)=\left(\frac{\sigma t^2}{\rho}\right)^{1/3}.
544: \ee
545: %
546: In our units $L_{\rm I}= \alpha \left(l\, t^2/\rho\right)^{1/3}$.
547:
548: \end{itemize}
549: %
550: By equating the growth laws $L_{\rm D} = L_{\rm V}$, we can construct
551: the characteristic lengthscale $\ldv$ at which we expect a crossover
552: from diffusive to viscous dynamics. Similarly by equating $L_{\rm
553: V}=L_{\rm I}$ we find the lengthscale $\lvi$ for crossover from
554: viscous to inertial dynamics. Together with the ``microscopic''
555: interfacial width $l$ and the system size $\Lambda_x,\Lambda_y$, we
556: then have the following basic lengthscales
557: %
558: \begin{itemize}
559: \item The interfacial width $l$.
560:
561: \item The characteristic lengthscale for crossover from diffusive to
562: viscous hydrodynamic coarsening:
563: %
564: \be
565: \ldv=\sqrt{M \eta}.
566: \ee
567: %
568: In our units $\ldv=l \sqrt{\eta}$.
569:
570: \item The characteristic lengthscale for crossover from viscous
571: to inertial hydrodynamic coarsening:
572: %
573: \be
574: \lvi=\frac{\eta^2}{\rho \sigma}.
575: \ee
576: %
577: In our units $\lvi=3\eta^2/2\sqrt{2}\rho l$.
578:
579: \item The system size $\Lambda_x,\Lambda_y$. In our units
580: $\Lambda_y=1$ always.
581: \end{itemize}
582:
583: Corresponding to these are the following timescales
584: %
585: \begin{itemize}
586:
587: \item The microscopic timescale for diffusion across
588: an interface
589: %
590: \be
591: \tau_0=\frac{2\sqrt{2}}{3}\frac{l^3}{M\sigma}=1.
592: \ee
593: %
594:
595: \item The characteristic timescale for crossover from
596: diffusive to viscous hydrodynamic coarsening
597: %
598: \be
599: \tdv=\frac{\sqrt{M \eta^3}}{\sigma}=\frac{3}{2\sqrt{2}}\eta^{3/2}.
600: \ee
601: %
602:
603: \item The characteristic timescale for crossover from viscous to
604: inertial hydrodynamic coarsening
605: %
606: \be
607: \tvi=\frac{\eta^3}{\rho\sigma^2}=\frac{9}{8}\frac{\eta^3}{\rho l^2}.
608: \ee
609: %
610:
611: \item The characteristic time at which the domain size attains the
612: system size in coarsening
613: %
614: \be
615: T_{\rm system}.
616: \ee
617: %
618: \end{itemize}
619:
620: %\vspace{0.4cm}
621: %In shear, we have a further timescale set by
622: %
623: %\begin{itemize}
624: %\item the reciprocal applied shear rate $\gdot^{-1}$.
625: %\end{itemize}
626: %
627:
628: So far we have discussed the growing structural lengthscale $L(t)$ in
629: general terms, without any specific definition. In fact, there exist
630: many possible measures of the characteristic domain size. In zero
631: shear we use the following one
632: %
633: \be
634: \label{eqn:L_structure}
635: L_s=\left(\frac{\int dq_x\int dq_y S(\vecv{q})}{\int dq_x\int dq_y |q|^{-1}S(\vecv{q})}\right)^{-1},
636: \ee
637: %
638: defined via the structure factor $S(\vecv{q})$, which, as usual, is
639: the 2D Fourier Transform of $\phi(x,y)$.
640:
641: The discussion so far in this section has related to unsheared
642: systems. Under an applied shear flow the domain morphology becomes
643: anisotropic. See Fig.~\ref{fig:steadyStates}, for example.
644: Accordingly, we consider the following
645: matrix~\cite{wagner-pre-59-4366-1999,stansell-prl-96--2006}
646: %
647: \be
648: \label{eqn:L_matrix}
649: D_{\alpha\beta}=\frac{l\int dx \int dy \,\partial_\alpha \phi \partial_\beta\phi}{\int dx \int dy \,\phi^2},
650: \ee
651: %
652: the reciprocal eigenvalues of which give us two lengthscales
653: $L_{\|},L_{\bot}$ characterising the long and short principal axes of
654: the domain morphology.
655:
656: \begin{table*}
657: \centering
658: \subtable[Zero applied shear. Viscous hydrodynamic to inertial hydrodynamic regime.]{
659: \begin{tabular}{|p{1.3cm}|p{1.4cm}|p{1.5cm}|p{1.5cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.2cm}|p{1.7cm}|p{2.0cm}|} \hline
660: Set & $\eta$ & $\rho$ & $l$ & $\gdot$ & $\Lambda_x$ & $N_x$ & $N_y$ & $\Dt$ & $L_{\rm DV}$ & $L_{\rm VI}$ \\ \hline \hline
661: R028u & 0.111 & 136.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.000519 & 0.0616 \\
662: R022u & 0.196 & 2510.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.000690 & 0.0104 \\
663: R029u & 0.0467 & 893.0 & 0.00156 & 0.0 & 1.0 & 512 &512 & 0.008 & 0.000337 & 0.00166 \\
664: R020u & 0.0785 & 16180.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.000437 & 0.000259 \\
665: R030u & 0.0122 & 6250.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.000172 & 0.0000162 \\
666: R019u & 0.00876 & 32814.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.000146 & 0.00000159 \\
667: R032u & 0.00391 & 20067.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.008 & 0.0000975 & 0.000000518 \\
668: \hline
669: \end{tabular}
670: \label{tab:firsttable}
671: }
672: \qquad\qquad
673: \subtable[Zero applied shear. Diffusive to viscous hydrodynamic regime.]{
674: \begin{tabular}{|p{1.3cm}|p{1.4cm}|p{1.5cm}|p{1.5cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.2cm}|p{1.7cm}|p{2.0cm}|} \hline
675: Set & $\eta$ & $\rho$ & $l$ & $\gdot$ & $\Lambda_x$ & $N_x$ & $N_y$ & $\Dt$ & $L_{\rm DV}$ & $L_{\rm VI}$ \\ \hline \hline
676: %DV1u & 1000.0 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.0791 & $\infty$ \\
677: %DV2u & 100.0 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.0250 & $\infty$ \\
678: %DV3u & 10.0 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.00791 & $\infty$ \\
679: %DV4u & 1.0 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.00250 & $\infty$ \\
680: %DV5u & 0.3 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.00137 & $\infty$ \\
681: %DV6u & 0.1 & 0.0 & 0.0025 & 0.0 & 1.0 & 512 &512 & 0.016 & 0.000791 & $\infty$ \\
682: DV1u & 1000.0 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.0494 & $\infty$ \\
683: DV2u & 100.0 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.0156 & $\infty$ \\
684: DV3u & 10.0 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.00494 & $\infty$ \\
685: DV4u & 1.0 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.00156 & $\infty$ \\
686: DV5u & 0.3 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 & 512 & 0.016 & 0.000855 & $\infty$ \\
687: DV6u & 0.1 & 0.0 & 0.00156 & 0.0 & 1.0 & 512 &512 & 0.016 & 0.000494 & $\infty$ \\
688: \hline
689: \end{tabular}
690: \label{tab:secondtable}
691: }
692: \qquad\qquad
693: \subtable[Applied shear, with inertia.]{
694: \begin{tabular}{|p{1.3cm}|p{1.4cm}|p{1.5cm}|p{1.5cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.2cm}|p{1.7cm}|p{2.0cm}|} \hline
695: Set & $\eta$ & $\rho$ & $l$ & $\gdot$ & $\Lambda_x$ & $N_x$ & $N_y$ & $\Dt$ & $L_{\rm DV}$ & $L_{\rm VI}$ \\ \hline \hline
696: R028s & 0.111 & 136.0 & 0.00156 & 0.0765 & 2.0 & 1024 & 512 & 0.008 & 0.000519 & 0.0616 \\
697: R028bs & 0.444 & 8704.0 & 0.00156 & 0.0765 & 2.0 & 1024 & 512 & 0.008 & 0.00104 & 0.0154 \\
698: R022s & 0.196 & 2510.0 & 0.00156 & 0.0205 & 2.0 & 1024 & 512 & 0.008 & 0.000690 & 0.0104 \\
699: R022bs & 0.339 & 22590.0 & 0.00156 & 0.0355 & 2.0 & 1024 & 512 & 0.008 & 0.000908 & 0.00349 \\
700: R029s & 0.0467 & 893.0 & 0.00156 & 0.0341 & 2.0 & 1024 &512 & 0.008 & 0.000337 & 0.00166 \\
701: R029bs & 0.0809 & 8037.0 & 0.00156 & 0.0591 & 2.0 & 1024 &512 & 0.008 & 0.000444 & 0.000554 \\
702: R020s & 0.0785 & 16180.0 & 0.00156 & 0.0256 & 2.0 & 1024 & 512 & 0.008 & 0.000437 & 0.000259 \\
703: R030s & 0.0122 & 6250.0 & 0.00156 & 0.0410 & 2.0 & 1024 & 512 & 0.004 & 0.000172 & 0.0000162 \\
704: R019s & 0.00876 & 32814.0 & 0.00156 & 0.0251 & 2.0 & 1024 & 512 & 0.004 & 0.000146 & 0.00000159 \\
705: R032s & 0.00391 & 20067.0 & 0.00156 & 0.051 & 2.0 & 1024 & 512 & 0.004 & 0.0000975 & 0.000000518 \\
706: \hline
707: \end{tabular}
708: \label{tab:thirdtable}
709: }
710: \qquad\qquad
711: \subtable[Applied shear, without inertia.]{
712: \begin{tabular}{|p{1.3cm}|p{1.4cm}|p{1.5cm}|p{1.5cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.0cm}|p{1.2cm}|p{1.7cm}|p{2.0cm}|} \hline
713: Set & $\eta$ & $\rho$ & $l$ & $\gdot$ & $\Lambda_x$ & $N_x$ & $N_y$ & $\Dt$ & $L_{\rm DV}$ & $L_{\rm VI}$ \\ \hline \hline
714: %DV1s & 1000.0 & 0.0 & 0.0025 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.0791 & $\infty$ \\
715: %DV2s & 100.0 & 0.0 & 0.0025 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.0250 & $\infty$ \\
716: %DV3s & 10.0 & 0.0 & 0.0025 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.00791 & $\infty$ \\
717: %DV4s & 1.0 & 0.0 & 0.0025 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.00250 & $\infty$ \\
718: %DV5s & 0.3 & 0.0 & 0.0025 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.00137 & $\infty$ \\
719: %DV6s & 0.1 & 0.0 & 0.0025 & 0.03 & 2.0 & 1024 &512 & 0.016 & 0.000791 & $\infty$ \\
720: DV1s & 1000.0 & 0.0 & 0.00156 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.0494 & $\infty$ \\
721: DV2s & 100.0 & 0.0 & 0.00156 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.0156 & $\infty$ \\
722: DV3s & 10.0 & 0.0 & 0.00156 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.00494 & $\infty$ \\
723: DV4s & 1.0 & 0.0 & 0.00156 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.00156 & $\infty$ \\
724: DV5s & 0.3 & 0.0 & 0.00156 & 0.01 & 2.0 & 1024 & 512 & 0.016 & 0.000855 & $\infty$ \\
725: DV6s & 0.1 & 0.0 & 0.00156 & 0.03 & 2.0 & 1024 &512 & 0.016 & 0.000494 & $\infty$ \\
726: \hline
727: \end{tabular}
728: \label{tab:fourthtable}
729: }
730: \caption{Parameter values used in simulation runs.}
731: \label{tab:master}
732: \end{table*}
733:
734: \section{Parameter steering}
735: \label{sec:parameters}
736:
737: As discussed in Sec.~\ref{sec:units}, the physical control parameters
738: that must be prescribed in any simulation run are the fluid density
739: $\rho$ and viscosity $\eta$, the interfacial width $l$, the applied
740: shear rate $\gdot$, and the system's aspect ratio $\Lambda_x$.
741: (Recall that we have set $\Lambda_y = 1$, $G = 1$ and $\tau_0 = 1$.)
742: We must also specify the number of numerical mesh points $N_x$ and
743: $N_y$, and the numerical timestep $\delta_t$. We now discuss
744: appropriate choices of these parameter values that will allow us to
745: access the physical regimes of interest.
746:
747: In the absence of an applied shear flow, $\gdot=0.0$, our aim is to
748: study coarsening of the isotropic domain morphology. See
749: Fig.~\ref{fig:state_u}, for example. In each of these runs, we
750: consider a square simulation box with $\Lambda_x=\Lambda_y =
751: \Lambda=1$, in our units. In any given run, at any time $t$, our aim
752: is to ensure a separation of the four lengthscales
753: %
754: \be
755: \delta\ll l \ll L(t)\ll \Lambda.
756: \ee
757: %
758: Here $\delta=1/N$ is the mesh size, prescribed by the reciprocal of
759: the number $N=N_x=N_y$ of numerical mesh points in each spatial
760: dimension. Recall that $l$ is the width of the interface between
761: domains; $L(t)$ is the growing domain size; and $\Lambda\equiv 1$ the
762: system size. We thereby restrict ourselves to physical lengthscales
763: $l$ and $L(t)$ that lie between the mesh and system size. To make this
764: window as large as possible, we take $\delta$ as small as possible,
765: using as many grid points $N^2=1/\delta^2$ as is numerically feasible.
766: The results presented below have $N=512$, checked for convergence to
767: the limit $N\to\infty$ at fixed $l$ against $N=1024$ (not shown). We
768: set the interfacial width $l$ to the minimum for which interfaces are
769: still fully resolved by this grid, taking $l=0.00156$, checked for
770: convergence to the limit $l/\Lambda
771: \to 0$ against $l=0.0025$ and $l=0.005$ (not shown).
772:
773: In each simulation run, this leaves a window of approximate size $0.05
774: < L(t) < 0.25$ in which the domain size $L(t)$ is much larger than the
775: width $l$ of the interface between domains, and much smaller than the
776: system size $\Lambda=1$. Accordingly, we are only able to study a
777: small piece of the full curve $L(t)$ in each run. In different runs,
778: therefore, we vary the viscosity $\eta$ and density $\rho$ to ensure
779: that we cover all regimes of interest. As discussed in
780: Refs.~\cite{kendon-prl-83-576-1999,pagonabarraga-jsp-107-39-2002}, we
781: are then able to construct composite scaling curves $L(t)/\ldv$ and
782: $L(t)/\lvi$, each spanning several decades of scaled length and time.
783:
784: Below we present results for two different series of runs. In the
785: first, R028u to R032u in table~\ref{tab:firsttable}, we explore the
786: viscous and inertial hydrodynamic regimes, and the crossover between
787: them, by varying the crossover lengthscale $\lvi$. These are in fact
788: the parameter values used previously by Kendon and coworkers in their
789: lattice Boltzmann simulations~\cite{kendon-jfm-440-147-2001}, converted
790: into our units.
791: %Note that $\ldv\ll l$ in each case,
792: %ensuring negligible diffusion on the lengthscale of a macroscopic
793: %domain.
794:
795: In the second series of runs, DV1u to DV6u in
796: table~\ref{tab:secondtable}, we explore the diffusive and (again)
797: viscous hydrodynamic regimes, and the crossover between them.
798: Accordingly, we set $\rho=0$ such that $\lvi$ is infinite and the
799: inertial regime is never attained. Across the different runs we vary
800: the predicted crossover lengthscale $\ldv=l\sqrt{\eta}$ by sweeping
801: the viscosity in the range $\eta=10^{-2}\cdots \eta=10^4$. For the
802: largest viscosity values, the system explores the diffusive regime.
803: For the smallest values, the system passes into the viscous regime as
804: soon as well defined domains have formed, without any identifiable
805: regime of diffusive domain growth.
806:
807: %All the runs in table~\ref{tab:master} have an interfacial width set
808: %by $l=0.00156$, following Ref. However we found these to be only
809: %marginally converged with respect to mesh fineness, for the maximum
810: %number of mesh points feasible within present computational resources.
811: %We therefore repeated all the runs in the table for the larger value
812: %$l=0.0025$, maintaining {\em fixed values} of the physical ratios
813: %$\ldv/l, \lvi/l$. (This necessitated dividing the densities by the
814: %factor $(0.00250/0.00156)^2$.) Accordingly, each corresponding run
815: %should have the same basic physics, but now with greater resolution at
816: %the grid scale, and at the expense of finite size effects setting in
817: %earlier. Performing the same comparison ($l=0.00156$ versus
818: %$l=0.0025$) in the case of an applied shear will allow us to comment
819: %on finite size scaling in our inertia-free runs, for which we find no
820: %steady state free of finite size effects.
821:
822: To study the effect of an applied shear flow, we consider the same
823: sets of parameter values as in zero shear but now with non-zero values
824: of $\gdot$. In anticipation of the anisotropy that shear will induce,
825: we also double the length of cell in the flow direction, along with
826: $N_x$ to maintain a constant density of mesh points. See
827: tables~\ref{tab:thirdtable} and~\ref{tab:fourthtable}: as just
828: discussed, the parameters of any sheared run (R032s, for example) are
829: the same as those of the corresponding unsheared run (R032u), apart
830: from the values of $\gdot, \Lambda_x$ and $N_x$. The parameters of
831: runs R028s, R022s, R029s, R020s, R030s, R019s and R032s are those used
832: in the lattice Boltzmann study of Ref.~\cite{stansell-prl-96--2006},
833: converted into our units. The appearance of the new sets R028b, R022b
834: and R029b will be explained in Sec.~\ref{sec:inertia} below.
835:
836: Assuming for the moment -- in some cases incorrectly, as we shall show
837: below -- that any sheared system will attain a steady state with a
838: typical domain size set by the reciprocal shear rate, our aim would
839: then be to construct scaling plots $L(1/\gdot)$ analogous to those of
840: $L(t)$ for the zero-shear coarsening regime in
841: Fig.~\ref{fig:domain_u}. To obtain as much information as in
842: Fig.~\ref{fig:domain_u}, however, we would need to run for very many
843: values of the shear rate $\gdot$ to produce corresponding near
844: continuous segments of $L$ versus $1/\gdot$. This is prohibitive
845: within available computing resources. Accordingly, we ran for just
846: one shear rate for each set of (other) parameter values, marking the
847: scaled reciprocal shear rates recorded in tables~\ref{tab:thirdtable}
848: and~\ref{tab:fourthtable} with vertical arrows in
849: Fig.~\ref{fig:domain_u}.
850:
851: \section{Numerical results}
852: \label{sec:results}
853:
854: We now present the results of our numerical simulations. We start in
855: Sec.~\ref{sec:zeroShear} with the case of coarsening in zero shear,
856: before considering an applied shear flow with and without inertia in
857: Secs.~\ref{sec:inertia} and~\ref{sec:noInertia} respectively.
858: Sec.~\ref{sec:link} contains a linking discussion. The results in
859: Sec.~\ref{sec:noInertia} are (we believe) novel. In contrast, our aim
860: in Secs.~\ref{sec:zeroShear} and~\ref{sec:inertia} is simply to
861: reproduce behaviour seen earlier in the literature by lattice
862: Boltzmann
863: techniques~\cite{wagner-prl-80-1429-1998,kendon-jfm-440-147-2001,stansell-prl-96--2006}.
864: The purpose of this comparative part of our study is threefold: First,
865: and foremost, to develop confidence in our own algorithm and the code
866: via which it is implemented; second, to demonstrate the method used
867: here, which is potentially simpler and faster than lattice Boltzmann,
868: to be equally capable of capturing the physics of demixing, in 2D
869: at least; and third, to provide an independent check of some recent
870: results concerning nonequilibrium steady states under
871: shear~\cite{stansell-prl-96--2006}.
872:
873: In each run, we take as an initial condition at each grid point a
874: value of $\phi$ chosen at random from a flat probability distribution
875: between $-0.01$ and $+0.01$. We checked for independence with respect
876: to this initial condition (not shown) the statistical properties of
877: the domain morphologies presented below. This independence holds as
878: long as many domains are present, making the system self-averaging.
879: Accordingly, for each set of parameters we give results below for a
880: single simulation run only.
881:
882: \subsection{Zero shear}
883: \label{sec:zeroShear}
884:
885: Following a temperature quench into the two phase regime, domains of
886: each phase form. These progressively coarsen in size through the
887: action of surface tension, passing through growth regimes that are
888: successively dominated by diffusive, viscous and inertial dynamics.
889: Within these regimes, dimensional analysis predicts a functional
890: dependence of the characteristic domain size upon time of $L_{\rm
891: D}\propto t^{1/3}$, $L_{\rm V} \propto t$ and $L_{\rm I}\propto
892: t^{2/3}$ respectively, as discussed in Sec.~\ref{sec:lengthscales}.
893: Our aim in this section is to investigate this behaviour numerically,
894: as done previously by lattice Boltzmann
895: methods~\cite{wagner-prl-80-1429-1998,kendon-jfm-440-147-2001}.
896:
897: \begin{figure}[tb]
898: \subfigure{\includegraphics[scale=0.36]{./DVu.png}}
899: % Diffusive to viscous regime without shear.
900: % results/noShear/DVsets/DV{1,2,3,4,5,6}u_l0.00156_Dt0.0*_CG22
901: % Processed by scripts/domainSize with option 2, y, 2. Lower time
902: % cutoff 50.0. Upper domain cutoff (by this measure of structure
903: % factor) 0.015. Tdv for sets DV{1,2,3,4,5,6}u are respectively 33541,
904: % 1061, 33.54, 1.061, 0.174, 0.03354. t_int/Tdv for sets
905: % DV{1,2,3,4,5,6}u are respectively 0.0003, 0.003, 0.17, 20.0, 180.0,
906: % ?? 1/(gdot*Tdv) for sets DV{1,2,3,4,5,6}u are respectively 0.00298,
907: % 0.0943, 2.98, 94.3, 575, 994.
908: \hspace{0.4cm}
909: \subfigure{\includegraphics[scale=0.36]{./VIu.png}}
910: % Viscous to inertial regimes without shear. results files
911: % results/noShear/R0sets/R0*l0.00156*22. Processed by
912: % scripts/domainSize.pl with option 2, y, 2. Lower time cutoff of
913: % 40.0. Upper domain cutoff (by this measure of structure factor)
914: % 0.02.
915: \caption{ a) Scaled domain size versus scaled time for coarsening
916: in zero shear in the diffusive and viscous hydrodynamic regimes.
917: Segments from left to right correspond to runs DV1u ... DV6u. The
918: vertical arrows show the values of scaled reciprocal shear rate used
919: in each corresponding run DV1s ... DV6s. Circles show the times of
920: the top two snapshots of Fig.~\ref{fig:state_u} below.
921: \\\hspace{0.4cm} b) Scaled domain size versus scaled time for
922: coarsening in zero shear in the viscous and inertial hydrodynamic
923: regimes. Segments from left to right correspond to runs R028u,
924: R022u, R029u, R020u, R030u, R019u, R032u. The vertical arrows show
925: the values of scaled reciprocal shear rate used in each
926: corresponding run R028s ... R032s. Circles show the times of the
927: bottom two snapshots of Fig.~\ref{fig:state_u} below.}
928: \label{fig:domain_u}
929: \end{figure}
930:
931:
932: \begin{figure}[tb]
933: \includegraphics[scale=0.6]{./stateD.png}
934: \includegraphics[scale=0.6]{./stateV1.png}
935: \vspace{0.05cm}
936: \includegraphics[scale=0.6]{./stateV2.png}
937: \includegraphics[scale=0.6]{./stateI.png}
938: \caption{Snapshots of domain morphology during coarsening in zero
939: shear at the times shown by circles in Fig.~\ref{fig:domain_u}. Top
940: left: diffusive regime run DV1u at $t=1052$. Top right: viscous
941: hydrodynamic regime run DV5u at $t=131$. Bottom left: viscous
942: hydrodynamic regime run R022 at $t=78.9$. Bottom right: inertial
943: hydrodynamic regime run R019 at $t=78.9$.}
944: % Diffusive:
945: % results/noShear/DVsets/states/DV1u_l0.00156_Dt0.016_CG22_state.
946: % State at time 1052. (t-tint)/tdv is then 0.029 Viscous 1:
947: % results/noShear/DVsets/states/DV5u_l0.00156_Dt0.016_CG22_state State
948: % at time 131.6. (t-tint)/tdv is then 572. Viscous 2:
949: % results/noShear/R0sets/states/R022_l0.00156_*22_state,
950: % state at time 78.96. At this time L=0.00775. For this set
951: % Lvi=0.0104. So L/Lvi=0.745.
952: % Inertial:
953: % results/noShear/R0sets/states/R019_l0.00156_*22_state,
954: % state at time 78.96. At this time L=0.0757. For this set
955: % Lvi=0.00000159. So L/Lvi=4761
956: \label{fig:state_u}
957: \end{figure}
958:
959:
960: As noted in Sec.~\ref{sec:parameters}, only a small segment of the
961: full coarsening curve $L(t)$ can be explored in any simulation run.
962: Accordingly, in different runs we vary the viscosity $\eta$ and
963: density $\rho$. In this way we correspondingly vary the crossover
964: length scales $\ldv$ and $\lvi$. We then combine the data sets from
965: the different runs into composite scaling curves of $L/\ldv$ and
966: $L/\lvi$, each spanning several decades in scaled length and time. We
967: choose as our measure of $L$ the one defined in
968: Eqn.~\ref{eqn:L_structure} via the structure factor.
969:
970: To explore the diffusive and viscous hydrodynamic regimes, and the
971: crossover between them, we performed a single simulation run for each
972: set of parameter values of table~\ref{tab:secondtable}. The composite
973: curve of $L/\ldv$ is shown in Fig.~\ref{fig:domain_u}a). For each
974: segment of the curve, three manipulations were performed. (i) The
975: first $t=0-40$ time units were discarded to allow for transient
976: dynamics during the initial process of domain formation. (ii) A
977: time-offset $t_{\rm int}$ was subtracted from the time $t$ to allow
978: for these non-universal dynamics during the initial transient. In each
979: case this subtraction was performed by eye. The appropriate $t_{\rm
980: int}$ was chosen as that which gives the most convincing straight
981: line on a log-log plot, without any known power present to bias the
982: eye during this process. (iii) The data set was cutoff at long times
983: once the typical domain size approaches the system size. For this
984: measure of the domain size, we used $L<0.02$ as a very conservative
985: criterion ensuring this. The same manipulations were performed for
986: runs in table~\ref{tab:firsttable}, to be described below.
987:
988: For small values of $L/\ldv$, in the diffusive regime, we recover the
989: power $L_{\rm D} \propto t^{1/3}$ predicted by dimensional analysis.
990: This was confirmed previously by lattice Boltzmann
991: simulations~\cite{wagner-prl-80-1429-1998}. For large values of
992: $L/\ldv$, in the viscous hydrodynamic regime, dimensional analysis
993: predicts the growth law $L_{\rm V}\propto t$. We instead find $L
994: \propto t^{2/3}$ for each run. This anomalous scaling been observed
995: previously, by lattice Boltzmann simulations. It is believed to stem
996: from a breakdown of scale invariance in this regime in 2D, due to the
997: formation of disconnected droplets~\cite{wagner-prl-80-1429-1998}. As
998: a result of this breakdown, different measures of $L$ depend
999: differently on time (not shown here). In 3D (not studied here) scale
1000: invariance is recovered, along with the $2/3$ growth
1001: exponent~\cite{kendon-jfm-440-147-2001}.
1002:
1003: To study the inertial hydrodynamic regime and (again) the viscous
1004: hydrodynamic regime, and the crossover between them, we use the
1005: parameter values of table~\ref{tab:firsttable}. The composite curve of
1006: $L/\lvi$ is shown in Fig.~\ref{fig:domain_u}b). For large values of
1007: $L/\lvi$, in the inertial regime, we clearly recover the predicted
1008: power $L_{\rm I} \propto t^{2/3}$. This has been seen before, in three
1009: dimensions, by lattice Boltzmann
1010: simulations~\cite{kendon-jfm-440-147-2001}. For small values of
1011: $L/\lvi$, in the viscous hydrodynamic regime, we again find the
1012: anomalous power $L \propto t^{2/3}$ for each run. This is consistent
1013: with the breakdown of scale invariance in 2D, and its attendant
1014: departure from the predictions of dimensional
1015: analysis~\cite{wagner-prl-80-1429-1998}. Because of this breakdown,
1016: segments R028, R022, R029 and R029 are not colinear: we do not have a
1017: single composite scaling curve in this regime. The fact that each
1018: segment follows a $t^{2/3}$ power, apparently the same as in the
1019: inertial regime, is a coincidence resulting from having chosen the
1020: particular measure of domain size given by Eqn.~\ref{eqn:L_structure}.
1021: As already noted, the lack of scaling in this regime results in
1022: different power laws for different measures of the domain
1023: size~\cite{wagner-prl-80-1429-1998}.
1024:
1025:
1026:
1027: \subsection{Under shear, with inertia}
1028: \label{sec:inertia}
1029:
1030:
1031: We now turn to phase separation in the presence of an applied shear
1032: flow. The main question of interest here is whether shear arrests
1033: coarsening and restores a nonequilibrium steady state with a finite
1034: domain size set by the reciprocal shear rate, independent of the
1035: system size; or whether coarsening persists indefinitely, up to the
1036: system size, even under shear.
1037:
1038: Despite previous
1039: experimental~\cite{hashimoto-jcp-88-5874-1988,takahashi-jr-38-699-1994,chan-pra-43-1826-1991,krall-prl-69-1963-1992,hashimoto-prl-74-126-1995,beysens-pra-28-2491-1983,chan-prl-61-412-1988,matsuzaka-prl-80-5441-1998,hobbie-pre-54-R5909-1996,qiu-pre-58-R1230-1998,lauger-prl-75-3576-1995},
1040: numerical~\cite{shou-pre-61-R2200-2000,zhang-jcp-113-8348-2000,yamamoto-pre-59-3223-1999,corberi-prl-81-3852-1998,corberi-prl-83-4057-1999,corberi-pre-61-6621-2000,corberi-pre-62-8064-2000,rothman-prl-65-3305-1990,rothman-el-14-337-1991,chan-el-11-13-1990,qiu-jcp-108-9529-1998,padilla-jcp-106-2342-1997,ohta-jcp-93-2664-1990,wagner-pre-59-4366-1999,lamura-pamia-294-295-2001,lamura-epjb-24-251-2001,berthier-pre-6305--2001}
1041: and
1042: theoretical~\cite{onuki-jpm-9-6119-1997,lamura-pamia-294-295-2001,corberi-prl-81-3852-1998,corberi-prl-83-4057-1999,corberi-pre-61-6621-2000,corberi-pre-62-8064-2000,bray-ptrslsapes-361-781-2003,cavagna-pre-62-4702-2000,bray-jpag-33-L305-2000,Onuk97,doi-jcp-95-1242-1991}
1043: work, this question remained open until the recent simulation study of
1044: Stansell et al. in Ref.~\cite{stansell-prl-96--2006}. Using lattice
1045: Boltzmann techniques, they found nonequilibrium steady states of the
1046: type reproduced by our own simulations in Fig.~\ref{fig:steadyStates}
1047: (discussed below). As expected under shear, the domain morphology is
1048: anisotropic. Accordingly, two length scales are needed to
1049: characterise it. The domain lengths $L_x$ and $L_y$ for the steady
1050: states were extracted in Ref.~\cite{stansell-prl-96--2006} via the
1051: curvature tensor of Eqn.~\ref{eqn:L_matrix}, and scaling plots of
1052: $L_{\{x,y\}}/\lvi$ versus $1/\gdot \tvi$ were constructed. These
1053: suggested apparent scaling exponents $L_x \sim \gdot^{-2/3}$ and
1054: $L_y\sim \gdot^{-3/4}$, sustained over six decades. The same scalings
1055: $L_{\|}\sim \gdot^{-2/3}$ and $L_{\bot}\sim \gdot^{-3/4}$ were
1056: suggested for the major and minor principal lengths of the domains.
1057: Slightly different scalings, discussed below, were obtained by the
1058: same group in a more recent 3D study~\cite{cates-3d-2007}.
1059:
1060: In this section we aim to show that our simulations, which are in 2D
1061: throughout, reproduce these results to good approximation. We will
1062: thereby gain confidence in our technique, before proceeding to the
1063: novel contribution of this work in Sec.~\ref{sec:noInertia} below.
1064: Accordingly, we perform a single simulation run for each set of
1065: parameter values used in Ref.~\cite{stansell-prl-96--2006}, converted
1066: into our units. See R028, R022, R029, R020, R030, R019 and R032 in
1067: table~\ref{tab:thirdtable}. The small discrepancy in the imposed
1068: values of $1/\gdot \tvi$, evident in Fig.~\ref{fig:scalingPlot}, stems
1069: from a slightly different definition adopted for the interfacial
1070: width, realised by this author only at a late stage of the present
1071: study.
1072:
1073: \begin{figure}[tbp]
1074: % results/shear/R0sets/R0{28b,22b,29b,30,20,19,32}*l0.00156*22 processed using scripts/domainSize.pl, with options 3 0 3 {l,s}
1075: \subfigure{\includegraphics[scale=0.36]{./larger.png}}
1076: \subfigure{\includegraphics[scale=0.36]{./smaller.png}}
1077: \caption{Top: larger domain length {\it vs.} strain $\gdot t$. Data sets for
1078: decreasing values of long term temporal average correspond to R020s,
1079: R030s, R028b, R029bs, R022bs, R019s, R032s. Bottom: smaller domain
1080: length {\it vs.} strain. Data sets for decreasing values of long
1081: term temporal average correspond to R020s, R030s, R019s, R029bs,
1082: R032s, R022bs, R028b. Colour online.}
1083: \label{fig:timeSeries}
1084: \end{figure}
1085:
1086: \begin{figure}[tbp]
1087: % results/shear/R0sets/R0{28b,22b,29b,30,20,19,32}*l0.00156*22 processed using scripts/scalingDomainPlot.pl, with options 3 2 b and lower strain cutoff of 75
1088: \subfigure{ \includegraphics[scale=0.38]{./scalingPlot.png}}
1089: \caption{Dimensionless scaling plot of lengths {\it vs.} shear
1090: rate. Solid symbols: average of time series of
1091: Fig.~\ref{fig:timeSeries} for $L_{\|}, L_{\bot}$ for strains
1092: $\gdot t>100$. Symbols from left to right correspond to R028bs,
1093: R022bs, R029bs, R020s, R030s, R019s, R032s. Errors bars show the
1094: standard deviation. Solid lines: power law fits to the solid
1095: symbols, suggesting $L_{\|} \sim \gdot^{-0.619}$ and $L_{\bot}\sim
1096: \gdot^{-0.693}$. Upper set of crosses shows the scaled system size
1097: $\Lambda_y/\lvi$; lower set shows the scaled interfacial width
1098: $l/\lvi$. Open symbols: data from
1099: Ref.~\cite{stansell-prl-96--2006} for $L_{\|}, L_{\bot}$,
1100: reproduced here for comparison by kind permission of the authors
1101: of that study.}
1102: \label{fig:scalingPlot}
1103: \end{figure}
1104:
1105: A typical run took one to two weeks of wall-clock time on a Linux box,
1106: given exclusive use of a single 3.4GHz Intel Xeon CPU with 2Mb cache
1107: and 400MHz DDR2 memory. Approximately $70\%$ of the runtime appears to
1108: be used switching back and forth between real and reciprocal space at
1109: each timestep.
1110:
1111: In each run, we monitored as a function of time the characteristic
1112: domain sizes $L_{\|}$ and $L_{\bot}$ extracted from the tensor of
1113: Eqn.~\ref{eqn:L_matrix}. For each of R020, R030, R019 and R032 we
1114: found $L_{\|}$ and $L_{\bot}$ to saturate at long times, showing
1115: temporal fluctuations about constant mean values
1116: (Fig.~\ref{fig:timeSeries}). For the remainder of this section we
1117: focus only on these statistically steady states, neglecting the first
1118: $\gdot t=100$ strain units of each run. (This cutoff was chosen by a
1119: simple visual inspection of Fig.~\ref{fig:timeSeries}.) In each case,
1120: finite size effects appear under control, as seen in the order
1121: parameter snapshots of Fig.~\ref{fig:steadyStates}. The mean values
1122: of the time series are shown on the scaling plot of $L/\lvi$ versus
1123: $1/\gdot\tvi$ in Fig.~\ref{fig:scalingPlot}, with error bars showing
1124: the standard deviation. A snapshot of the order parameter for run R032
1125: is shown in Fig.~\ref{fig:steadyStates} (bottom).
1126:
1127: For runs R028, R022 and R029, in contrast, we found the domains to
1128: wrap completely round the system in the flow direction, eventually
1129: comprising trivial horizontal stripes connected at the edges of the
1130: cell by the periodic boundary conditions. We believe this to be due to
1131: the horizontal system size $\Lambda_x$ being dangerously close to the
1132: expected values of $L_{\|}$ for these runs, leading eventually to
1133: nucleation of these stripes.
1134:
1135: To eliminate this effect, we performed new runs R028b, R022b and R029b
1136: at the same prescribed values of $1/\gdot \tvi$ as for R028, R022 and
1137: R029, but now for larger values of the scaled system size
1138: $\Lambda_x/\lvi$. This was achieved by adjusting $\lvi$ at fixed
1139: $\Lambda_x, \gdot\tvi$. Effectively, the left three sets of crosses in
1140: Fig.~\ref{fig:scalingPlot} have been slightly shifted upward with
1141: respect to those in Ref.~\cite{stansell-prl-96--2006} These new runs
1142: produced nonequilibrium steady states, as seen in the time series of
1143: Fig.~\ref{fig:timeSeries}. The long term time averages of these series
1144: are shown in the scaling plot of Fig.~\ref{fig:scalingPlot}. A
1145: snapshot of the order parameter for run R022b is shown in
1146: Fig.~\ref{fig:steadyStates} (top).
1147:
1148: Of course we cannot rule out eventual nucleation of system-wrapped
1149: stripes at times exceeding those accessed numerically. The same
1150: comment applies to runs R028, R029 and R022 of
1151: Ref.~\cite{stansell-prl-96--2006}. Conversely, the system-wrapped
1152: stripes seen in the corresponding runs in the recent 3D lattice
1153: Boltzmann study~\cite{cates-3d-2007} might well be eliminated by
1154: adjusting the scaled system size as done here.
1155:
1156: % Reasonable quantitative agreement is found with the data of Ref. for
1157: % $L_x$ and $L_y$. Complete agreement is not expected, since we
1158: % instead give $L_{\|}$ and $L_{\bot}$.
1159:
1160: \begin{figure}[tbp]
1161: \subfigure{ \includegraphics[scale=0.56]{./R022bs_l0p00156_Dt0p008_CG22_state_time03157p92_phi.png}}
1162: \subfigure{ \includegraphics[scale=0.56]{./R032s_l0p00156_Dt0p004_CG22_state_time03157p896_phi.png}}
1163: \caption{Snapshots of the steady state order parameter for R022b at strain $\gdot t=112$ (top) and R032 at strain $\gdot t=161$ (bottom).}
1164: \label{fig:steadyStates}
1165: \end{figure}
1166:
1167:
1168:
1169:
1170: Power law fits to our data in the scaling plot of
1171: Fig.~\ref{fig:scalingPlot} suggest exponents $L_{\|}\sim
1172: \gdot^{-0.619\pm 0.01}$ and $L_{\bot}\sim ^{-0.693\pm 0.007}$. The
1173: quoted uncertainties are the standard deviations given by the
1174: automated regression package, and do not include systematic errors in
1175: the simulations. In contrast Ref.~\cite{stansell-prl-96--2006} found
1176: exponents for $L_{\|}, L_{\bot}$ of $-0.678\pm 0.039$ and $-0.756\pm
1177: 0.03$, by fitting to the open symbols that are reproduced by kind
1178: permission in Fig.~\ref{fig:scalingPlot} for comparison with our data.
1179: It also quoted exponents for $L_{x}, L_{y}$ of $-0.678\pm 0.042$ and
1180: $-0.759\pm 0.029$. The more recent 3D study~\cite{cates-3d-2007} by
1181: the same group found exponents for $L_{\|}, L_{\bot}$ of $-0.64\pm
1182: 0.06$ and $-0.67\pm 0.03$, with $-0.53\pm0.04$ in the third dimension.
1183: Within their $O(10\%)$ margins of error, these agree with the
1184: exponents for $L_{\|}$ and $L_{\bot}$ found in the 2D study of the
1185: present work.
1186:
1187: \subsection{Discussion}
1188: \label{sec:link}
1189:
1190: In this section, we will discuss in more detail the shape of the
1191: scaling plot of $L/\lvi$ versus $1/\gdot \tvi$ in
1192: Fig.~\ref{fig:scalingPlot}. In doing so, we will motivate the further
1193: numerical study of Sec.~\ref{sec:noInertia} below.
1194:
1195: Translated into equivalent scaled times $t/\tvi$, the range of scaled
1196: shear rates $1/\gdot\tvi$ explored in Fig.~\ref{fig:scalingPlot} would
1197: span the viscous and inertial hydrodynamic regimes in the coarsening
1198: plot of Fig.~\ref{fig:domain_u}b). One possibility for the shape of
1199: the scaling curve $L/\lvi$ versus $1/\gdot\tvi$ under shear is that it
1200: should be the same as that of the corresponding coarsening plot of
1201: $L/\lvi$ versus $t/\tvi$ in zero shear, given the simple substitution
1202: $t=\gdot^{-1}$. In view of this, it is instructive to compare
1203: Fig.~\ref{fig:scalingPlot} with Fig.~\ref{fig:domain_u}b) in some
1204: detail.
1205:
1206: In the coarsening case (Fig.~\ref{fig:domain_u}b), two distinct
1207: regimes are apparent. The inertial hydrodynamic regime, on the right
1208: side, shows the $t^{2/3}$ exponent predicted by dimensional analysis.
1209: Each segment neatly lines up with the next, consistent with dynamical
1210: scaling. In contrast, in the viscous hydrodynamic regime on the left
1211: hand side the segments do not align. Furthermore, each departs from
1212: the predicted $t^1$ power. As noted above, these anomalous features
1213: stem from non-scaling effects that arise in 2D. In 3D these are
1214: suppressed and the predicted $t^1$ power law is recovered. The
1215: counterpart of Fig.~\ref{fig:domain_u}b) for 3D systems thus comprises
1216: a clean $t^1$ in the viscous regime on the left hand side, and
1217: $t^{2/3}$ in the inertial regime on the right hand side, with a rather
1218: slow crossover in between. See Fig. 9 of
1219: Ref.~\cite{kendon-jfm-440-147-2001}. An applied shear flow is also
1220: anticipated to suppress these non-scaling effects. Accordingly, the
1221: 2D results of Fig.~\ref{fig:scalingPlot} are expected capture the
1222: basic physics and, indeed, compare favourably with the recent 3D
1223: sheared study of Ref.~\cite{kendon-jfm-440-147-2001}.
1224:
1225: With these remarks in mind, it is now clear that we should in fact
1226: compare Fig.~\ref{fig:scalingPlot} for sheared systems with Fig. 9 of
1227: Ref.~\cite{kendon-jfm-440-147-2001} for coarsening systems. As noted
1228: above, one might naively expect the two plots to have the same shape,
1229: up to the substitution $t=\gdot^{-1}$.
1230:
1231: Instead, two differences are clearly apparent. First,
1232: Fig.~\ref{fig:scalingPlot} has two characteristic lengths, in contrast
1233: to the single length that characterises unsheared systems in
1234: Fig.~\ref{fig:domain_u}b). With hindsight this is obvious: the domain
1235: morphology is anisotropic in sheared systems, and so characterised by
1236: two lengths.
1237:
1238: %The slightly different scaling with shear rate of $L_{\|}$ and
1239: %$L_{\bot}$ suggests that these two lengthscales will eventually merge
1240: %for large values of $1/\gdot\tvi$, and separate in the limit $1/\gdot
1241: %\tvi \to 0$. In principle, this separation would lead to a crisis in
1242: %which the domains assume an infinite aspect ratio. This possibility
1243: %is to be treated with some caution, however, because systematic
1244: %uncertainties not accounted for in our data analysis do not rule out a
1245: %common scaling for both lengths; and because a crisis-avoiding
1246: %crossover might occur for values of $1/\gdot\tvi$ smaller than those
1247: %accessed in Fig.~\ref{fig:scalingPlot}.
1248:
1249: The second, and more subtle, difference is that sheared systems
1250: apparently lack a distinction between viscous and inertial regimes,
1251: with a single power law spanning all six decades in
1252: Fig.~\ref{fig:scalingPlot}. A possible explanation for this is that
1253: the range of scaled shear rates in Fig.~\ref{fig:scalingPlot} actually
1254: occupies a window of extremely slow crossover between a true viscous
1255: regime at smaller $1/\gdot\tvi$ and a true inertial regime at larger
1256: $1/\gdot\tvi$, reminiscent of the slow crossover seen in the
1257: coarsening plot in Ref.~\cite{kendon-jfm-440-147-2001}. Another is
1258: that the value of $1/\gdot \tvi$ at which the crossover occurs in the
1259: sheared case is much smaller than the corresponding crossover value of
1260: $t/\tvi$ in the unsheared case, which is formally $O(1)$ but in
1261: practice lies around $10^4$. The results in Fig.~\ref{fig:scalingPlot}
1262: would then all lie in the inertial regime, to the right of this
1263: crossover.
1264:
1265: To test these ideas, in the next section we perform simulations at
1266: strictly zero Reynolds number, $\rho=0$. We thereby take the limit
1267: $\gdot \tvi \to \infty$ at the outset and perform simulations for
1268: several finite values of $1/\gdot \tdv$. When converted into
1269: equivalent scaled times $t/\tdv$ these span the diffusive and viscous
1270: hydrodynamic regimes, as shown by the vertical arrows in
1271: Fig.~\ref{fig:domain_u}a). By analogy with the fact that the 3D
1272: counterpart of the viscous regime of Fig.~\ref{fig:domain_u}a) would
1273: be expected in the limit $t/\tdv \to\infty$ to match into that of the
1274: viscous regime in Fig. 9 of Ref.~\cite{kendon-jfm-440-147-2001} in the
1275: limit $t/\tvi\to 0$, we might then expect to access, for large
1276: $1/\gdot \tdv \to\infty$, the scalings that $L_{\|}$ and $L_{\bot}$
1277: would attain in the limit $1/\gdot\tvi \to 0$ at the extreme left of
1278: an enlarged Fig.~\ref{fig:scalingPlot}.
1279:
1280: %Any discrepancy between these
1281: %scalings and the scalings $L_{\|}\sim \gdot^{-0.619}, L_{\bot}\sim
1282: %^{-0.693}$ already observed might indeed point to a slow crossover in
1283: %Fig.~\ref{fig:scalingPlot}; or alternatively to a much smaller
1284: %crossover value of $1/\gdot\tvi$ for sheared systems than of $t/\tvi$
1285: %in the unsheared case.
1286:
1287: Instead, however, we will find no evidence for non equilibrium steady
1288: states in inertialess systems. In contrast, coarsening appears to
1289: persist indefinitely, up to the system size, despite the applied shear
1290: flow. Given the existence of non equilibrium steady states for values
1291: of $1/\gdot\tvi$ at the left hand edge of Fig.~\ref{fig:scalingPlot},
1292: where the effects of inertia are non-zero but likely to be small, this
1293: suggests that inertia plays the role of a singular perturbation in
1294: this problem. Indeed, we will argue in Sec.~\ref{sec:argument} below
1295: that the single scaling seen for each of $L_{\|}$ and $L_{\bot}$
1296: across Fig.~\ref{fig:scalingPlot} results from a {\em mixed}
1297: visco-inertial regime across the entire plot. The crisis of infinite
1298: aspect ratio $L_{\|}/L_{\bot}\to \infty$ in the limit $1/\gdot/\tvi\to
1299: 0$, suggested by the data in Fig.~\ref{fig:scalingPlot}, is consistent
1300: with our suggestion that inertia plays a singular role.
1301:
1302:
1303: \begin{figure}[tbp]
1304: \subfigure{\includegraphics[scale=0.36]{./larger_zero_inertia.png}}
1305: \subfigure{\includegraphics[scale=0.36]{./smaller_zero_inertia.png}}
1306: \caption{Top: larger domain length $L_{\|}$ {\it vs.} strain $\gdot
1307: t$. Bottom: smaller domain length $L_{\bot}$ {\it vs.} strain
1308: $\gdot t$. In each case, decreasing values of $L$ at fixed $\gdot
1309: t=15$ correspond to runs DV5s, DV6s, DV4s, DV3s, DV2s, DV1s.}
1310: \label{fig:timeSeries_zero_inertia}
1311: \end{figure}
1312:
1313: \begin{figure*}[tbp]
1314: \subfigure[DV1s, $\gdot t=78.9$]{\includegraphics[scale=0.4]{./DV1s_l0p00156_Dt0p008_CG22_state_time07894p8_phi.png}}
1315: \subfigure[DV2s, $\gdot t=78.9$]{\includegraphics[scale=0.4]{./DV2s_l0p00156_Dt0p008_CG22_state_time07894p8_phi.png}}
1316: \subfigure[DV3s, $\gdot t=78.9$]{\includegraphics[scale=0.4]{./DV3s_l0p00156_Dt0p008_CG22_state_time07894p8_phi.png}}
1317: \subfigure[DV4s, $\gdot t=63.2$]{\includegraphics[scale=0.4]{./DV4s_l0p00156_Dt0p004_CG22_state_time06315p808_phi.png}}
1318: \subfigure[DV5s, $\gdot t=55.3$]{\includegraphics[scale=0.4]{./DV5s_l0p00156_Dt0p004_CG22_state_time05526p332_phi.png}}
1319: \subfigure[DV6s, $\gdot t=47.4$]{\includegraphics[scale=0.4]{./DV6s_l0p00156_Dt0p002_CG22_state_time01578p948_phi.png}}
1320: \caption{Snapshot domain morphologies after many strain units $S=\gdot
1321: t$ for reciprocal shear rates that, as equivalent times $\gdot^{-1}$,
1322: would be shown by the arrows in Fig.~\ref{fig:domain_u}a).
1323: These runs have zero inertia, $\rho=0$.}
1324: \label{fig:spanning}
1325: \end{figure*}
1326:
1327: \subsection{Under shear, without inertia}
1328: \label{sec:noInertia}
1329:
1330: Motivated by the discussion of the previous section we now consider
1331: phase separation under shear in inertialess systems, setting $\rho=0$.
1332: As just discussed, by studying the limit $t/\tdv\to\infty$ we might
1333: then, a priori, have expected to gain some insight into the far left
1334: hand side of an enlarged Fig.~\ref{fig:scalingPlot}, $1/\gdot \tvi \to
1335: 0$. Accordingly, we perform a single simulation run for each of the
1336: parameter sets DV1s to DV6s in table~\ref{tab:fourthtable}. The values
1337: of the scaled reciprocal shear rates $1/\gdot \tdv$ are marked as
1338: corresponding scaled times $t/\tdv$ on the coarsening plot of
1339: Fig.~\ref{fig:domain_u}a), where they are seen to span both the
1340: diffusive and viscous hydrodynamic regimes.
1341:
1342: In each run we monitored the characteristic domain sizes $L_{\|}$ and
1343: $L_{\bot}$ as a function of time. As can be seen in
1344: Fig.~\ref{fig:timeSeries_zero_inertia}, in each case the larger length
1345: $L_{\|}$ grows without bound until it attains the system size $O(1)$.
1346: Correspondingly, a snapshot of the order parameter after many strain
1347: units reveals system-wrapped domains, with pronounced finite-size
1348: effects. See Fig.~\ref{fig:spanning}. The snapshots for DV1s to DV3s
1349: strongly resemble those reported in
1350: Ref.~\cite{berthier-pre-6305--2001} for model B, which lacks
1351: hydrodynamics. This is consistent with the fact that these runs occupy
1352: the diffusive regime when marked as equivalent scaled times in
1353: Fig.~\ref{fig:domain_u}b).
1354:
1355: Beyond the parameter sets of table~\ref{tab:fourthtable}, we
1356: furthermore performed runs for shear rates $\gdot=0.1, 0.03, 0.01,
1357: 0.003$ at each value of the viscosity $\eta=1000.0, 100.0, 10.0, 1.0,
1358: 0.3, 0.1$, thereby covering a complete rectangle in
1359: $(\eta,\gdot)$-space, in contrast to the single slice taken by DV1s to
1360: DV6s. (For historical reasons these runs had a slightly larger value
1361: $l=0.0025$ for the interfacial width, but we do not expect this to
1362: make a qualitative difference.) We found no evidence in any run for a
1363: nonequilibrium steady state, unlimited by finite-size effects.
1364:
1365: These results clearly suggest that the limit $1/\gdot\tdv\to\infty$,
1366: which is approximated by runs DV5s and DV6s, does not match the limit
1367: $1/\gdot\tvi\to 0$, which is approximated by runs R028b, R022b. In
1368: the next section, we propose that this is because the limit
1369: $1/\gdot\tdv\to\infty$ corresponds to a pure viscous regime in which
1370: no steady state exists; while the limit $1/\gdot\tvi\to 0$ corresponds to a
1371: mixed visco-inertial regime in which a steady state does exist.
1372:
1373: \section{Absence of nonequilibrium steady states in inertialess
1374: systems}
1375: \label{sec:argument}
1376:
1377: We now present an analytical argument in support of the above
1378: numerical observation: that nonequilibrium steady states are not
1379: attained in inertialess systems. As we shall find, a closely related
1380: argument supports the existence of a single power law for each of
1381: $L_{\|}, L_{\bot}$ across the whole of Fig.~\ref{fig:scalingPlot}, as
1382: seen first in Ref.~\cite{stansell-prl-96--2006}. Recall that, a
1383: priori, we might have expected this plot to comprise separate viscous
1384: and inertial regimes.
1385:
1386: As a preliminary step, we recall the case of domain coarsening in an
1387: unsheared system. As discussed above, as time proceeds the system
1388: passes through regimes that are successively dominated by diffusive,
1389: viscous and inertial dynamics. In each regime, the system is aware of
1390: the surface tension $\sigma$; the time $t$; and either the mobility
1391: $M$ (diffusive regime), the viscosity $\eta$ (viscous regime) or the
1392: density $\rho$ (inertial regime). In each case, there exists only one
1393: possible combination of these relevant parameters that has the
1394: dimensions of a length. Thus we have the following predicted growth
1395: laws for the characteristic domain size:
1396: %
1397: \be
1398: \label{eqn:coarsening}
1399: L_{\rm D}=(M\sigma t)^{1/3},\;L_{\rm V}=\frac{\sigma t}{\eta},\;\;\textrm{and}\;\;L_{\rm I}=\left(\frac{\sigma t^2}{\rho}\right)^{1/3}.
1400: \ee
1401: %
1402:
1403: Consider next a nonequilibrium steady state under shear. Neglecting
1404: fluctuations, the time $t$ is now irrelevant. In its place, however,
1405: we gain the reciprocal shear rate as a relevant parameter. If
1406: ``pure'' diffusive, viscous and inertial regimes were still to exist,
1407: then we could again only construct a {\em single} length scale to
1408: characterise each:
1409: %
1410: \be
1411: L_{\rm Ds}=(M\sigma \gdot^{-1})^{1/3},\;\;\;L_{\rm Vs}=\frac{\sigma \gdot^{-1}}{\eta},\;\;\textrm{and}\;\;L_{\rm Is}=\left(\frac{\sigma \gdot^{-2}}{\rho}\right)^{1/3},
1412: \ee
1413: %
1414: by direct comparison with (\ref{eqn:coarsening}).
1415:
1416: In contrast, however, our numerical results clearly show the domain
1417: morphology to be anisotropic, as expected in a sheared system. It is
1418: therefore characterised by {\em two} lengths, which (according to
1419: Fig.~\ref{fig:scalingPlot}) scale differently with the applied shear
1420: rate. A steady state under shear therefore cannot exist in a ``pure''
1421: diffusive, viscous or inertial regime, because in each there are
1422: insufficient parameters out of which to construct the two lengthscales
1423: needed to characterise it.
1424:
1425: How can we retain enough parameters, out of the candidates $\sigma,
1426: \gdot, M, \rho, \eta$ and $t$, to construct the two lengthscales
1427: needed to characterise an anisotropic domain morphology under shear?
1428: Assuming that $\sigma$ and $\gdot$ are always relevant, two options
1429: are as follows.
1430:
1431: \begin{enumerate}
1432:
1433: \item
1434:
1435: If the system is to exist in a {\em pure diffusive, viscous or
1436: inertial regime}, with knowledge of just one of $M, \eta$ or
1437: $\rho$ (as well as $\sigma$ and $\gdot$), it must retain dependence
1438: on time $t$. Therefore, it must {\em fail to attain a steady
1439: state}.
1440:
1441: Assuming power laws, a purely diffusive regime would then have
1442: %
1443: \be
1444: L_{\rm Ds}=(M \sigma t)^{1/3}(\gdot t)^a,
1445: \ee
1446: %
1447: in which $a$ assumes two different values, which would be prescribed
1448: by more detailed physics than is contained in the present argument.
1449: Likewise, a purely viscous regime would have
1450: %
1451: \be
1452: L_{\rm Vs}=\frac{\sigma t}{\eta}(\gdot t)^b,
1453: \ee
1454: %
1455: again with two different values for $b$.
1456:
1457:
1458: \item
1459:
1460: If the system is to attain a {\em steady state}, thereby losing
1461: knowledge of the time $t$, it must retain dependence upon at least
1462: two of $M,\eta$ and $\rho$. For example, a shear-induced steady
1463: state that is free of diffusion on the lengthscale of domains (no
1464: $M$) must exist in a {\em mixed viscous-inertial regime} with
1465: knowledge of both viscosity $\eta$ and density $\rho$:
1466: %
1467: \be
1468: L_{\rm VIs}=\left(\rho^{-c-1}\eta^{c}\sigma \gdot^{-2-c}\right)^{\frac{1}{2c+3}},
1469: \ee
1470: %
1471: with two different values for $c$.
1472:
1473: \end{enumerate}
1474:
1475: We propose that both of these cases are seen in our numerical
1476: simulations. When inertia is present, for finite values of $\rho$
1477: (however small), the system exists in a mixed visco-inertial steady
1478: state, as discussed in option 2. This is consistent with the
1479: observation of a single power law scaling for each of $L_{\|}$ and
1480: $L_{\bot}$ across all six decades in the plot of $L/\lvi$ versus
1481: $1/\gdot\tvi$ in Fig.~\ref{fig:scalingPlot} and in
1482: Ref.~\cite{stansell-prl-96--2006}. In contrast, when inertia is
1483: strictly absent ($\rho=0$, infinite $\gdot\tvi$), the system exists in
1484: a pure diffusive or viscous regime and never attains a steady state,
1485: as in option 1.
1486:
1487: To summarise: we suggest that the limit $1/\gdot\tdv\to\infty$
1488: corresponds to a pure viscous regime with no steady state, while the
1489: limit $1/\gdot\tvi\to 0$ corresponds to a mixed visco-inertial steady
1490: state. Thus we propose finally that inertia provides the role of a
1491: singular perturbation in this problem.
1492:
1493: \section{Summary and outlook}
1494: \label{sec:conclusions}
1495:
1496: We have studied numerically phase separation in binary fluids with
1497: full hydrodynamics in two dimensions, considering both (1) unsheared
1498: and (2) sheared systems, both (a) with and (b) without inertia. Of
1499: these, cases (1a), (1b) and (2a) have been studied by previous authors
1500: using Lattice Boltzmann (LB)
1501: methods~\cite{wagner-prl-80-1429-1998,wagner-pre-59-4366-1999,stansell-prl-96--2006,kendon-jfm-440-147-2001}.
1502: In this paper, we have introduced an alternative simulation technique
1503: that uses finite-differencing and spectral methods. We have also used
1504: a convenient transformation to render trivial the implementation of
1505: Lees-Edwards sheared periodic boundary
1506: conditions~\cite{onuki-jpsj-66-1836-1997}.
1507:
1508: In unsheared systems, phase separation occurs via a process of domain
1509: coarsening. Our simulation method successfully recovers results
1510: obtained previously by LB for this process, both with and without
1511: inertia (1a, 1b above). In particular, it finds the familiar power
1512: law $L_{\rm D}\sim t^{1/3}$ characterising the growth of the typical
1513: domain size in the diffusive regime~\cite{wagner-prl-80-1429-1998}. It
1514: also recovers $L_{\rm I}\sim t^{2/3}$ in the inertial hydrodynamic
1515: regime~\cite{kendon-jfm-440-147-2001}. In the viscous hydrodynamic
1516: regime it finds the anomalous power $L \sim t^{2/3}$, compared to the
1517: predicted one of $L_{\rm v}\sim t^1$. As noted by previous authors,
1518: this is due to subtle non scaling effects that arise in 2D from the
1519: formation of disconnected droplets, also seen in LB
1520: studies~\cite{wagner-prl-80-1429-1998}. Such effects are eliminated
1521: in 3D~\cite{kendon-jfm-440-147-2001} and also seem suppressed under
1522: shear\cite{wagner-pre-59-4366-1999,stansell-prl-96--2006}.
1523:
1524: We have also successfully reproduced the observations of existing LB
1525: studies for sheared systems that have non-zero inertia (case 2a
1526: above)~\cite{stansell-prl-96--2006}. Here, an applied shear flow
1527: arrests domain coarsening and restores a nonequilibrium steady state
1528: with domains of a finite size set by the inverse shear rate. The
1529: domain morphology is anisotropic, characterised by two lengthscales
1530: $L_{\|}, L_{\bot}$. Scaling exponents $L_{\|}\sim \gdot^{-0.619}$ and
1531: $L_{\bot}\sim \gdot^{-0.693}$ found here agree with those of LB, to
1532: within margins of error. An outstanding puzzle, however, is why these
1533: 2D exponents agree better with the exponents $L_{\|}\sim
1534: \gdot^{-0.64}, L_\bot \sim \gdot^{-0.67}$ of the 3D LB
1535: study~\cite{cates-3d-2007}, than those of the 2D LB
1536: study~\cite{stansell-prl-96--2006}, $L_{\|}\sim \gdot^{-0.678}$ and
1537: $L_\bot\sim\gdot^{-0.756}$. To investigate this, it would be
1538: interesting to study the role of systematic errors in our simulations;
1539: to consider the possible eventual nucleation of system wrapping
1540: stripes in the 2D LB study of Ref.~\cite{stansell-prl-96--2006} for the
1541: smaller values of $1/\gdot\tvi$; and to extend the present
1542: finite-differencing work to 3D.
1543:
1544: Our successful recovery of these important existing results in regimes
1545: (1a, 1b, 2a) provides some confidence in our simulation method.
1546: Beyond thereby having demonstrated this method to be capable of
1547: capturing the physics of demixing, the other main contribution of this
1548: paper has been a novel study of phase separation in sheared systems
1549: that are strictly inertialess (case 2b). Here we found no evidence of
1550: nonequilibrium steady states, free of pronounced finite size effects.
1551: Instead coarsening appears to persist indefinitely until the typical
1552: domain size attains the system size, as in zero shear.
1553:
1554: To support this observation, we have suggested by means of a simple
1555: analytical argument that sheared inertialess systems adopt either a
1556: pure diffusive or pure viscous regime, in each of which there are
1557: insufficient parameters out of which to construct the two lengthscales
1558: needed to characterise an anisotropic domain morphology in steady
1559: state. By extending this argument slightly, we have also suggested
1560: that sheared systems with any amount of inertia, however small, exist
1561: in a mixed visco-inertial steady state. This provides a possible
1562: explanation for the observation of a single scaling with shear rate
1563: for each of $L_{\|}$ and $L_{\bot}$ across all six decades in
1564: Fig.~\ref{fig:scalingPlot}, and in the corresponding plot of
1565: Ref.~\cite{stansell-prl-96--2006}
1566:
1567: If this suggestion is correct, it remains unclear why the viscous and
1568: inertial regimes should mix to yield a steady state, while the
1569: diffusive and viscous regimes apparently remain separate, precluding
1570: nonequilibrium steady states in truly inertialess systems. A possible
1571: explanation lies in the more severe nonlinearity (in $\vecv{v}$) of
1572: the inertial terms in the equation of motion.
1573:
1574: In future work, we aim to investigate whether the absence of
1575: nonequilibrium steady states in inertialess systems persists in 3D.
1576: In extending our method to 3D, several challenges are to be faced. Of
1577: these, the main ones appear to be contained in the basic Navier-Stokes
1578: equations of incompressible fluid flow, and are not complicated
1579: significantly by any additional order parameters~\cite{mixedLB}:
1580: $\phi$ in this model. However, relatively standard methods do exist
1581: for finite-differencing the incompressible Navier-Stokes equations in
1582: 3D, as discussed in Ref.~\cite{Pozrikidis}. At each timestep these
1583: involve updating the vorticity via a slightly modified equation of
1584: motion $\partial_t \omega=\cdots$; then updating the velocity field
1585: either in the velocity-vorticity formulation (as effectively done here
1586: in 2D via the intermediate of the streamfunction) or in the vector
1587: potential-vorticity formulation. In neither case does the pressure
1588: field need to be calculated directly. An outstanding issue before any
1589: such algorithm could be run efficiently concerns its scaling with
1590: system size, and its corresponding ease of parallelisation. The same
1591: question is also relevant in 2D, and a direct comparison of our
1592: algorithm with that of Ref.~\cite{stansell-prl-96--2006} would clearly
1593: be interesting. In the longer term, the outcome of such studies might
1594: help determine whether finite differencing can emerge as a useful tool
1595: in such problems, alongside the already tested and reliable LB
1596: methods.
1597:
1598: Other open questions concern the role of initial conditions in sheared
1599: systems. All the simulations reported here consider a temperature
1600: quench performed in the presence of a shear flow. Future work should
1601: consider an already demixed system, with either a flat interface or a
1602: minimal-surface droplet, subsequently subject to a sudden shear
1603: startup. We also aim to address demixing in complex macromolecular
1604: fluids, focusing on the role of viscoelasticity.
1605:
1606: \section*{Acknowledgements}
1607:
1608: The author thanks Prof. Alan Bray, Prof. Mike Cates and Prof. Peter
1609: Sollich for helpful discussions and feedback, and the UK's EPSRC
1610: GR/S29560/02 for funding.
1611:
1612: \appendix
1613:
1614: \bw
1615:
1616: \section{Transformation to the cosheared frame}
1617: \label{sec:appendix1}
1618:
1619: Here we give details of the transformation to the cosheared
1620: frame~\cite{onuki-jpsj-66-1836-1997}. As discussed in
1621: Sec.~\ref{sec:numerics} above, this is performed in order to render
1622: trivial the numerical implementation of sheared periodic boundary
1623: conditions. As a first step, we separate the velocity field into a
1624: constant affine contribution $\gdot y\xhat$ and a fluctuating part
1625: $\vt$
1626: %
1627: \be
1628: \vecv{v}(\vecv{x},t)=\gdot y\xhat + \vt(\vecv{x},t).
1629: \ee
1630: %
1631: Noting that $\gdot y \xhat$ automatically satisfies the
1632: incompressibility condition, we define the fluctuating parts of the
1633: streamfunction and vorticity via
1634: %
1635: \be
1636: \label{eqn:streamT}
1637: \vt=\nablu\wedge \st \zhat,
1638: \ee
1639: %
1640: and
1641: %
1642: \be
1643: \label{eqn:streamVortT}
1644: \wt=-\nabla^2\st.
1645: \ee
1646: %
1647: Our equation set then comprises (\ref{eqn:streamT}) and
1648: (\ref{eqn:streamVortT}) together with
1649: %
1650: \be
1651: \label{eqn:vorticityT}
1652: \rho\left(\partial_t\,\wt+\vt.\nablu\,\wt\right)+\rho \gdot y\, \partial_x\wt=\eta\nabla^2\wt-\left[\nablu\wedge\phi\nablu\mu\right]\cdot\zhat,
1653: \ee
1654: %
1655: from Eqn.~\ref{eqn:vorticity},
1656: %
1657: \be
1658: \label{eqn:phiT}
1659: \partial_t\,\phi + \vt.\nablu\,\phi +\gdot y\,\partial_x\phi=l^2\nabla^2\mu,
1660: \ee
1661: %
1662: from Eqn.~\ref{eqn:phi1}, and
1663: %
1664: %
1665: \be
1666: \label{eqn:muAppendix}
1667: \mu= \phi(\phi^2-1)- l^2 \nabla^2\phi,
1668: \ee
1669: %
1670: from Eqn.~\ref{eqn:mu1}, unchanged.
1671:
1672:
1673: We then make a transformation to the cosheared frame
1674: %
1675: \be
1676: \label{eqn:trans1}
1677: (x,y,t)\to(x'=x-\gdot t y, y'=y,t'=t).
1678: \ee
1679: %
1680: %
1681: The various partial derivatives then become
1682: %
1683: \be
1684: (\partial_x,\partial_y,\partial_t) = (\partial_{x'},-\gdot t\partial_{x'}+\partial_{y'},-\gdot y\partial_{x'}+\partial_{t'}).
1685: \ee
1686: %
1687: Accordingly, we define the cosheared 2D gradient operator
1688: %
1689: \be
1690: \nablu_c=\xhat\, \partial_{x'}+\yhat\,(-\gdot t\partial_{x'}+\partial_{y'}).
1691: \ee
1692: %
1693: Finally, for any function $a$ we write
1694: %
1695: \be
1696: \label{eqn:trans4}
1697: a(x,y,t)=A(x',y',t').
1698: \ee
1699: %
1700: Throughout we continue to work with velocity components
1701: $\tilde{v}_x,\tilde{v}_y$ and not $\tilde{v}_{x'},\tilde{v}_{y'}$.
1702:
1703: Inserting Eqn.~\ref{eqn:streamT} into the $\vt\cdot\nablu$ terms on
1704: the LHS of Eqns.~\ref{eqn:vorticityT} and~\ref{eqn:phiT}, and
1705: performing the transformation (\ref{eqn:trans1}) to (\ref{eqn:trans4})
1706: on Eqns.~\ref{eqn:streamVortT} to~\ref{eqn:muAppendix}, we get the
1707: equation set
1708: %
1709: \be
1710: \tilde{\Omega}=-\nabla_c^2\tilde{\Psi},
1711: \ee
1712: %
1713: together with
1714: %
1715: \be
1716: \label{eqn:T2}
1717: \rho\,\partial_{t'}\tilde{\Omega}+\rho\left(\partial_{y'}\tilde{\Psi}\partial_{x'}\tilde{\Omega}-\partial_{x'}\tilde{\Psi}\partial_{y'}\tilde{\Omega}\right)=\eta\,\nabla_c^2\tilde{\Omega}-\left[\partial_{x'}\Phi\partial_{y'} M-\partial_{y'}\Phi \partial_{x'} M\right],
1718: \ee
1719: %
1720: \be
1721: \label{eqn:T3}
1722: \partial_{t'}\Phi+\left(\partial_{y'}\tilde{\Psi}\partial_{x'}\Phi-\partial_{x'}\tilde{\Psi}\partial_{y'}\Phi\right)=l^2\nabla_c^2M,
1723: \ee
1724: %
1725: and
1726: %
1727: \be
1728: M=\Phi(\Phi^2-1)-l^2\nabla_c^2\Phi.
1729: \ee
1730: %
1731: %
1732: In these $M$ represents upper case $\mu$, not mobility. A priori, the
1733: bracketed expressions in Eqns.~\ref{eqn:T2} and~\ref{eqn:T3} contain
1734: terms in the applied shear rate $\gdot$. However in each bracket these
1735: are equal and opposite, and so cancel. For clarity we finally drop the
1736: tildes and dashes, and revert from upper to lower case. The final
1737: governing equations are then as summarised in Sec.~\ref{sec:numerics}
1738: above.
1739:
1740: \section{Numerical method}
1741: \label{sec:appendix2}
1742:
1743: Here we discuss the numerical algorithm used to study the dynamical
1744: evolution of $\phi(\vecv{x},t),\omega(\vecv{x},t)$ and
1745: $\psi(\vecv{x},t)$, as specified by Eqns.~\ref{eqn:equation1}
1746: to~\ref{eqn:equation4} and~\ref{eqn:coshearedNabluFinal}. Our basic
1747: strategy is to step along a grid of time values $t^n=n\Delta t$ for
1748: $n=1,2,3\cdots$. Discretization with respect to time of any quantity
1749: $f$ is denoted $f(t^n)=f^n$, or sometimes by $f|^n$. At each timestep,
1750: we update $\phi^n, \psi^n, \omega^n \to \phi^{n+1}, \psi^{n+1},
1751: \omega^{n+1}$ in three separate stages. First, we update the
1752: compositional order parameter $\phi^n \to \phi^{n+1}$ according to
1753: Eqns.~\ref{eqn:equation3} and~\ref{eqn:equation4} with fixed, old
1754: values of the stream-function $\psi^n$. We then update $\omega^n \to
1755: \omega^{n+1}$ using Eqn.~\ref{eqn:equation2} at fixed
1756: $\phi^{n+1},\psi^n$. Finally we update the streamfunction
1757: $\psi^n\to\psi^{n+1}$ using Eqn.~\ref{eqn:equation1} at fixed
1758: $\omega^{n+1}$.
1759:
1760: \begin{enumerate}
1761:
1762: \item The update $\phi^n\to\phi^{n+1}$ using Eqns.~\ref{eqn:equation3}
1763: and~\ref{eqn:equation4} is performed in two successive partial
1764: updates. In the first we implement the advective term in
1765: Eqn.~\ref{eqn:equation3} to give $\phi^n \to \phi^{n+1/2}$. In the
1766: second we implement the diffusive term to give $\phi^{n+1/2} \to
1767: \phi^{n+1}$. The advective term is handled using an explicit Euler
1768: algorithm~\cite{PreTeuVetFla92}. Temporarily setting aside the issue
1769: of spatial discretization, this can be written
1770: %
1771: \be
1772: \label{eqn:advective}
1773: \phi^{n+1/2}(x,y)=\phi^n - \Delta t \; \left(\partial_{y}{\psi}^n\partial_{x}\phi^n-\partial_{x}{\psi}^n\partial_{y}\phi^n\right).
1774: \ee
1775: %
1776: This is then spatially discretized on a rectangular grid of
1777: $(\Lambda_x N/\Lambda_y) \times N$ mesh points in real space $(x-y)$,
1778: with constant mesh intervals $\Dx=\Dy=\Lambda_y/N$. Using indices
1779: $i=1\cdots \Lambda_x N/\Lambda_y$ and $j=1\cdots N$, we denote any
1780: discretized function $f(x_i,y_j)=f_{ij}$, or sometimes $f|_{ij}$.
1781: Periodic boundary conditions are imposed by setting
1782: $f_{i(-1)}=f_{i({N}-1)}$, $f_{i0}=f_{iN}$, $f_{i(N+1)}=f_{i1}$,
1783: $f_{i({N}+2})=f_{i2}$, and similarly in the $x$ direction. The
1784: derivatives of $\psi$ in Eqn.~\ref{eqn:advective} are discretized as
1785: follows:
1786: %
1787: \be
1788: \label{eqn:p1y}
1789: \partial_x\psi|_{ij}^n=\frac{1}{2\Dx}\left[\psi_{(i+1)j}-\psi_{(i-1)j}\right],
1790: \ee
1791: %
1792: with
1793: %
1794: \be
1795: \partial_y\psi|_{ij}^n=\frac{1}{2\Dy}\left\{\left(1-\frac{s}{2}\right)\left(\psi_{i(j+1)}-\psi_{i(j-1)}\right) + \frac{s}{2}\left(\psi_{(i-2)(j+1)}-\psi_{(i+2)(j-1)}\right)\right\}\;\;\textrm{if}\;\;\; s\ge 0,
1796: \ee
1797: %
1798: and
1799: %
1800: \be
1801: \partial_y\psi|_{ij}^n=\frac{1}{2\Dy}\left\{\left(1+\frac{s}{2}\right)\left(\psi_{i(j+1)}-\psi_{i(j-1)}\right) - \frac{s}{2}\left(\psi_{(i+2)(j+1)}-\psi_{(i-2)(j-1)}\right)\right\}\;\;\textrm{if}\;\;\; s<0.
1802: \ee
1803: %
1804: The derivatives of $\phi$ in Eqn.~\ref{eqn:advective} are discretized
1805: in the same way.
1806:
1807: %We verified
1808: %that this gives the same results as using more sophisticated
1809: %third-order upwinding~\cite{Pozrikidis}:
1810: %%
1811: %\be
1812: %\partial_y\phi_{ij}^n=\frac{1}{6\Dy}\left[\phi^n_{i(j-2)} -6 \phi^n_{i(j-1)} +3
1813: % \phi^n_{ij} +2 \phi^n_{i(j+1)}
1814: %\right]\;\;\;\textrm{if}\;\;\; v_y|^n_{ij}>0,
1815: %\ee
1816: %
1817: %while
1818: %
1819: %\be
1820: %\partial_y\phi_{ij}^n=\frac{1}{6\Dy}\left[-\phi^n_{i(j+2)} +6 \phi^n_{i(j+1)} -3
1821: % \phi^n_{ij} -2 \phi^n_{i(j-1)}
1822: %\right]\;\;\;\textrm{if}\;\;\; v_y|^n_{ij}<0,
1823: %\ee
1824: %
1825: %with analogous expressions for the derivative of $\phi$ with
1826: %respect to $x$.
1827:
1828:
1829: It then remains to implement the diffusive part of Eqn.~\ref{eqn:equation3}:
1830: %
1831: \be
1832: \partial_t\phi=-l^2\,\nabla_c^2\phi-l^4\,\nabla_c^4\phi+l^2\,\nabla_c^2 f
1833: \ee
1834: %
1835: with $f=\phi^3$. After calculating $f$ on our rectangular grid in real
1836: space, this equation is handled in reciprocal space by taking fast
1837: Fourier transforms $x\to q_x$ and $y\to q_y$ using a standard NAG
1838: routine~\cite{NAG}. The transformation in each dimension generates a
1839: real and an imaginary part, so for each mode $\vecv{q}=(q_x,q_y)$ we
1840: need to consider a vector of the (transposed) form
1841: $\vecv{\phi}^T=(\phi_{\rm rr}, \phi_{\rm ir}, \phi_{\rm ri}, \phi_{\rm
1842: ii})$ where subscript ``r'' denotes real part, and ``i'' imaginary.
1843: The respective transforms $\tens{D}_2$ and $\tens{D}_4$ of the
1844: operators $l^2\,\nabla_c^2$ and $l^4\,\nabla_c^4$ can easily be found
1845: analytically:
1846: %
1847: \be
1848: \tens{D}_2=l^2\left(
1849: \begin{array}{cccc}
1850: D & 0 & 0 & -\Delta \\
1851: 0 & D & \Delta & 0 \\
1852: 0 & \Delta & D & 0 \\
1853: -\Delta & 0 & 0 & D
1854: \end{array}
1855: \right)
1856: \;\;\;\;\textrm{and}\;\;\;\;
1857: \tens{D}_4=l^4\left(
1858: \begin{array}{cccc}
1859: D^2+\Delta^2 & 0 & 0 & -2D\Delta \\
1860: 0 & D^2+\Delta^2 & 2D\Delta & 0 \\
1861: 0 & 2D\Delta & D^2+\Delta^2 & 0 \\
1862: -2D\Delta & 0 & 0 & D^2+\Delta^2
1863: \end{array}
1864: \right),
1865: \ee
1866: %
1867: in which
1868: %
1869: \vspace{-0.4cm}
1870: \be
1871: D=-(aq_x^2+q_y^2),\;\;\;\Delta=bq_xq_y\;\;\;\textrm{with}\;\;\;a=1+[s(t)]^2,\;\;\;b=2s(t).
1872: \ee
1873: %
1874: For each $\vecv{q}$-mode, we then have
1875: %
1876: \be
1877: \partial_t\vecv{\phi}=-\tens{D}_2\cdot\vecv{\phi}-\tens{D}_4\cdot\vecv{\phi}+\tens{D}_2\cdot\vecv{f}.
1878: \ee
1879: %
1880: To evolve this in time, we use an explicit Euler algorithm for the
1881: first and third terms, and a semi-implicit Crank-Nicolson
1882: algorithm for the second term. Thus we have
1883: %
1884: \be
1885: \vecv{\phi}^{n+1}-\vecv{\phi}^{n+1/2}=-\tilde{\tens{D}}_2\cdot\vecv{\phi}^{n+1/2}-\tfrac{1}{2}\tilde{\tens{D}}_4\cdot(\vecv{\phi}^{n+1}+\vecv{\phi}^{n+1/2})+\tilde{\tens{D}}_2\cdot\vecv{f}^n,
1886: \ee
1887: %
1888: in which $\tilde{\tens{D}}_m=\Dt\,\tens{D}_m$ for $m=2,4$. Rearranging gives finally
1889: %
1890: \be
1891: \vecv{\phi}^{n+1}=\vecv{\phi}^{n+1/2} + (\tens{\delta}+\tfrac{1}{2}\tilde{\tens{D}}_4)^{-1}\cdot\left( - \tilde{\tens{D}}_2\cdot\vecv{\phi}^{n+1/2} - \tilde{\tens{D}}_4 \cdot\vecv{\phi}^{n+1/2} +\tilde{\tens{D}}_2\cdot\vecv{f}^n\right).
1892: \ee
1893: %
1894:
1895: \item We now update $\omega^n \to \omega^{n+1}$ using
1896: Eqn.~\ref{eqn:equation2} at fixed $\phi^{n+1},\psi^n$. The
1897: advective term on the LHS is updated in the same way as its
1898: counterpart in the $\phi$ equation above. To avoid inefficient
1899: multiple switching between real and Fourier space, this is in fact
1900: done at the same time as the corresponding update of $\phi$ in 1.
1901: above. (This reordering leaves the algorithm exactly unchanged.)
1902: We then update the RHS of Eqn.~\ref{eqn:equation2}. Divided across
1903: by $\rho$, this reads:
1904: %
1905: \be
1906: \label{eqn:vortNum}
1907: \partial_t\,\omega=\nu\,\nabla_c^2 \omega+ g,
1908: \ee
1909: %
1910: in which $\nu=\eta/\rho$ and
1911: %
1912: \be
1913: \label{eqn:G}
1914: g=-\frac{1}{\rho}\left[\partial_{x}\phi\,\partial_{y} \mu-\partial_{y}\phi \,\partial_{x} \mu\right].
1915: \ee
1916: %
1917: As a first step, we calculate $g(x,y)$ using the newly updated
1918: $\phi^{n+1}$ from 1. To do so, we first calculate
1919: %
1920: \be
1921: \mu=\phi(\phi^2-1)-l^2(1+[s(t)]^2)\partial_x^2\phi - l^2\partial_y^2\phi + 2l^2s\partial_x\partial_y\phi.
1922: \ee
1923: %
1924: At each grid point $i,j$, we spatially discretize the derivatives in
1925: this expression according to
1926: %
1927: \be
1928: \partial^2_x\phi|_{ij}=\frac{1}{(\Dx)^2}\left[\phi_{(i+1)j}-2\phi_{ij}+\phi_{(i-1)j}\right],
1929: \ee
1930: %
1931: similarly for $\partial^2_y\phi$, and
1932: %
1933: \be
1934: \partial_x\partial_y\phi|_{ij}=\frac{1}{4 \Dx\Dy}\left[\phi_{(i+1)(j+1)}-\phi_{(i+1)(j-1)}-\phi_{(i-1)(j+1)}+\phi_{(i-1)(j-1)}\right].
1935: \ee
1936: %
1937: The first order derivatives of $\phi$ and $\mu$ with respect to $x$
1938: and $y$ in~(\ref{eqn:G}) are then discretized as
1939: in~(\ref{eqn:p1y}).
1940:
1941: Having calculated $g(x,y)$ in real space, we then Fourier transform
1942: Eqn.~\ref{eqn:vortNum} to get
1943: %
1944: \be
1945: \partial_t\,\vecv{\omega}=\tens{C}\cdot\vecv{\omega}+\vecv{g},
1946: \ee
1947: %
1948: in which we have used the same vector/matrix notation as in 1 above,
1949: with $\tens{C}=\nu \tens{D}_2/l^2$. To evolve this in time, we use a
1950: semi-implicit Crank-Nicolson algorithm for the first term on the RHS
1951: to get
1952: %
1953: \be
1954: \label{eqn:forOmega}
1955: \vecv{\omega}^{n+1}-\vecv{\omega}^n=\tfrac{1}{2}\tilde{\tens{C}}\cdot(\vecv{\omega}^n+\vecv{\omega}^{n+1})+\tilde{\vecv{g}}^{n+1}
1956: \ee
1957: %
1958: with $\tilde{\tens{C}}=\Dt\, \tens{C}$ and $\tilde{\vecv{g}}=\Dt\,
1959: \vecv{g}$. The superscript $n+1$ on the last term (\ref{eqn:forOmega})
1960: serves to remind us that $g$ was calculated using the new $\phi^{n+1}$
1961: from part 1. Rearranging, we get finally
1962: %
1963: \be
1964: \vecv{\omega}^{n+1}=\vecv{\omega}^n+(\tens{\delta}-\tfrac{1}{2}\tilde{\tens{C}})^{-1}\cdot(\tilde{\tens{C}}\cdot\vecv{\omega}^n+\tilde{\vecv{g}}^{n+1}).
1965: \ee
1966: %
1967:
1968: \item We finally update the streamfunction $\psi^n\to\psi^{n+1}$ using
1969: Eqn.~\ref{eqn:equation1} at fixed $\omega^{n+1}$. For each
1970: $\vecv{q}-$mode, we have
1971: %
1972: \be
1973: \vecv{\psi}^{n+1}=-\tens{E}\cdot\vecv{\omega}^{n+1}\;\;\;\textrm{in which}\;\;\;\tens{E}=\frac{1}{D^2-\Delta^2}\left(
1974: \begin{array}{cccc}
1975: D & 0 & 0 & \Delta \\
1976: 0 & D & -\Delta & 0 \\
1977: 0 & -\Delta & D & 0 \\
1978: \Delta & 0 & 0 & D
1979: \end{array}
1980: \right).
1981: \ee
1982: %
1983: \end{enumerate}
1984: %
1985: Finally, we transform all functions back to real space and return to
1986: step 1 to start the next timestep.
1987:
1988: \vspace{-0.4cm}
1989:
1990: \subsection*{Algorithm at zero Reynolds number}
1991:
1992: \vspace{-0.2cm}
1993:
1994: The algorithm discussed so far is suited to non-zero values of the
1995: fluid density $\rho$. At zero Reynolds number, with $\rho=0$,
1996: Eqn.~\ref{eqn:equation1} and~\ref{eqn:equation2} of our basic equation
1997: set combine to give the simpler equation
1998: %
1999: \be
2000: \label{eqn:equationNew}
2001: 0=-\eta\,\nabla_c^4{\psi}-\left[\partial_{x}\phi\,\partial_{y} \mu-\partial_{y}\phi \,\partial_{x} \mu\right],
2002: \ee
2003: %
2004: and we need no longer consider the vorticity field $\omega$.
2005: Equations~\ref{eqn:equation3} and~\ref{eqn:equation4} remain
2006: unchanged. Correspondingly, step 1 of our algorithm is also
2007: unchanged. Steps 2 and 3 now combine to give
2008: %
2009: \be
2010: \vecv{\psi}^{n+1}=\tens{E}.\vecv{g}^{n+1},
2011: \ee
2012: %
2013: As above, we calculate $g$ in real space then take a Fourier transform.
2014: In Fourier space
2015: %
2016: \be
2017: \tens{E}=\frac{1}{(D^2+\Delta^2)^2-(2D\Delta)^2}\left(
2018: \begin{array}{cccc}
2019: D^2+\Delta^2 & 0 & 0 & 2D\Delta \\
2020: 0 & D^2+\Delta^2 & -2D\Delta & 0 \\
2021: 0 & -2D\Delta & D^2+\Delta^2 & 0 \\
2022: 2D\Delta & 0 & 0 & D^2+\Delta^2
2023: \end{array}
2024: \right).
2025: \ee
2026: %
2027: After calculating $\psi$, we revert to real space to start the next
2028: timestep.
2029:
2030: \ew
2031:
2032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2033: %\bibliographystyle{prsty}
2034: %\bibliography{ackerson,actin,ajdari,articles,banding,barham,Berret,berrportdecruppe,berthier,books,bray,callaghan,cates,chandcits,cook,crystal,crystal_theory,Decruppe,deGennes,demixing,dhont,dnatheory,elasticTurbulence,fielding,fischer,Fischer,flowcryst,fredrickson,furukawa,gelbart,gonnella,goveas,graham,Groisman,head,hebraud,helfrich,HinchRallison,hsiao,Kadoma,larson,LCtheory,leal,lerougeDecruppe,lifshitz,likhtman,line_tension,maffettone,malkus,master,mccoy,membs,Mexican,modelH,noirez,notes,olmsted,onions,onuki,otherRelated,pagonabarraga,phanThien,phd1,phd,pine,PineHu,pomeau,poon,pratt,psolutions,ramaswamy,recent,rheochaos,rheofolks,ryan,salmon,SalmonManneville,savedrecs.txt,schoot,semenov,sgrband,shaqfeh,sood,sriram,stein,sureshkumar,tanaka,vansaarloos,vorticity,Wang,warren,weitz,wilson,worms2,worms3,worms,yeomans,yuan,zubarev}
2035:
2036: \begin{thebibliography}{10}
2037:
2038: \bibitem{hohenberg77}
2039: P.~C. Hohenberg and B.~I. Halperin, Rev. Mod. Phys. {\bf 49}, 435 (1977).
2040:
2041: \bibitem{bray-aip-43-357-1994}
2042: A.~J. Bray, Adv. In Phys. {\bf 43}, 357 (1994).
2043:
2044: \bibitem{cates-fd--1-1999}
2045: M.~E. Cates, V.~M. Kendon, P. Bladon, and J.~C. Desplat, Faraday Discussions
2046: {\bf 112}, 1 (1999).
2047:
2048: \bibitem{hashimoto-jcp-88-5874-1988}
2049: T. Hashimoto, T. Takebe, and S. Suehiro, J. Chem. Phys. {\bf 88}, 5874
2050: (1988).
2051:
2052: \bibitem{takahashi-jr-38-699-1994}
2053: Y. Takahashi, N. Kurashima, I. Noda, and M. Doi, J. Rheology {\bf 38}, 699
2054: (1994).
2055:
2056: \bibitem{chan-pra-43-1826-1991}
2057: C.~K. Chan, F. Perrot, and D. Beysens, Phys. Rev. A {\bf 43}, 1826 (1991).
2058:
2059: \bibitem{krall-prl-69-1963-1992}
2060: A.~H. Krall, J.~V. Sengers, and K. Hamano, Phys. Rev. Lett. {\bf 69}, 1963
2061: (1992).
2062:
2063: \bibitem{hashimoto-prl-74-126-1995}
2064: T. Hashimoto, K. Matsuzaka, E. Moses, and A. Onuki, Phys. Rev. Lett. {\bf 74},
2065: 126 (1995).
2066:
2067: \bibitem{beysens-pra-28-2491-1983}
2068: D. Beysens, M. Gbadamassi, and B. Moncefbouanz, Phys. Rev. A {\bf 28}, 2491
2069: (1983).
2070:
2071: \bibitem{chan-prl-61-412-1988}
2072: C.~K. Chan, F. Perrot, and D. Beysens, Phys. Rev. Lett. {\bf 61}, 412 (1988).
2073:
2074: \bibitem{matsuzaka-prl-80-5441-1998}
2075: K. Matsuzaka, T. Koga, and T. Hashimoto, Phys. Rev. Lett. {\bf 80}, 5441
2076: (1998).
2077:
2078: \bibitem{hobbie-pre-54-R5909-1996}
2079: E.~K. Hobbie, S.~H. Kim, and C.~C. Han, Phys. Rev. E {\bf 54}, R5909 (1996).
2080:
2081: \bibitem{qiu-pre-58-R1230-1998}
2082: F. Qiu, J.~D. Ding, and Y.~L. Yang, Phys. Rev. E {\bf 58}, R1230 (1998).
2083:
2084: \bibitem{lauger-prl-75-3576-1995}
2085: J. Lauger, C. Laubner, and W. Gronski, Phys. Rev. Lett. {\bf 75}, 3576
2086: (1995).
2087:
2088: \bibitem{shou-pre-61-R2200-2000}
2089: Z.~Y. Shou and A. Chakrabarti, Phys. Rev. E {\bf 61}, R2200 (2000).
2090:
2091: \bibitem{zhang-jcp-113-8348-2000}
2092: Z.~L. Zhang, H.~D. Zhang, and Y.~L. Yang, J. Chem. Phys. {\bf 113}, 8348
2093: (2000).
2094:
2095: \bibitem{yamamoto-pre-59-3223-1999}
2096: R. Yamamoto and X.~C. Zeng, Phys. Rev. E {\bf 59}, 3223 (1999).
2097:
2098: \bibitem{corberi-prl-81-3852-1998}
2099: F. Corberi, G. Gonnella, and A. Lamura, Phys. Rev. Lett. {\bf 81}, 3852
2100: (1998).
2101:
2102: \bibitem{corberi-prl-83-4057-1999}
2103: F. Corberi, G. Gonnella, and A. Lamura, Phys. Rev. Lett. {\bf 83}, 4057
2104: (1999).
2105:
2106: \bibitem{corberi-pre-61-6621-2000}
2107: F. Corberi, G. Gonnella, and A. Lamura, Phys. Rev. E {\bf 61}, 6621 (2000).
2108:
2109: \bibitem{corberi-pre-62-8064-2000}
2110: F. Corberi, G. Gonnella, and A. Lamura, Phys. Rev. E {\bf 62}, 8064 (2000).
2111:
2112: \bibitem{rothman-prl-65-3305-1990}
2113: D.~H. Rothman, Phys. Rev. Lett. {\bf 65}, 3305 (1990).
2114:
2115: \bibitem{rothman-el-14-337-1991}
2116: D.~H. Rothman, Europhysics Lett. {\bf 14}, 337 (1991).
2117:
2118: \bibitem{chan-el-11-13-1990}
2119: C.~K. Chan and L. Lin, Europhysics Lett. {\bf 11}, 13 (1990).
2120:
2121: \bibitem{qiu-jcp-108-9529-1998}
2122: F. Qiu, H.~D. Zhang, and Y.~L. Yang, J. Chem. Phys. {\bf 108}, 9529 (1998).
2123:
2124: \bibitem{padilla-jcp-106-2342-1997}
2125: P. Padilla and S. Toxvaerd, J. Chem. Phys. {\bf 106}, 2342 (1997).
2126:
2127: \bibitem{ohta-jcp-93-2664-1990}
2128: T. Ohta, H. Nozaki, and M. Doi, J. Chem. Phys. {\bf 93}, 2664 (1990).
2129:
2130: \bibitem{wagner-pre-59-4366-1999}
2131: A.~J. Wagner and J.~M. Yeomans, Phys. Rev. E {\bf 59}, 4366 (1999).
2132:
2133: \bibitem{lamura-pamia-294-295-2001}
2134: A. Lamura and G. Gonnella, Physica A-statistical Mechanics Its Applications
2135: {\bf 294}, 295 (2001).
2136:
2137: \bibitem{lamura-epjb-24-251-2001}
2138: A. Lamura, G. Gonnella, and F. Corberi, European Phys. J. B {\bf 24}, 251
2139: (2001).
2140:
2141: \bibitem{berthier-pre-6305--2001}
2142: L. Berthier, L.~F. Cugliandolo, and J.~L. Iguain, Phys. Rev. E {\bf 6305},
2143: (2001).
2144:
2145: \bibitem{onuki-jpm-9-6119-1997}
2146: A. Onuki, J. Physics-condensed Matter {\bf 9}, 6119 (1997).
2147:
2148: \bibitem{bray-ptrslsapes-361-781-2003}
2149: A.~J. Bray, Philosophical Transactions Royal Soc. London Series A-mathematical
2150: Phys. Engineering Sciences {\bf 361}, 781 (2003).
2151:
2152: \bibitem{cavagna-pre-62-4702-2000}
2153: A. Cavagna, A.~J. Bray, and R.~D.~M. Travasso, Phys. Rev. E {\bf 62}, 4702
2154: (2000).
2155:
2156: \bibitem{bray-jpag-33-L305-2000}
2157: A.~J. Bray and A. Cavagna, J. Phys. A-mathematical General {\bf 33}, L305
2158: (2000).
2159:
2160: \bibitem{Onuk97}
2161: A. Onuki, J. Phys. Cond. Matt. {\bf 9}, 6119 (1997).
2162:
2163: \bibitem{doi-jcp-95-1242-1991}
2164: M. Doi and T. Ohta, J. Chem. Phys. {\bf 95}, 1242 (1991).
2165:
2166: \bibitem{stansell-prl-96--2006}
2167: P. Stansell, K. Stratford, J.~C. Desplat, R. Adhikari, and M.~E. Cates, Phys.
2168: Rev. Lett. {\bf 96}, (2006).
2169:
2170: \bibitem{cates-3d-2007}
2171: K. Stratford, J.-C. Desplat, P. Stansell, and M.~E. Cates, Phys. Rev. E {\bf
2172: 76}, 030501 (2007).
2173:
2174: \bibitem{onuki-jpsj-66-1836-1997}
2175: A. Onuki, J. Phys. Soc. Japan {\bf 66}, 1836 (1997).
2176:
2177: \bibitem{wagner-prl-80-1429-1998}
2178: A.~J. Wagner and J.~M. Yeomans, Phys. Rev. Lett. {\bf 80}, 1429 (1998).
2179:
2180: \bibitem{kendon-jfm-440-147-2001}
2181: V.~M. Kendon, M.~E. Cates, I. Pagonabarraga, J.~C. Desplat, and P. Bladon, J.
2182: Fluid Mechanics {\bf 440}, 147 (2001).
2183:
2184: \bibitem{swift-pre-54-5041-1996}
2185: M.~R. Swift, E. Orlandini, W.~R. Osborn, and J.~M. Yeomans, Phys. Rev. E {\bf
2186: 54}, 5041 (1996).
2187:
2188: \bibitem{siggia}
2189: E.~D. Siggia, Phys.~Rev. {\bf A20}, 595 (1979).
2190:
2191: \bibitem{furukawa-aip-34-703-1985}
2192: H. Furukawa, Adv. In Phys. {\bf 34}, 703 (1985).
2193:
2194: \bibitem{kendon-prl-83-576-1999}
2195: V.~M. Kendon, J.~C. Desplat, P. Bladon, and M.~E. Cates, Phys. Rev. Lett. {\bf
2196: 83}, 576 (1999).
2197:
2198: \bibitem{pagonabarraga-jsp-107-39-2002}
2199: I. Pagonabarraga, A.~J. Wagner, and M.~E. Cates, J. Statistical Phys. {\bf
2200: 107}, 39 (2002).
2201:
2202: \bibitem{mixedLB}
2203: D. Marenduzzo, E. Orlandini, M.~E. Cates and J.~M.
2204: Yeomans, preprint arXiv:0708.2062 (2007).
2205:
2206: \bibitem{Pozrikidis}
2207: C. Pozrikidis, {\em Introduction to Theoretical and Computation Fluid Dynamics}
2208: (Oxford University Press, New York, 1997).
2209:
2210: \bibitem{PreTeuVetFla92}
2211: W.~H. Press, S.~A. Teukolsky, W.~T. Vetterling, and B.~P. Flannery, {\em
2212: Numerical Recipes in C (2nd ed.)} (Cambridge University Press, Cambridge,
2213: 1992).
2214:
2215: \bibitem{NAG}
2216: Numerical Algorithms Group Ltd., Wilkinson House, Jordan Hill Road, Oxford, OX2
2217: 8DR, UK.
2218:
2219: \end{thebibliography}
2220:
2221:
2222: \end{document}
2223: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2224:
2225: