cond-mat0703151/p3.tex
1: %% p2b.tex resubmitted in response to round 1 of refs comments Sept 2006
2: %% p2.tex is version submitted to PRL 26th June 2006
3: %% p3.tex is version submitted to PRE March 2007
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \documentclass[twocolumn,showpacs,amsmath,amssymb,bibnotes]{revtex4}
6: %\documentclass[preprint,showpacs,amsmath,amssymb,bibnotes]{revtex4}
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: \usepackage{bm}
9: \usepackage{graphicx,epsfig,subfigure,afterpage}
10: \usepackage{amsfonts}
11: %\usepackage{mcite}
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: 
14: \newcommand{\be}{\begin{equation}}
15: \newcommand{\ee}{\end{equation}}
16: \newcommand{\bea}{\begin{eqnarray}}
17: \newcommand{\eea}{\end{eqnarray}}
18: \newcommand{\gdot}{\dot{\gamma}}
19: \newcommand{\gdotbar}{\overline{\dot{\gamma}}}
20: \newcommand{\ie}{{\it i.e.\/}}
21: \newcommand{\eg}{{\it e.g.\/}}
22: \newcommand{\etc}{{\it etc.\/}}
23: \newcommand{\versus}{{\it vs.\/}}
24: \newcommand{\etal}{{\it et al.\/}}
25: \newcommand{\bw}{\begin{widetext}}
26: \newcommand{\ew}{\end{widetext}}
27: \newcommand{\lae}{\stackrel{<}{\scriptstyle\sim}}
28: \newcommand{\gae}{\stackrel{>}{\scriptstyle\sim}}
29: 
30: \newcommand{\tmax}{t_{\rm max}}
31: \newcommand{\Tband}{T_{\rm b}}
32: \newcommand{\Nbase}{N_{\rm base}}
33: 
34: \newcommand{\ommax}{\omega_{\rm max}}
35: 
36: \newcommand{\xhat}{\vecv{\hat{x}}}
37: \newcommand{\yhat}{\vecv{\hat{y}}}
38: \newcommand{\zhat}{\vecv{\hat{z}}}
39: 
40: \newcommand{\vecv}[1]{\mathbf{{#1}}}
41: \newcommand{\tens}[1]{\mathbf{{#1}}}
42: \newcommand{\nablu}{{\bf \nabla}}
43: 
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: \begin{document}
46: 
47: \title{Vorticity structuring and Taylor-like velocity rolls triggered by gradient shear bands}
48: \author{Suzanne M. Fielding}
49: \email{suzanne.fielding@manchester.ac.uk}
50: \affiliation{School of Mathematics, University of Manchester, Booth
51:   Street East, Manchester M13 9EP, United Kingdom }
52: 
53: \date{\today}
54: \begin{abstract}
55: 
56: We suggest a novel mechanism by which vorticity structuring and
57: Taylor-like velocity rolls can form in complex fluids, triggered by
58: the linear instability of one dimensional gradient shear banded
59: flow. We support this with a numerical study of the diffusive Johnson-Segalman
60: model. In the steady vorticity structured state, the thickness of the
61: interface between the bands remains finite in the limit of zero stress
62: diffusivity, presenting a possible challenge to the accepted theory of
63: shear banding.
64: 
65: 
66: \end{abstract}
67: \pacs{{47.50.+d}, %{ Non-Newtonian fluid flows}--
68:      {47.20.-k}, %{ Hydrodynamic stability}--
69:      {36.20.-r}.%{ Macromolecules and polymer molecules}
70:      } 
71: \maketitle
72: 
73: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74: 
75: \section{Introduction}
76: \label{sec:intro}
77: 
78: Many complex fluids exhibit flow instabilities that result in
79: spatially heterogeneous, shear banded states.  Examples include
80: wormlike~\cite{BritCall97c,berret-pre-55-1668-1997,BecManCol04} and
81: onion~\cite{diat93,wilkinsOlms2006,SalManCol03} surfactants;
82: side-chain liquid crystalline polymers~\cite{pujolle01}; viral
83: suspensions~\cite{lettinga-jpm-16-S3929-2004,dhont-fd-123-157-2003};
84: telechelic polymers~\cite{berret-prl-8704--2001}; soft
85: glasses~\cite{CRBMGHJL02}; polymer solutions~\cite{HilVla02}; and
86: colloidal suspensions~\cite{ChenZAHSBG92}. In many cases, the
87: instability is explained by a region of negative slope
88: $dT_{xy}/d\gdot<0$ in the constitutive relation $T_{xy}(\gdot)$
89: between shear stress and shear rate for homogeneous flow, as shown in
90: Fig~\ref{fig:schematic}a. In this regime, homogeneous flow is unstable
91: ~\cite{Yerushalmi70} with respect to the formation of bands of
92: differing shear rates $\gdot_1$ and $\gdot_2$, with layer-normals in
93: the flow-gradient direction.  Force balance requires that the shear
94: stress $T_{xy}$ is uniform across the gap, and therefore common to
95: both the bands. Any change in the overall applied shear rate
96: $\gdotbar$ causes a change in the relative volume fraction $f$ of the
97: bands according to a lever rule $\gdotbar=f\gdot_1+(1-f)\gdot_2$,
98: while $\gdot_1,\gdot_2$ and $T_{xy}$ remain constant. In bulk
99: rheometry, this leads to a plateau in the steady state flow curve at
100: some stress $T_{xy}=T^*$ (Fig.~\ref{fig:schematic}a), the value of
101: which is selected by accounting for spatial non-locality in the
102: constitutive dynamics of the viscoelastic
103: stress~\cite{lu-prl-84-642-2000}.  In what follows, we shall refer to
104: the effect just described as gradient shear banding, or simply
105: gradient banding.
106: 
107: In some systems, flow induced heterogeneity has been reported in the
108: vorticity direction. By analogy with the above discussion, this effect
109: is often termed vorticity banding. It has been observed in onion
110: surfactants~\cite{diat93,Bonn+98} and colloidal
111: crystals~\cite{ChenZAHSBG92}, accompanied by a steep stress ``cliff''
112: in the flow curve. Multiple turbid and clear vorticity bands also
113: occur in some polymeric~\cite{HilVla02} and
114: micellar~\cite{fischer-ra-41-35-2002,fischer-ra-39-234-2000}
115: solutions, accompanied by shear thickening. In these cases, the bands
116: not only alternate in space but also oscillate in time. In shear
117: thinning viral suspensions, multiple stationary vorticity bands can
118: arise in the regime of isotropic-nematic microphase
119: separation~\cite{lettinga-jpm-16-S3929-2004,dhont-fd-123-157-2003}.
120: 
121: In comparison with gradient banding, vorticity banding is poorly
122: understood theoretically.  To date, the main attempts to model it have
123: invoked the analogy with gradient
124: banding~\cite{goveas-epje-6-79-2001}. As discussed above, gradient
125: bands have different shear rates $\gdot_1,\gdot_2$, and coexist at a
126: common shear stress $T^*$. For vorticity bands, the moving rotor
127: imposes a shear rate that is common to each band. Pursuing the analogy
128: with gradient banding, it is natural to invoke an underlying
129: constitutive curve of the form in Fig.~\ref{fig:schematic}b, in the
130: case of shear thickening systems.  This allows bands (A and B) of
131: differing shear {\em stresses} to coexist, having layer normals in the
132: vorticity direction.  In steady state, one then expects a flow curve
133: with a steep stress cliff, consistent with the experimental
134: observations discussed above~\cite{diat93,Bonn+98,ChenZAHSBG92}. This
135: scenario can be adapted to shear thinning systems by invoking a
136: constitutive curve of the form sketched in Fig.~\ref{fig:schematic}c.
137: In fact, at the level of 1D calculations performed separately in the
138: flow-gradient and vorticity directions, such a curve can support
139: either gradient or vorticity banded
140: states~\cite{olmsted-pre-60-4397-1999}.  Which of these (if either)
141: would be selected in a full 3D calculation remains an outstanding
142: question.  Indeed, any concrete calculation of vorticity banding to
143: date has taken a simplified one-dimensional (1D) approach, allowing
144: spatial variations only in the direction of the layer normals, and
145: thereby imposing axial and radial symmetry.
146: 
147: Beyond the traditional shear banding literature, other flow
148: instabilities are well known to trigger structuring in the vorticity
149: direction, in both simple and complex fluids.  For simple liquids in
150: curved Couette flow, the unstable centripetal interplay of fluid
151: inertia with cell curvature gives rise to Taylor velocity
152: rolls~\cite{drazin} stacked in the vorticity direction. An analogous
153: inertia-free instability occurs in viscoelastic fluids, here triggered
154: by viscoelastic hoop stresses~\cite{larson-jfm-218-573-1990,Lars92c},
155: and again leading to velocity rolls stacked along the axis of the
156: Couette cylinder.  The rolls are typically observed as bandlike
157: structures, imaged by seeding the fluid with mica flakes.
158: 
159: In contrast to the 1D vorticity banding scenario discussed above, both
160: traditional and viscoelastic Taylor-Couette instabilities are
161: inherently 2D (at least), with velocity rolls comprising a circulation
162: of fluid in the flow-gradient/vorticity
163: plane~\cite{drazin,larson-jfm-218-573-1990}.  Not unexpectedly, given
164: this roll-like structure, the wavelength of the associated banding in
165: the vorticity direction is roughly set by the width of the gap in the
166: flow-gradient direction~\cite{drazin}.
167: 
168: %
169: \begin{figure}[tb]
170: \subfigure{
171:   \includegraphics[scale=0.40]{./flowCurvea0.3.eps}
172: }
173: \hspace{-0.3cm}
174: \subfigure{
175:   \includegraphics[scale=0.35]{./schematic.eps}
176: }
177: \caption{a) Homogeneous constitutive curve and steady state flow curve for 1D planar gradient banded flow in the  DJS model at $a=0.3, \eta=0.05$. Inset: schematic arrangement of the bands in curved Couette flow. b) Schematic constitutive curve giving shear thickening vorticity banding. c) A shear thinning curve allowing gradient or vorticity banding in 1D. }
178: \label{fig:schematic}
179: \vspace{-0.5cm}
180: \end{figure}
181: %
182: 
183: In view of the above discussion, it is natural to ask whether any link
184: exists between traditional (1D) vorticity banding and 2D viscoelastic
185: Taylor-Couette instabilities; or whether the two effects are entirely
186: distinct. Given that both can be accompanied by shear thickening in
187: the bulk flow curve, this is a difficult question to address
188: experimentally. Even if they are distinct in theory, it seems feasible
189: that some experimental observations that have traditionally been
190: interpreted as 1D vorticity banding in fact comprise 2D viscoelastic
191: Taylor Couette rolls. Likely candidates include those systems in which
192: the wavelength of the alternating vorticity bands is comparable to the
193: width of the cell in the flow-gradient direction, suggesting an
194: underlying roll
195: structure~\cite{HilVla02,fischer-ra-41-35-2002,fischer-ra-39-234-2000,lettinga-jpm-16-S3929-2004,dhont-fd-123-157-2003}.
196: This was recently suggested in the context of viral suspensions in
197: Ref.~\cite{kang-pre-74--2006}. In other systems, particularly those
198: showing a very marked stress cliff in flow curve, the traditional 1D
199: scenario of Fig.~\ref{fig:schematic}b or c remains more likely.
200: 
201: In this paper, we suggest a novel mechanism by which vorticity
202: structuring can emerge in complex fluids. A key feature of our
203: approach is that, to some extent, it unifies traditional 1D (gradient)
204: banding descriptions with those of 2D roll-like instabilities. The
205: basic idea is as follows. The constitutive curve of
206: Fig.~\ref{fig:schematic}a gives rise initially to 1D gradient bands,
207: via the conventional instability in the region of negative slope,
208: $dT_{xy}/d\gdot<0$. These then undergo a secondary linear
209: instability~\cite{fielding-prl-95--2005,wilson-jnfm-138-181-2006}, due to the
210: action of normal stresses across the interface between the bands. This
211: leads finally to pronounced undulations along the interface, with
212: wavevector in the vorticity direction. These are accompanied by 2D
213: Taylor-like velocity rolls stacked in the vorticity direction, and
214: undulatory vorticity stress structuring superposed on the underlying
215: gradient bands.  In contrast to the conventional
216: inertial~\cite{drazin} and viscoelastic~\cite{larson-jfm-218-573-1990}
217: Taylor mechanisms, the vorticity instability introduced here does not
218: rely on cell curvature, but occurs even in the limit of planar shear,
219: to which our calculations are confined for simplicity.
220: 
221: The results to be presented are in good agreement with recent
222: experiments in which a gradient banded solution of wormlike micelles
223: was found to be unstable with respect to interfacial undulations with
224: wavevector in the vorticity direction~\cite{lerouge-prl-96--2006}. We
225: will return to a detailed comparison with these experiments later in
226: the manuscript.  Our results might also apply to systems in which
227: shear thickening~\cite{decruppe-pre-73--2006} and/or vorticity
228: banding~\cite{fischer-ra-41-35-2002,fischer-ra-39-234-2000,HilVla02,Bonn+98}
229: is reported to set in at the right hand edge of a stress plateau in
230: the flow curve (suggestive of underlying gradient banding, as in
231: Fig.~\ref{fig:schematic}a); or in which gradient and vorticity banding
232: have actually been observed
233: concomitantly~\cite{britton-prl-78-4930-1997}.
234: 
235: The paper is structured as follows. In Sec.~\ref{sec:model} we
236: introduce the diffusive Johnson Segalman model, within which the
237: subsequent calculations are to be performed. In Sec.~\ref{sec:1D} we
238: discuss 1D calculations, confined to the flow-gradient direction.
239: These predict gradient banding for applied shear rates in the regime
240: of negative slope in the homogeneous constitutive curve. In
241: Sec.~\ref{sec:linear} we switch to two dimensions -- the
242: flow-gradient/vorticity plane -- and show an initially 1D gradient
243: banded ``base state'' to be linearly unstable with respect to
244: undulations along the interface with wavevector in the vorticity
245: direction. We then perform a full 2D nonlinear numerical study of the
246: subsequent growth and eventual saturation of these undulations.
247: Details of the numerical method are discussed in
248: Sec.~\ref{sec:numerics}, followed by presentation of the results in
249: Sec.~\ref{sec:nonlinear}. Finally we summarise our findings and
250: discuss some directions for future study.
251: 
252: \section{Model and geometry}
253: \label{sec:model}
254: 
255: The generalised Navier--Stokes equation for a viscoelastic material in
256: a Newtonian solvent of viscosity $\eta$ and density $\rho$ is
257: %
258: \begin{equation} \label{eqn:NS} \rho(\partial_t +
259:   \vecv{v}.\nablu)\vecv{v} = \nablu .(\tens{\Sigma} + 2\eta\vecv{D}
260:   -P\tens{I}),
261: \end{equation} 
262: %
263: where $\vecv{v}(\vecv{r},t)$ is the velocity field and $\tens{D}$ is
264: the symmetric part of the velocity gradient tensor, $(\nablu
265: \vecv{v})_{\alpha\beta}\equiv
266: \partial_\alpha v_\beta$. The pressure field
267: $P(\vecv{r},t)$ is determined by enforcing incompressibility,
268: %
269: \be
270: \label{eqn:incomp}
271: \vecv{\nabla}\cdot\vecv{v}=0.
272: \ee
273: %
274: The quantity $\vecv{\Sigma}(\vecv{r},t)$ in Eqn.~\ref{eqn:NS} is the
275: extra stress contributed by the viscoelastic component. In principle,
276: the dynamics of this quantity should be explicitly derived by
277: averaging over the underlying microscopic dynamics of the viscoelastic
278: component. This was done in Refs.~\cite{cates90} for wormlike
279: micelles.  For simplicity, however, we use the phenomenological
280: Johnson-Segalman (JS) model~\cite{johnson-jnfm-2-255-1977}
281: %
282: \begin{widetext}
283: \be
284: \label{eqn:DJS}
285: (\partial_t
286: +\vecv{v}\cdot\nablu )\,\tens{\Sigma} 
287: = a(\tens{D}\cdot\tens{\Sigma}+\tens{\Sigma}\cdot\tens{D}) +
288: (\tens{\Sigma}\cdot\tens{\Omega} + \tens{\Omega}\cdot\tens{\Sigma})  
289:  + 2 G\tens{D}-\frac{\tens{\Sigma}}{\tau}+ \frac{\ell^2}{\tau }\nablu^2 
290:  \tens{\Sigma}.
291: \ee
292: \end{widetext}
293: %
294: In this equation, $G$ is a plateau modulus, $\tau$ is the viscoelastic
295: relaxation time, and $\tens{\Omega}$ is the antisymmetric part of the
296: velocity gradient tensor.  For $a=1$ and $\ell=0$, Eqn.~\ref{eqn:DJS}
297: reduces to the Oldroyd B model, which can be derived by considering
298: the dynamics of an ensemble of Hookean dumbbells in solution. For
299: $|a|<1$, the JS model captures non-affine slip between the dumbbells
300: and the solvent, leading to the drastic shear thinning of
301: Fig.~\ref{fig:schematic}a. Accordingly, $a$ is called the slip
302: parameter. The JS model is the simplest tensorial model to exhibit a
303: regime of negative slope in the homogeneous constitutive curve, and so
304: to predict a shear banding instability. As discussed further below,
305: the diffusive term $\nablu^2 \tens{\Sigma}$ in Eqn.~\ref{eqn:DJS} is
306: needed to correctly describe the ultimate shear banded
307: flow~\cite{lu-prl-84-642-2000}.
308: 
309: Within this model, we study planar shear between parallel plates at
310: $y=0,L$, with the top plate driven at velocity $V \xhat$.  At the
311: plates we assume boundary conditions of
312: $\partial_y\Sigma_{\alpha\beta}=0\;\forall\;\alpha,\beta$ for the
313: viscoelastic stress, with no slip and no permeation for the fluid
314: velocity.  In the linear stability analysis of Sec.~\ref{sec:linear},
315: we consider small values of the Reynolds number $Re=\rho L^2/\eta$. In
316: the nonlinear study of Secs.~\ref{sec:numerics}
317: and~\ref{sec:nonlinear}, we set $Re=0$ at the outset.  Throughout we
318: use units in which $G=1,\tau=1$ and $L=1$. 
319: 
320: %
321: \begin{figure}[b]
322:   \centering
323: \includegraphics[scale=0.4]{./dispersion.eps}
324: \includegraphics[scale=0.42]{./peak.eps}
325:  \caption{a) Dispersion relation for perturbations about a 1D banded
326:  state for $\ell=0.00375,\, 0.00250,\, 0.00125$. $a=0.3$, $\eta=0.05$,
327:  $\rho/\eta=0.02$. Inset: Linear dynamics of 2D code, starting from a
328:  flat interface. Solid lines: weight in modes, $q_zL_z/2\pi=1,2$;
329:  dashed: analytical prediction. b) Peak $\omega^*$ ($q^*$ in inset) in
330:  dispersion relation {\it vs.} $\gdotbar$ across the plateau of
331:  Fig.~\ref{fig:schematic}a. Vertical dashed lines denote the edges of
332:  the stress plateau. }
333: \label{fig:dispersion}
334: \vspace{-0.5cm}
335: \end{figure}
336: %
337: 
338: 
339: \section{1D gradient bands}
340: \label{sec:1D}
341: 
342: As noted above, to capture shear thinning the DJS model invokes a slip
343: parameter $a$ with $|a|<1$, giving non affine deformation of the
344: viscoelastic component~\cite{johnson-jnfm-2-255-1977}. The homogeneous
345: constitutive curve $T_{xy}=\Sigma_{xy}(\gdot, a)+\eta\gdot$ is then
346: capable of non-monotonicity, as in Fig.~\ref{fig:schematic}a.  For an
347: imposed shear rate $\gdotbar\equiv V/L$ in the region of decreasing
348: stress, homogeneous flow is unstable with respect to fluctuations with
349: wavevector in the flow-gradient direction
350: $y$~\cite{yerushal.j-ces-25-1891-1970}. A 1D calculation then predicts
351: separation into gradient bands of differing shear rates $\gdot_1,
352: \gdot_2$, with a flat interface in between. The diffusive term in
353: Eqn.~\ref{eqn:DJS} is needed to account for spatial gradients of the
354: shear rate and viscoelastic stress across the interface, which has a
355: characteristic thickness $O(\ell)$. It also ensures a unique,
356: history-independent banding stress
357: $T_{xy}=T^*$~\cite{lu-prl-84-642-2000}, as seen experimentally.  We
358: expect $\ell=O(10^{-4})$, set by the typical micellar mesh size, in
359: units of the (typical) gap size.
360: 
361: %$\gdot_1=0.66,\gdot_2=7.09$, at a shear stress $T^{\ast}_{xy}=0.506$.
362: 
363: \section{Linear instability}
364: \label{sec:linear}
365: 
366: In Refs.~\cite{fielding-prl-95--2005,wilson-jnfm-138-181-2006}, we
367: considered the linear stability of this 1D gradient banded state with
368: respect to 3D ($x, y, z$) perturbations of infinitesimal amplitude. In
369: the flow direction $\xhat$ and vorticity direction $\zhat$ these are
370: decomposed into Fourier modes with wavevectors
371: $\vecv{q}=q_x\xhat+q_z\zhat$.  (Ref.~\cite{renardy} had previously
372: considered $\vecv{q}=q_x\xhat$ in the pathological limit $\ell=0$,
373: assuming ``top jumping''.)  For diffuse interfaces, $l\gae 0.015$, the
374: 1D state is linearly stable. For $l\lae 0.015$, we find it to be
375: linearly unstable with respect to modes with wavevector
376: $\vecv{q}=q_x\xhat$ in the flow
377: direction~\cite{fielding-prl-95--2005,wilson-jnfm-138-181-2006}.  The
378: associated eigenfunction essentially corresponds to undulations along
379: the interface. For $l\lae 0.005$, the 1D state is also unstable with
380: respect to undulations with wavevector $\vecv{q}=q_z\zhat$ in the
381: vorticity direction. However, these modes are predicted to grow much
382: more slowly than those with $\vecv{q}=q_x\xhat$.  Accordingly, in
383: Refs.~\cite{fielding-prl-95--2005,wilson-jnfm-138-181-2006}, we
384: focused mainly on the dominant modes, $\vecv{q}=q_x\xhat$.
385: Subsequently in Ref.~\cite{fielding-prl-96--2006}, however, we showed
386: that these undulations are cut off, once they attain finite amplitude,
387: by the nonlinear effects of shear. In contrast, the vorticity
388: direction is neutral with respect to the shear. Accordingly, the modes
389: with $\vecv{q}=q_z\vecv{\hat{z}}$ should not suffer this cutoff and
390: are therefore likely to contribute significantly to the ultimate
391: nonlinear state, despite their much slower initial growth rate. With
392: this motivation, in this paper we study the dynamics of the model in
393: the flow-gradient/vorticity $(y-z)$ plane. For simplicity and
394: computational efficiency, we will assume uniformity in the flow
395: direction $x$, returning in Sec.~\ref{sec:nonlinear} below to comment
396: on the validity of this simplification.  It corresponds to taking a
397: vertical slice through one side of an axisymmetric flow state in the
398: planar limit of a Couette device.
399: 
400: The growth rates $\omega$ of the vorticity modes
401: $\vecv{q}=q_z\vecv{\hat{z}}$ are shown in Fig.~\ref{fig:dispersion}a,
402: at a single value of the imposed shear rate. States with thinner
403: interfaces (smaller $\ell$) are more unstable (larger $\omega>0$).
404: Fig.~\ref{fig:dispersion}b shows the growth rate of the maximally
405: unstable mode for shear rates across the stress plateau of
406: Fig.~\ref{fig:schematic}a. The corresponding wavelength
407: $\lambda^*=O(1)$ is of order the rheometer gap $L\equiv 1$ (inset). At
408: small $\ell$, the instability persists across most of the plateau, so
409: is likely to be unavoidable experimentally.  The mechanism of
410: instability is not fully understood, but is likely to stem from steep
411: gradients in the normal stress and shear rate across the
412: interface~\cite{hinch-jnfm-43-311-1992,mcleish87,fielding-prl-95--2005}.
413: 
414: \section{Numerical method}
415: \label{sec:numerics}
416: 
417: To study the undulations once they have grown to attain a finite
418: amplitude, beyond the regime of linear instability, we solve the
419: model's full nonlinear dynamics numerically.  In this section we
420: discuss the details of our numerical method, which is adapted from
421: that of Refs.~\cite{chilcott-jnfm-29-381-1988,fielding-prl-96--2006}.
422: Readers who are not interested in these issues can skip straight to
423: Sec.~\ref{sec:nonlinear} without loss of thread.
424: 
425: The model equations have already been specified in
426: Sec.~\ref{sec:model}, together with the flow geometry, boundary
427: conditions and choice of adimensionalisation. For computational
428: efficiency, our numerical study is confined to the $(y-z)$ plane,
429: assuming translational invariance along the flow direction $x$. In the
430: vorticity direction we take a cell of length $L_z$, with periodic
431: boundary conditions.
432: 
433: We consider the limit of zero Reynolds number, in which
434: Eqn.~\ref{eqn:NS} reduces to
435: %
436: \be
437: \label{eqn:zeroRey}
438: 0 = \nablu .(\tens{\Sigma} + 2\eta\vecv{D} -P\tens{I}).
439: \ee
440: %
441: To ensure that the incompressibility constraint of
442: Eqn.~\ref{eqn:incomp} is satisfied always, we express the velocity in terms
443: of stream functions $\phi$ and $\psi$:
444: %
445: \be
446: \label{eqn:stream}
447: v_x=\partial_y\phi,\;\; v_y=\partial_z\psi,\;\; v_z=-\partial_y\psi.
448: \ee
449: %
450: In this way, Eqn.~\ref{eqn:incomp} need no longer be considered: it
451: remains only solve Eqns.~\ref{eqn:DJS} and~\ref{eqn:zeroRey}, with the
452: velocity expressed as in Eqn.~\ref{eqn:stream}.
453: 
454: 
455: To solve these, the basic strategy is to step along a grid of time
456: values $t^n=n\Delta t$ for $n=1,2,3\cdots$, at each step updating
457: $\tens{\Sigma}^n, \phi^n, \psi^n\to \tens{\Sigma}^{n+1}, \phi^{n+1},
458: \psi^{n+1}$.  Discretization with respect to time of any quantity $f$
459: is denoted $f(t^n)=f^n$, or sometimes below by $f|^n$. At each
460: time-step, we first update the viscoelastic stress $\tens{\Sigma}^n \to
461: \tens{\Sigma}^{n+1}$ using the constitutive equation (\ref{eqn:DJS})
462: with fixed, old values of the stream-functions $\phi^n,\psi^n$. We
463: then update $\phi^n,\psi^n \to \phi^{n+1},\psi^{n+1}$ using the force
464: balance equation (\ref{eqn:zeroRey}) with the new values of
465: $\tens{\Sigma}^{n+1}$.
466: 
467: The update $\tens{\Sigma}^n\to\tens{\Sigma}^{n+1}$ using the
468: viscoelastic constitutive equation (\ref{eqn:DJS}) is performed as
469: follows. As a preliminary step, we rewrite (\ref{eqn:DJS}) in the form
470: %
471: \be
472: \label{eqn:DJSsep}
473: \partial_t\tens{\Sigma}= 
474:  f(\nablu \vecv{v},\tens{\Sigma}) - \vecv{v}\cdot\nablu \tens{\Sigma}
475:  + \ell^2 \nablu^2 \tens{\Sigma}, 
476: \ee
477: %
478: in which $f(\nablu \vecv{v},\tens{\Sigma})$ comprises the
479: non-diffusive terms from the right hand side of Eqn.~\ref{eqn:DJS}.
480: In what follows, the three terms on the right hand side of
481: Eqn.~\ref{eqn:DJSsep} are referred to as the local, advective and
482: diffusive terms respectively. Numerically, they are dealt with in
483: three successive partial updates $\tens{\Sigma}^n \to
484: \tens{\Sigma}^{n+1/3}$, $\tens{\Sigma}^{n+1/3} \to
485: \tens{\Sigma}^{n+2/3}$ and $\tens{\Sigma}^{n+2/3} \to
486: \tens{\Sigma}^{n+1}$.
487: 
488: In the first of these, the local term is handled using an explicit
489: Euler algorithm~\cite{PreTeuVetFla92}, checked for consistency against
490: a fourth order Runge-Kutta algorithm~\cite{PreTeuVetFla92}.
491: Temporarily setting aside the issue of spatial discretization, the
492: Euler algorithm can be written
493: %
494: \be
495: \label{eqn:local}
496: \tens{\Sigma}^{n+1/3}(y,z)=\tens{\Sigma}^n + \Delta t \; f(\nablu
497: \vecv{v}^n,\tens{\Sigma}^n).
498: \ee
499: %
500: In terms of the stream-functions $\phi$ and $\psi$, the
501: velocity-gradient tensor $\nablu \vecv{v}$ has Cartesian components
502: %
503: \be
504: \label{eqn:localM}
505: \nablu \vecv{v}= \left(\begin{array}{ccc}
506:             0 & 0 & 0 \\
507:             \partial_y^2\phi & \partial_y\partial_z\psi &
508:             -\partial_y^2\psi \\
509:             \partial_y\partial_z\phi & \partial_z^2\psi &
510:             -\partial_y\partial_z\psi \end{array}\right),
511: \ee
512: %
513: in which we have omitted the superscripts $n$ for clarity.
514: Eqns.~\ref{eqn:local} and~\ref{eqn:localM} are then spatially
515: discretized on a rectangular grid in real space. In some runs of the
516: code, the grid points are linearly spaced, with $z_i=i\Delta z$ for
517: $i=1\cdots N_z$ and $y_j=j\Delta y$ for $j=1\cdots N_y$.  In others,
518: we used a nonlinear mapping in the $y$ direction to focus attention on
519: the region explored by the interface. For simplicity, most of the
520: description of this section will concern the linear grid, though we
521: will return briefly at the end of the section to discuss the nonlinear
522: modification.  In either case, any spatially discretized function $f$
523: is denoted $f(y^j,z^i)=f_{ij}$, or sometimes $f|_{ij}$.  (The
524: apparently unusual order of the indices is a historical convention on
525: the part of the author, stemming from a previous study in the $x-y$
526: plane.)  Eqn.~\ref{eqn:local} then becomes
527: %
528: \begin{widetext}
529: %
530: \be
531: \label{eqn:localD}
532: \tens{\Sigma}^{n+1/3}_{ij}=\tens{\Sigma}^n_{ij} + \Delta t \; f(\nablu
533: \vecv{v}^n_{ij},\tens{\Sigma}^n_{ij}),
534: \ee
535: %
536: The derivatives in the components of $\nablu\vecv{v}^n_{ij}$ are
537: discretized (in the case of a rectangular grid) as follows:
538: %
539: \be
540: \label{eqn:p2y}
541: \partial_y^2\psi|^n_{ij}=\frac{1}{ \Delta
542:   y^2}\left[\psi_{i(j+1)}^n-2\psi_{ij}^n+\psi_{i(j-1)}^n\right],
543: \ee
544: %
545: \be
546: \partial_z^2\psi|^n_{ij}=\frac{1}{ \Delta
547:   z^2}\left[\psi_{(i+1)j}^n-2\psi_{ij}^n+\psi_{(i-1)j^n}\right],
548: \ee
549: %
550: and
551: %
552: \be
553: \partial_y\partial_z \psi|^n_{ij}=\frac{1}{4\Delta x\Delta
554:   y}\left[\psi_{(i+1)(j+1)}-\psi_{(i+1)(j-1)}-\psi_{(i-1)(j+1)}+\psi_{(i-1)(j-1)}\right].
555: \ee
556: %
557: Corresponding derivatives of $\phi$ are obtained in the same way,
558: replacing $\psi$ by $\phi$ in the above equations. For values of $ij$
559: at the edges of the flow domain, these formulae link to values of the
560: flow variables at ``phantom'' grid points that lie just outside the
561: domain.  These values are specified by imposing the boundary
562: conditions, the spatial discretization of which is discussed at the
563: end of this section.
564: 
565: The advective term is also handled using an explicit Euler
566: algorithm~\cite{PreTeuVetFla92}, on the same real space grid:
567: %
568: \bea
569: \label{eqn:advect}
570: \tens{\Sigma}^{n+2/3}_{ij}&=&\tens{\Sigma}^{n+1/3}_{ij} - \Delta t
571: \left(
572: v_{y}|_{ij}^n \, \partial_y\tens{\Sigma}|^n_{ij}+v_{z}|_{ij}^n\,\partial_z\tens{\Sigma}|^n_{ij}\right)\nonumber\\
573:   &=&\tens{\Sigma}^{n+1/3}_{ij} - \Delta t
574: \left(
575: \partial_z\psi |_{ij}^n \, \partial_y\tens{\Sigma}|^n_{ij}-\partial_y\psi|_{ij}^n\,\partial_z\tens{\Sigma}|^n_{ij}\right).
576: \eea
577: %
578: The derivatives of $\psi$ in this equation are discretized as follows:
579: %
580: \be
581: \label{eqn:p1y}
582: \partial_y\psi|_{ij}^n=\frac{1}{2\Delta
583:   y}\left[\psi_{i(j+1)}-\psi_{i(j-1)}\right]\;\;\;\textrm{and}\;\;\; \partial_z\psi|_{ij}^n=\frac{1}{2\Delta
584:   z}\left[\psi_{(i+1)j}-\psi_{(i-1)j}\right].
585: \ee
586: %
587: %
588: The derivative of $\tens{\Sigma}$ with respect to $y$ in
589: Eqn.~\ref{eqn:advect} was discretized using third-order
590: upwinding~\cite{Pozrikidis}:
591: %
592: \be
593: \partial_y\tens{\Sigma}_{ij}^n=\frac{1}{6\Delta
594:   y}\left[\tens{\Sigma}^n_{i(j-2)} -6 \tens{\Sigma}^n_{i(j-1)} +3
595:   \tens{\Sigma}^n_{ij} +2 \tens{\Sigma}^n_{i(j+1)}
596: \right]\;\;\;\textrm{if}\;\;\; v_y|^n_{ij}>0,
597: \ee
598: %
599: while
600: %
601: \be
602: \partial_y\tens{\Sigma}_{ij}^n=\frac{1}{6\Delta
603:   y}\left[-\tens{\Sigma}^n_{i(j+2)} +6 \tens{\Sigma}^n_{i(j+1)} -3
604:   \tens{\Sigma}^n_{ij} -2 \tens{\Sigma}^n_{i(j-1)}
605: \right]\;\;\;\textrm{if}\;\;\; v_y|^n_{ij}<0,
606: \ee
607: %
608: with analogous expressions for the derivative of $\tens{\Sigma}$ with
609: respect to $z$.
610: 
611: The diffusive term is handled by discretizing on $y$ in real space as
612: above, taking a fast Fourier transform $z\to q_i$ in the vorticity
613: dimension using a standard NAG routine~\cite{NAG}, and solving the
614: resulting problem using a semi-implicit Crank-Nicolson
615: algorithm~\cite{PreTeuVetFla92}:
616: %
617: \be
618: \label{eqn:Crank}
619: \tens{\Sigma}^{n+1}_{ij}-\tens{\Sigma}^{n+2/3}_{ij} =\frac{1}{2} l^2
620: \Delta t
621: \left(\partial_y^2\tens{\Sigma}_{ij}^{n+2/3}-q_i^2\tens{\Sigma}_{ij}^{n+2/3}\right)+\frac{1}{2}
622: l^2 \Delta t
623: \left(\partial_y^2\tens{\Sigma}_{ij}^{n+1}-q_i^2\tens{\Sigma}_{ij}^{n+1}\right),
624: \ee
625: %
626: in which the index $i$ now labels the Fourier mode number. The
627: derivatives $\partial_y^2$ are discretized as in Eqn.~\ref{eqn:p2y}
628: above. Note that Eqn.~\ref{eqn:Crank} contains no mixing of the
629: Cartesian components $\Sigma_{\alpha\beta}$ for any
630: $\alpha\beta=xx,xy,yy\cdots$, so can be solved for each one
631: separately. Bringing all terms in the unknown
632: $\tens{\Sigma}^{n+1}_{ij}$ across to the left hand side, and putting
633: all terms in the known $\tens{\Sigma}^{n+2/3}_{ij}$ on the right hand
634: side, we obtain a sparse set of linear equations characterized by a
635: tridiagonal matrix on the left hand side. These are then solved for
636: the $\tens{\Sigma}^{n+1}_{ij}$ using standard NAG routines~\cite{NAG}.
637: %
638: 
639: Having updated $\tens{\Sigma}^n \to \tens{\Sigma}^{n+1}$ using the
640: viscoelastic constitutive equation, we now update the stream-functions
641: $\phi^n,\psi^n \to \phi^{n+1},\psi^{n+1}$ using the $x,y$ and $z$
642: components of the force balance equation (\ref{eqn:zeroRey}). Again,
643: we work in real flow-gradient space and reciprocal vorticity space. To
644: eliminate the pressure from Eqn.~\ref{eqn:zeroRey}, we subtract
645: $\partial_y$ of the $z$ component from $\partial_z$ of the $y$
646: component to get the following equations, written separately for
647: $q_i=0$ and $q_i\neq 0$:
648: %
649: \be
650: \label{eqn:fb1}
651: \partial_y^3\phi|^{n+1}_{0(j+1/2)}=-\frac{1}{\eta}\partial_y\Sigma_{xy}|_{0(j+1/2)}^{n+1}\;\;\;\textrm{for}\;\;\; q_i=0,
652: \ee
653: %
654: \be
655: \label{eqn:fb2}
656: \partial^3_y\phi|^{n+1}_{i(j+1/2)}-q_i^2 \partial_y\phi|^{n+1}_{i(j+1/2)}=-\frac{1}{\eta}\left(\partial_y\Sigma_{xy}|^{n+1}_{i(j+1/2)}  + iq_i \Sigma_{xz}|^{n+1}_{i(j+1/2)}\right)\;\;\;\textrm{for}\;\;\; q_i\neq 0,
657: \ee
658: %
659: \be
660: \label{eqn:fb3}
661: \partial_y^3\psi|^{n+1}_{0(j+1/2)}=\frac{1}{\eta}\partial_y\Sigma_{yz}|^{n+1}_{0(j+1/2)}\;\;\;\textrm{for}\;\;\;q_i=0,
662: \ee
663: %
664: \be
665: \label{eqn:fb4}
666: \nablu^4\psi|^{n+1}_{ij}=-\frac{1}{\eta}\left[iq_i\partial_y\left(\Sigma_{yy}-\Sigma_{zz}\right)|^{n+1}_{ij}-\left(q_i^2+\partial_y^2\right)\Sigma_{yz}|^{n+1}_{ij}\right]\;\;\;\textrm{for}\;\;\;
667: q_i\neq
668: 0,
669: \ee
670: %
671: with $\nablu^2=(\partial_y^2-q_i^2)$. The real and imaginary parts of
672: these equations are treated separately. The third order equations
673: (\ref{eqn:fb1}), (\ref{eqn:fb2}) and (\ref{eqn:fb3}) are discretized
674: at staggered half grid points $y_{j+1/2}$ for $j=1\cdots N_y-1$, with
675: derivatives calculated as follows:
676: %
677: \be
678: \partial_y f|_{j+1/2}=\frac{1}{\Delta y}\left(f_{j+1}-f_j\right),
679: \ee
680: %
681: and 
682: %
683: \be
684: \partial_y^3 f|_{j+1/2}=\frac{1}{\Delta
685:   y^3}\left(f_{j+2}-3f_{j+1}+3f_{j}-f_{j-1}\right),
686: \ee
687: %
688: for any quantity $f$. (For clarity, the subscript $i$ and the
689: superscript $n$ have been omitted from these expressions.)  The fourth
690: order equation is implemented at full grid points $y_j$, excluding
691: those at the very edge of the domain ($y_0$ and $y_N$). In it,
692: $\partial^2_y$ and $\partial_y$ are discretized as in
693: Eqns.~\ref{eqn:p2y} and~\ref{eqn:p1y} respectively, and $\partial_y^4$
694: according to
695: %
696: \be
697: \partial_y^4 f|_{j}=\frac{1}{\Delta y^4}\left(f_{j+2}-4f_{j+1}+6f_j
698:   -4f_{j-1}+f_{j-2}\right).
699: \ee
700: %
701: Each of Eqns.~\ref{eqn:fb1} to~\ref{eqn:fb4}, for each mode index $i$,
702: then takes the form of a sparse set of linear equations for
703: $\phi_{ij}^{n+1}$ or $\psi_{ij}^{n+1}$. These equations are solved
704: using standard NAG routines~\cite{NAG}.
705: 
706: \end{widetext}
707: 
708: It remains finally to specify the spatial discretization of the
709: boundary conditions. In turn, this will prescribe the values of the
710: flow variables on the phantom grid points that lie just outside the
711: flow domain.
712: 
713: In the vorticity direction $z$, the boundary conditions are periodic.
714: For any quantity $f$ on the grid $z_1\cdots z_{N_Z}$ in real space
715: $z_i$, we thus have $f_{-1}=f_{{N_z}-1}$, $f_{0}=f_{{N_Z}}$, $f_{{N_z}+1}=f_0$,
716: $f_{{N_z}+2}=f_1$. In reciprocal space $q_i$, the periodic boundary conditions
717: are always satisfied.
718: 
719: In the flow-gradient direction $y$, the boundary conditions for the
720: fluid velocity at the plates $y=0,1$ are those of no slip and no
721: permeation. In terms of the stream-function $\phi$, the no slip
722: condition gives $\partial_y\phi=0$ at $y=0$ and
723: $\partial_y\phi=\gdotbar$ at $y=1$. We also note that $\phi$ is only
724: defined up to an arbitrary additive constant, and accordingly choose
725: $\phi=0$ at $y=0$. The third order Eqns.~\ref{eqn:fb1}
726: and~\ref{eqn:fb2} then have three boundary conditions, as required.
727: After discretizing real flow-gradient space $y_j$ in reciprocal
728: vorticity space $q_i$, we then have
729: %
730: $$\phi_{i1}=0,\;\;\;\phi_{i0}=\phi_{i2},\;\;\;\textrm{and}\;\;\;\phi_{i({N_y}+1)}=\phi_{i({N_y}-1)}+\delta_{i0}\gdotbar \Delta y,$$
731: %
732: in which $\delta_{ij}$ is the usual Kronecker delta function.
733: 
734: In terms of the stream-function $\psi$, the no-slip condition gives
735: $\partial_y\psi=0$ at $y=0,1$, and the no-permeation condition gives
736: $\partial_z\psi=0$ at $y=0,1$. We also note that $\psi$ is only
737: defined up to an arbitrary additive constant, and choose $\psi=0$ at
738: $y=0$. In the $q_i=0$ equation (\ref{eqn:fb3}), we then have
739: %
740: $$\psi_{i1}=0,\;\;\;\psi_{i0}=\psi_{i2}\;\;\;\textrm{and}\;\;\;\psi_{i({N_y}+1)}=\psi_{i({N_y}-1)}.$$
741: %
742: These also hold for the $q_i\neq 0$ equation (\ref{eqn:fb4}), which
743: obeys the additional condition
744: %
745: $$\psi_{i{N_y}}=0.$$
746: %
747: The zero-gradient boundary condition for the viscoelastic stress, after
748: discretization on the flow-gradient grid $y_1,y_2\cdots y_{N_y}$, gives
749: $f_{i0}=f_{i2}$ and $f_{i({N_y}+1)}=f_{i({N_y}-1)}$ for all components
750: $f=\Sigma_{xx}, \Sigma_{xy}, \Sigma_{yy}\cdots$. These apply in both
751: real $z_i$ and reciprocal $q_i$ vorticity spaces.
752: 
753: 
754: Extension to the case of a nonlinear grid in the $y$ direction is
755: straightforward in principle, but cumbersome in detail. Discretized
756: derivatives are calculated via the usual Taylor expansions. For
757: example, to first order accuracy, centred second derivatives with
758: respect to $y$ become
759: %
760: \be
761: \partial_y^2
762: f_j=\frac{2}{y_{j+1}-y_{j-1}}\left[\frac{f_{j+1}-f_j}{y_{j+1}-y_j}-\frac{f_j-f_{j-1}}{y_j-y_{j-1}}\right].
763: \ee
764: %
765: We checked our nonlinear mapping carefully by performing a few runs
766: for identical parameter sets with both linear and highly nonlinear
767: grids.  All the numerical results in this paper are converged with
768: respect to grid and timestep, to within the accuracy resolvable on the
769: plots presented.
770: 
771: \begin{figure}[t]
772: \vspace{-0.8cm}\includegraphics[scale=0.63,
773: angle=90]{./gdot2.0_eta0.05_a0.3_l0.00375_Lz4.0_state_time506.34999999494_OPWxx_border150.ps}
774: \includegraphics[scale=0.44, angle=90]{./flowMap.eps}
775: \caption{Steady state at $a=0.3$, $\eta=0.05$, $\gdotbar=2.0$, $\ell=0.00375$, $L_z=4.0$. Left: greyscale of $\Sigma_{xx}$ in $y-z$ plane. Middle: $(y-z)$ velocity vectors, showing Taylor-like rolls. Right: vorticity banding of viscoelastic shear stress.}
776: \label{fig:steady}
777: \vspace{-0.5cm}
778: \end{figure}
779: 
780: 
781: 
782: \section{Nonlinear steady state}
783: \label{sec:nonlinear}
784: 
785: In each simulation run, we input as an initial condition the 1D
786: gradient-banded state discussed in Sec.~\ref{sec:1D}, superposed with
787: Fourier perturbations of tiny random amplitudes.  As expected, under
788: conditions where the linear analysis of Sec.~\ref{sec:linear} predicts
789: the 1D initial state to be unstable with respect to perturbations with
790: wavevectors $\vecv{q}=q_z\zhat$ in the vorticity direction, we find
791: that these initial disturbances grow in time. Full agreement between
792: (i) the early-time growth rate and functional form of the fastest
793: growing mode and (ii) the most unstable eigenvalue and eigenfunction
794: of the linear stability analysis provides a stringent check of our
795: numerical method.
796: 
797: During the instability, the initially flat interface between the bands
798: develops undulations that grow in time.  At long times, once nonlinear
799: effects become important, these saturate in a finite amplitude
800: interfacial undulation to give a 2D steady state
801: (Fig.~\ref{fig:steady}).  The wavelength of the steady undulations
802: corresponds to that of the maximally unstable mode of the linear
803: analysis. For the parameters of Fig.~\ref{fig:steady}, this is roughly
804: twice the gap width.  Associated with these undulations are
805: Taylor-like velocity rolls stacked in the vorticity direction
806: (Fig.~\ref{fig:steady}, middle), accompanied by undulations of the
807: stress along the wall (Fig.~\ref{fig:steady}, right).  The results of
808: Fig.~\ref{fig:steady} could be tested experimentally as follows.
809: Optical measurements should reveal birefringent stripes stacked in the
810: vorticity direction, each of height comparable to the gap width.
811: Likewise, the velocity rolls could potentially be measured using
812: velocimetry. This is a challenging task, however, because the highest
813: speed in Fig.~\ref{fig:steady}c is only $O(0.01)$.
814: 
815: Our results capture recent experimental observations in which an
816: initially 1D gradient banded state of a wormlike surfactant solution
817: was found to destabilise with respect to interfacial undulations with
818: wavevector in the vorticity direction~\cite{lerouge-prl-96--2006}.
819: Indeed, several features of our results can be directly compared with
820: these experiments, as follows. In the ultimate steady state, the
821: wavelength $O(L)$ and amplitude $O(L/10)$ of the undulations in our
822: Fig.~\ref{fig:steady} are comparable to those in Fig.  2 of
823: Ref.~\cite{lerouge-prl-96--2006}, measured in units of the gap width
824: $L$.  The wavelength of these undulations was furthermore reported to
825: increase with increasing average applied shear rate
826: $\gdotbar$~\cite{lerouge-prl-96--2006}, consistent with the inset of
827: our Fig.~\ref{fig:dispersion}b.
828: %
829: \begin{figure}[t]
830:   \centering
831: \includegraphics[scale=0.42]{./selected.eps}
832:  \caption{Selected stress: 1D initial state and 2D steady
833:  state. $L_z=2.0$, $a=0.3$, $\eta=0.05$, $\gdotbar=2.0$. At $\ell=0$, the
834:  stress varies erratically about an average that depends on the
835:  grid. Inset: evolution to steady state at $l=0.00375$. }
836: \label{fig:selected}
837: \vspace{-0.5cm}
838: \end{figure}
839: %
840: 
841: The kinetics of the instability can also be compared, via the temporal
842: evolution of the stress signal. In the
843: experiments~\cite{lerouge-prl-96--2006}, a shear startup protocol was
844: followed. Accordingly, the stress signal showed an initial overshoot
845: followed by a decay (at $\gdotbar=30s^{-1}$) on a timescale $O(\tau)$
846: to a plateau value. This part of the dynamics corresponded to the
847: initial formation of 1D gradient bands.  It is absent from our
848: simulations, because we take as our initial condition an already 1D
849: gradient banded flow.  Subsequently, the stress signal in
850: Ref.~\cite{lerouge-prl-96--2006} slowly increased by about $1\%$ on a
851: timescale $O(100\tau)$.  This part of the dynamics was associated with
852: the 1D gradient banded state destabilising to exhibit vorticity
853: undulations. As shown in Fig.~\ref{fig:selected}, it is captured very
854: well by our simulations: we find a slow stress increase $O(1\%)$ on a
855: time-scale $O(100\tau)$, consistent with the experiments.
856: 
857: Some differences between our work and the experiments of
858: Ref.~\cite{lerouge-prl-96--2006} are noted as follows.  In
859: Ref.~\cite{lerouge-prl-96--2006}, the instability was studied using
860: light scattering techniques, which couple to concentration
861: fluctuations. In the present manuscript, we do not consider
862: concentration coupling. In future work, it might be interesting to
863: perform analogous simulations in the concentration coupled model of
864: Ref.~\cite{fielding-epje-11-65-2003}.  However, an important finding
865: of the present work is that concentration coupling is not actually
866: needed to trigger the basic undulatory instability.  Indeed, we
867: believe this to stem instead from normal stress effects, with
868: concentration coupling a sub-dominant feature. Finally, we have not
869: seen the exotic dynamics reported at the edges of the stress plateau
870: in Ref.~\cite{lerouge-prl-96--2006}. We cannot access small enough
871: $\ell$ for the instability to persist here.  (Recall
872: Fig.~\ref{fig:dispersion}b.)
873: 
874: \begin{figure}[t]
875: \subfigure{
876:   \includegraphics[scale=0.47]{./interface2.eps}
877: }
878: \caption{Profile $\Sigma_{xx}$ normal to the interface; $\ell=0.00375$ (thick lines), $\ell=0.00250, 0.00125$ (thin). $a=0.3$, $\eta=0.05$, $\gdotbar=2.0$, $L_z=2.0$. a,b) 1D state with interfacial thickness $d\propto \ell$. (c,d), (e,f) 2D  state at the $z-$coordinate where the interface has maximal $+$ and $-$ $y$ displacements. The  thickness $d$ appears {\em independent} of $\ell$, but note the  bump of thickness $O(l)$. The offset $y_0(\ell)$ is chosen to centre the profile at the origin in each case.}
879: % results files as follows
880: % gdot2.0_eta0.05_a0.3_l0.00375_Lz2.0_temperature1.0e-4_Dt0.001_Nz200_Nin1000_Nout100_adapt1_tmax800.0 
881: % gdot2.0_eta0.05_a0.3_l0.0025_Lz2.0_temperature1.0e-4_Dt0.001_Nz400_Nin1600_Nout160_adapt1_tmax800.0
882: %gdot2.0_eta0.05_a0.3_l0.00125_Lz2.0_temperature1.0e-4_Dt0.002_Nz600_Nin2000_Nout200_adapt1_tmax600.0
883: \label{fig:interface}
884: \vspace{-0.5cm}
885: \end{figure}
886: 
887: We return to comment on the validity of restricting our study to the
888: $y-z$ plane, which was done mainly for computational efficiency. As
889: seen from our results, the ultimate amplitude of the interfacial
890: undulation in this ($y-z$) plane is in fact comparable to that
891: reported in the ($x-y$) plane in
892: Ref.~\cite{fielding-prl-96--2006} (to within $10\%$ at $a=0.3$,
893: $\eta=0.05$, $\gdotbar=2.0$, $\ell=0.00375$, $L_{\{x,z\}}=2$). The
894: present study therefore shows that vorticity structuring is indeed
895: important, but also that a full 3D simulation should be performed in
896: future work.
897: 
898: Finally, we discuss briefly the thickness $d$ of the interface between
899: the bands.  In the 1D initial state, $d=O(\ell)$. In the limit
900: $\ell\to 0$, this gives an unphysically sharp interface $d\to
901: 0$. Associated with this is a pathological steady state that strongly
902: depends strongly on the flow history~\cite{olmsted-jr-44-257-2000}. In
903: 2D, in contrast, $d$ appears virtually independent of $\ell$, as shown
904: in Fig.~\ref{fig:interface}c-f.  This is an important finding that
905: could potentially obviate the gradient term $l^2\nabla^2\tens\Sigma$
906: in Eqn.~\ref{eqn:DJS}, which is needed to give a finite interfacial
907: thickness in 1D~\cite{lu-prl-84-642-2000}.  Nonetheless, the
908: interfacial profile does retain a small bump of thickness $O(\ell)$,
909: Fig.~\ref{fig:interface}e-f, suggesting that the local case $\ell=0$
910: remains pathological even in 2D. Indeed, at $\ell=0$ the stress signal
911: varies erratically about an average that varies between runs,
912: Fig.~\ref{fig:selected}, though purely numerical instability cannot be
913: ruled out. This important issue will be pursued further in future
914: work.
915: 
916: To summarise, we have identified a mechanism by which vorticity stress
917: bands and Taylor-like velocity rolls can form in a complex fluid,
918: triggered by the instability of gradient shear banded flow with
919: respect to interfacial undulations. In any real startup experiment, we
920: would expect the vorticity instability to commence during the final
921: stages of the initial band formation: above we assumed a complete
922: separation of timescales between the processes. In future work, we
923: will study the true dynamics of shear startup experiment in curved
924: Couette flow. We will also extend to 3D, to study the interplay of
925: vorticity banding with the dynamics of
926: Ref.~\cite{fielding-prl-96--2006}. Robustness of the mechanism in
927: other models will also be studied.
928: 
929:  The author thanks Peter Olmsted, Paul Callaghan, Georgina Wilkins and
930:  Helen Wilson for discussions, and the UK's EPSRC for financial
931:  support, GR/S29560/01.
932: 
933:  \begin{thebibliography}{10}
934: 
935: \bibitem{BritCall97c}
936: M.~M. Britton and P.~T. Callaghan, Phys. Rev. Lett. {\bf 78},  4930  (1997).
937: 
938: \bibitem{berret-pre-55-1668-1997}
939: J.~F. Berret, G. Porte, and J.~P. Decruppe, Phys. Rev. E {\bf 55},  1668
940:   (1997).
941: 
942: \bibitem{BecManCol04}
943: L. Becu, S. Manneville, and A. Colin, Phys.\ Rev.\ Lett. {\bf 93},  art. no.
944:   (2004).
945: 
946: \bibitem{diat93}
947: O. Diat, D. Roux, and F. Nallet, J.~Phys.~II (France) {\bf 3},  1427  (1993).
948: 
949: \bibitem{wilkinsOlms2006}
950: G.~M.~H. Wilkins and P.~D. Olmsted, Submitted for publication  (2006).
951: 
952: \bibitem{SalManCol03}
953: J.~B. Salmon, S. Manneville, and A. Colin, Phys.\ Rev.\ E {\bf 68},  art. no.
954:   (2003).
955: 
956: \bibitem{pujolle01}
957: C. Pujolle-Robic and L. Noirez, Nature {\bf 409},  167  (2001).
958: 
959: \bibitem{lettinga-jpm-16-S3929-2004}
960: M.~P. Lettinga and J.~K.~G. Dhont, J. Physics-condensed Matter {\bf 16},  S3929
961:    (2004).
962: 
963: \bibitem{dhont-fd-123-157-2003}
964: J.~K.~G. Dhont, M.~P. Lettinga, Z. Dogic, T.~A.~J. Lenstra, H. Wang, S.
965:   Rathgeber, P. Carletto, L. Willner, H. Frielinghaus, and P. Lindner, Faraday
966:   Discussions {\bf 123},  157  (2003).
967: 
968: \bibitem{berret-prl-8704--2001}
969: J.~F. Berret and Y. Serero, Phys. Rev. Lett. {\bf 8704},    (2001).
970: 
971: \bibitem{CRBMGHJL02}
972: P. Coussot, J.~S. Raynaud, F. Bertrand, P. Moucheront, J.~P. Guilbaud, H.~T.
973:   Huynh, S. Jarny, and D. Lesueur, Phys.\ Rev.\ Lett. {\bf 88},  art. no.
974:   (2002).
975: 
976: \bibitem{HilVla02}
977: L. Hilliou and D. Vlassopoulos, Ind.\ Eng.\ Chem.\ Res. {\bf 41},  6246
978:   (2002).
979: 
980: \bibitem{ChenZAHSBG92}
981: L.~B. Chen, C.~F. Zukoski, B.~J. Ackerson, H.~J.~M. Hanley, G.~C. Straty, J.
982:   Barker, and C.~J. Glinka, Phys. Rev. Lett. {\bf 69},  688  (1992).
983: 
984: \bibitem{Yerushalmi70}
985: J. Yerushalmi, S. Katz, and R. Shinnar, Chemical Engineering Science {\bf 25},
986:   1891  (1970).
987: 
988: \bibitem{lu-prl-84-642-2000}
989: C.~Y.~D. Lu, P.~D. Olmsted, and R.~C. Ball, Phys. Rev. Lett. {\bf 84},  642
990:   (2000).
991: 
992: \bibitem{Bonn+98}
993: D. Bonn, J. Meunier, O. Greffier, A. Alkahwaji, and H. Kellay, Phys. Rev. {\bf
994:   E} {\bf 58},  2115  (1998).
995: 
996: \bibitem{fischer-ra-41-35-2002}
997: P. Fischer, E.~K. Wheeler, and G.~G. Fuller, Rheologica Acta {\bf 41},  35
998:   (2002).
999: 
1000: \bibitem{fischer-ra-39-234-2000}
1001: P. Fischer, Rheologica Acta {\bf 39},  234  (2000).
1002: 
1003: \bibitem{goveas-epje-6-79-2001}
1004: J.~L. Goveas and P.~D. Olmsted, European Phys. J. E {\bf 6},  79  (2001).
1005: 
1006: \bibitem{olmsted-pre-60-4397-1999}
1007: P.~D. Olmsted and C.~Y.~D. Lu, Phys. Rev. E {\bf 60},  4397  (1999).
1008: 
1009: \bibitem{drazin}
1010: P.~G. Drazin and W.~H. Reid, {\em Hydrodynamic stability} (Cambridge University
1011:   Press, Cambridge, 2004).
1012: 
1013: \bibitem{larson-jfm-218-573-1990}
1014: R.~G. Larson, E.~S.~G. Shaqfeh, and S.~J. Muller, J. Fluid Mechanics {\bf 218},
1015:    573  (1990).
1016: 
1017: \bibitem{Lars92c}
1018: R.~G. Larson, Rheol. Acta. {\bf 31},  213  (1992).
1019: 
1020: \bibitem{kang-pre-74--2006}
1021: K.~G. Kang, M.~P. Lettinga, Z. Dogic, and J.~K.~G. Dhont, Phys. Rev. E {\bf
1022:   74},    (2006).
1023: 
1024: \bibitem{fielding-prl-95--2005}
1025: S.~M. Fielding, Phys. Rev. Lett. {\bf 95},    (2005).
1026: 
1027: \bibitem{wilson-jnfm-138-181-2006}
1028: H.~J. Wilson and S.~M. Fielding, J. Non-newtonian Fluid Mechanics {\bf 138},
1029:   181  (2006).
1030: 
1031: \bibitem{lerouge-prl-96--2006}
1032: S. Lerouge, M. Argentina, and J.~P. Decruppe, Phys. Rev. Lett. {\bf 96},
1033:   (2006).
1034: 
1035: \bibitem{decruppe-pre-73--2006}
1036: J.~P. Decruppe, O. Greffier, S. Manneville, and S. Lerouge, Phys. Rev. E {\bf
1037:   73},    (2006).
1038: 
1039: \bibitem{britton-prl-78-4930-1997}
1040: M.~M. Britton and P.~T. Callaghan, Phys. Rev. Lett. {\bf 78},  4930  (1997).
1041: 
1042: \bibitem{cates90}
1043: M.~E. Cates, J.~Phys. Chem. {\bf 94},  371  (1990).
1044: 
1045: \bibitem{johnson-jnfm-2-255-1977}
1046: M.~W. Johnson and D. Segalman, J. Non-newtonian Fluid Mechanics {\bf 2},  255
1047:   (1977).
1048: 
1049: \bibitem{yerushal.j-ces-25-1891-1970}
1050: W. Yerushal.j, S. Katz, and R. Shinnar, Chem. Engineering Science {\bf 25},
1051:   1891  (1970).
1052: 
1053: \bibitem{renardy}
1054: Y.~Y. Renardy, The. Comp. Fl.~Dyn. {\bf 7},  463  (1995).
1055: 
1056: \bibitem{fielding-prl-96--2006}
1057: S.~M. Fielding and P.~D. Olmsted, Phys. Rev. Lett. {\bf 96},    (2006).
1058: 
1059: \bibitem{hinch-jnfm-43-311-1992}
1060: E.~J. Hinch, O.~J. Harris, and J.~M. Rallison, J. Non-newtonian Fluid Mechanics
1061:   {\bf 43},  311  (1992).
1062: 
1063: \bibitem{mcleish87}
1064: T.~C.~B. McLeish, J.~Poly. Sci. B-Poly. Phys. {\bf 25},  2253  (1987).
1065: 
1066: \bibitem{chilcott-jnfm-29-381-1988}
1067: M.~D. Chilcott and J.~M. Rallison, J. Non-newtonian Fluid Mechanics {\bf 29},
1068:   381  (1988).
1069: 
1070: \bibitem{PreTeuVetFla92}
1071: W.~H. Press, S.~A. Teukolsky, W.~T. Vetterling, and B.~P. Flannery, {\em
1072:   Numerical Recipes in C (2nd ed.)} (Cambridge University Press, Cambridge,
1073:   1992).
1074: 
1075: \bibitem{Pozrikidis}
1076: C. Pozrikidis, {\em Introduction to Theoretical and Computation Fluid Dynamics}
1077:   (Oxford University Press, New York, 1997).
1078: 
1079: \bibitem{NAG}
1080: Numerical Algorithms Group Ltd., Wilkinson House, Jordan Hill Road, Oxford, OX2
1081:   8DR, UK.
1082: 
1083: \bibitem{fielding-epje-11-65-2003}
1084: S.~M. Fielding and P.~D. Olmsted, European Phys. J. E {\bf 11},  65  (2003).
1085: 
1086: \bibitem{olmsted-jr-44-257-2000}
1087: P.~D. Olmsted, O. Radulescu, and C.~Y.~D. Lu, J. Rheology {\bf 44},  257
1088:   (2000).
1089: 
1090: \end{thebibliography}
1091: 
1092: 
1093: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1094: %\bibliographystyle{prsty}
1095: %\bibliography{ackerson,actin,articles,banding,barham,Berret,berrportdecruppe,berthier,books,bray,callaghan,cates,chandcits,cook,crystal,crystal_theory,Decruppe,dhont,dnatheory,elasticTurbulence,fielding,fischer,Fischer,flowcryst,fredrickson,gelbart,goveas,graham,Groisman,head,hebraud,helfrich,HinchRallison,hsiao,Kadoma,larson,LCtheory,leal,lerougeDecruppe,lifshitz,line_tension,maffettone,malkus,master,mccoy,membs,Mexican,noirez,notes,olmsted,onions,otherRelated,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,vansaarloos,vorticity,Wang,weitz,wilson,worms2,worms3,worms,yuan,zubarev}
1096: 
1097: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1098: 
1099: \end{document}
1100: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1101: 
1102: